Phylogeography and Conservation Genetics of Two Endangered Amphibians, Blanchard's Cricket Frog (Acris crepitans blanchardi) and the Puerto Rican Crested Toad (Peltophryne lemur)

Kaela B. Beauclerc

A thesis submitted to the Committee on Graduate Studies in partial fulfillment of the requirements for the degree of Doctor of Philosophy in the Faculty of Arts and Science

TRENT UNIVERSITY

Peterborough, Ontario, Canada

© Kaela B. Beauclerc 2009

Environmental and Life Sciences Ph.D. Program

May 2009

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Abstract

Phylogeography and Conservation Genetics of Two Endangered Amphibians, Blanchard's Cricket Frog (Acris crepitans blanchardi) and the Puerto Rican Crested Toad (Peltophryne lemur)

Kaela B. Beauclerc

I investigated the genetic diversity and structure of two endangered amphibians with the

goal of developing conservation recommendations. The Puerto Rican crested toad

(Peltophryne lemur) is a critically endangered tropical bufonid, for which one wild

population remains. Declines are primarily due to habitat modification, and recovery

efforts include captive breeding of northern and southern populations. In contrast,

Blanchard's cricket frog (Acris crepitans blanchardi) is a temperate hylid that is

declining at the northern edge of its range, but for which many large southern populations

exist. The causes of declines are not well understood, and few conservation efforts have

been initiated. I profiled the captive and wild populations of P. lemur, and 185 locations

encompassing all Acris taxa, at the mitochondrial control region and several novel

microsatellite loci. P. lemur had moderate microsatellite allelic diversity, but northern

and southern populations were each fixed for a different mitochondrial haplotype. The

northern population possessed many private microsatellite alleles that, when combined

with habitat differences, may be indicative of adaptive variation. Given the highly inbred

nature of the northern population and its likely extirpation, I recommend that a third

breeding colony be established in which northern and southern individuals are mixed to

preserve any unique northern traits. Profiling of A. c. blanchardi revealed a relatively

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. uniform northern clade and a substructured southern clade. Low genetic diversity in the

north may be contributing to declines. The south was likely a refugium during

Pleistocene glaciations, while the north experienced repeated population expansions and

contractions. Phylogenetic analysis of all Acris species and subspecies indicated that

substantial revisions to current and distributions are required, including

elevation of A. blanchardi to species status. Diversification within this complex

primarily reflects large rivers, changes in elevation, and Pleistocene glaciations.

Recognition of A. blanchardi as a species increases the urgency of conservation actions,

and the distinctiveness of northern populations suggests that they may possess adaptive

traits that should be conserved. Phylogeographic and spatial analysis provide a

framework for actions such as reintroductions. Overall, this research demonstrates the

importance of incorporating genetic information into recovery recommendations and

management actions for threatened species.

Keywords: amphibian, anuran, declining amphibian populations, endangered species,

Acris, Peltophryne lemur, cricket frog, Puerto Rican crested toad, microsatellite,

mitochondrial DNA, control region, phylogeography, genetic structure, conservation

genetics, Pleistocene glaciation

11

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Acknowledgements

Immense thanks go to my advisor, Dr. Brad White, for having the patience to see this through to the end. He provided the right amount of prodding, was always available for discussions, and could be counted on to keep the "bigger picture" in focus. I am also grateful to my committee members, Drs. Michael Berrill and David Galbraith, for providing a fresh perspective and forcing me to think outside the DNA. Bob Johnson at the Toronto Zoo is the inspiration for this work and life in general. His passion for these little frogs and their conservation is contagious. I am indebted to him for providing me with these fascinating projects, and for letting me tag along to Puerto Rico and participate in recovery efforts for the crested toad. This was full of unforgettable experiences, including swimming in the Caribbean Sea, maniac drivers, bomba with some locals, rainforests, tiny toad metamorphs, and roadside roasted chicken. Very special thanks go to Drs. Chris Wilson and Paul Wilson for countless chats about my research and my future. Dr. Tim Frasier cannot be thanked enough for his time explaining the minutiae to me and critiquing some of this work. Linda Cardwell of the ENLS program helped to smooth out the endless bumps encountered along the way. Her enthusiasm is refreshing. Andrew Lentini coordinated samples from the Toronto Zoo and endured many annoying emails from me. The cast and crew of the NRDPFC have made this experience enjoyable and unforgettable. Many technicians were instrumental in generating the data and I am sincerely grateful to all of them: Sue White-Sobey, Karen Mills, Kristyne Wozney, Smolly Coulson, and Jen Dart have variously sequenced, genotyped, and picogreened samples. Blane Bell is the automation wiz that maintained my sanity by extracting almost one thousand samples. Very special thanks go to my cohort of grad students, particularly Drs. Cathy Cullingham, Tim Frasier, Brenna McLeod, Roxanne Gillet, and Mark Ball for assistance, sympathy, rides to school, and life outside the lab. Dr. Stephen Petersen and his Mac were always ready to run ModelTest for me.

111

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Ariana Rickard and Edi Sonntag are thanked for dragging me along during their field work in Michigan so I could have some perspective on cricket frogs in real life. I am indebted to the following people for providing tissue samples of Acris: Joe Bartoszek, Neil Bass, Neil Bernstein, Will Bird, Luis Bonachea, Ron Bonnett, Jeff Briggler, Liz Burton, Brian Butterfield, Jeffrey Camper, Keith Coleman, Jeffrey Davis, Kevin Enge, Amy Estep, Eugenia Farrar, Dan Fogell, Carl Franklin, David Garcia, Dave Golden, Wayne Hildebrand, John Himes, John Jensen, Glen Johnson, Bob Jones, John Kleopfer, Kenny Krysko, Mike Lannoo, Jim Lee, Lisa Lehnhoff, Richard Lehtinen, Gregory Lipps, Mike Lodato, Malcolm McCallum, Brian Metts, Jonathan Micancin, Richard Montanucci, Nate Nazdrowicz, Jeff Parmelee, Steve Perrill, John Petzing, Stephen Richter, Ariana Rickard, Leslie Rissler, Floyd Scott, Don Shepard, Dirk Stevenson, David Swanson, Ashley Traut, Stan Trauth, and Zack Walker. Special thanks go to Donna Dittman at Louisiana State University Museum of Natural Science and Travis Taggart at the Sternberg Museum of Natural History, Fort Hays State University for arranging substantial tissue loans of Acris. Many thanks go to Miguel Canals of the Guanica State Forest and Rick Dietz of the Audubon Nature Institute for providing tissue samples of the Puerto Rican crested toad. Dr. Juan Rivero graciously allowed me to sample the historic specimens in his private collection, Jaime Matos-Torres provided additional specimens, and Gail Ross tromped around Puerto Rico collecting historic tissue samples, but these unfortunately were not incorporated. My family deserves special recognition for enduring me throughout my seemingly infinite university education. My parents, Murray and Marianne Jacobs, have encouraged and supported my education goals and love of biology and animals since I first protected that acre of rainforest in grade 9. Parker made sure I knew there was a world outside of my laptop and got some fresh air and exercise twice a day. My husband Michael encouraged, supported, listened, laughed, and vacuumed the house every week. He has seen me at my worst throughout this, and stuck it out anyways. He is my emotional stability and my best friend. Thank you for believing in me.

IV

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Table of Contents

Abstract ...... ii

Acknowledgements ...... iv

List ofFigures ...... vii

List of Tables ...... viii

Chapter 1 General Introduction ...... 1

Declining amphibian populations ...... 1 Conservation genetics ...... 3 Biology and conservation of Peltophryne lemur ...... 7 Biology and conservation of A eris crepitans blanchardi ...... 14 Thesis objectives and overview ...... 22

Chapter 2 Genetic Rescue of an Inbred Captive Population of the Critically Endangered Puerto Rican Crested Toad (Peltophryne lemur) by Mixing Lineages ...... 27

Abstract ...... 28 Introduction ...... 29 Materials and Methods ...... 34 Results ...... 39 Discussion ...... 46

Chapter 3 Characterisation, Multiplex Conditions, and Cross-Species Utility of Tetranucleotide Microsatellite Loci for Blanchard's Cricket Frog (Acris crepitans blanchardi) ...... 55

Abstract ...... 56 Introduction ...... 57 Materials and Methods ...... 57 Results and Discussion ...... 63

V

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Chapter 4 Phylogeography and Spatial Structure of Blanchard's Cricket Frog (Acris crepitans blanchardi): Conservation Issues at the Northern Edge of its Range ...... 68

Abstract ...... 69 Introduction ...... 70 Materials and Methods ...... 74 Results ...... 79 Discussion ...... 90

Chapter 5 Molecular Phylogenetics of the North American Cricket Frog Complex (Genus Acris) Across its Range ...... 99

Abstract ...... 100 Introduction ...... 101 Materials and Methods ...... 108 Results ...... 116 Discussion ...... 131

Chapter 6 General Discussion ...... 157

Conservation genetics of Peltophryne lemur ...... 157 Synthesis of Acris data and conservation genetics of A. blanchardi ...... 158 Value of genetics to conservation biology ...... 164 Conclusion ...... 166

Literature Cited ...... 167

Appendices ...... 199

Vl

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. List of Figures

Figure 1.1. Male and female P. lemur ...... 9

Figure 1.2. Large costume character of P. lemur used for public education ...... 13

Figure 1.3. Currently accepted species and subspecies of A eris ...... 15

Figure 2.1. Map of Puerto Rico and the British Virgin Islands with extirpated and extant populations of P. lemur ...... 31

Figure 2.2. Microsatellite allele frequencies for three P. lemur populations ...... 42

Figure 4.1. Location of sampling sites for A. c. blanchardi ...... 71

Figure 4.2. Linear regression of nucleotide diversity with latitude and longitude for A. c. blanchardi ...... 81

Figure 4.3. Median joining network of A. c. blanchardi haplotypes ...... 82

Figure 4.4. FcT and Fsc for groupings of A. c. blanchardi sites using spatial analysis of molecular variance ...... 83

Figure 4.5. Correlograms of genetic and geographic distance for A. c. blanchardi ...... 86

Figure 4.6. Pairwise mismatch distributions for A. c. blanchardi groups ...... 89

Figure 5.1. Distribution of currently accepted A eris taxa ...... 104

Figure 5.2. Distribution of sample collection sites and major clades of Acris ...... 109

Figure 5.3. Unrooted Bayesian phylogenetic tree for Acris haplotypes ...... 118

Figure 5.4. Bayesian phylogenetic tree of A. c. blanchardi haplotypes ...... 126

Figure 5.5. Bayesian phylogenetic tree of A. gryllus haplotypes ...... 129

Figure 5.6. Geographic boundaries for Acris identified using BARRIER ...... 130

Figure 5.7. The Mississippi Delta showing localities containing A. blanchardi and A. gryllus specimens ...... 138

Figure 5.8. Major physiographic divisions of the United States ...... 140

Figure 5.9. Geographic distribution of allozymes among A eris populations ...... 150

Vll

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. List of Tables

Table 2.1. Primers and amplification conditions for six microsatellite loci isolated from P. lemur ...... 38

Table 2.2. Characteristics of six microsatellite loci in three populations of P. lemur ...... 41

Table 2.3. Diversity indices for six microsatellite loci in three populations of P. lemur ...... 43

Table 2.4. Genetic distance measures using six microsatellite loci for three populations of P. lemur ...... 45

Table 3.1. Characteristics of tetranucleotide microsatellite loci and multiplex amplification conditions for A. c. blanchardi ...... 59

Table 3.2. Genetic diversity indices for 16 tetranucleotide microsatellite loci in two populations of A. c. blanchardi ...... 62

Table 3.3. Genetic distance measures for 16 microsatellite loci in two populations of A. c. blanchardi ...... 64

Table 3.4. Cross-amplification ability of A c. blanchardi microsatellite loci in related A eris subspecies and species ...... 67

Table 4.1. Summary of sampling sites for A. c. blanchardi ...... 75

Table 4.2. Characteristics of seven A. c. blanchardi groups identified using spatial analysis of molecular variance ...... 85

Table 4.3. Neutrality tests and parameters of the mismatch distribution for A. c. blanchardi groups ...... 87

Table 5.1. Summary of samples collected for five Acris subspecies ...... 111

Table 5.2. Characteristics of three Acris phylogenetic clades ...... 119

Table 5.3. Sequence divergences between various Acris lineages ...... 120

Table 5.4. Analysis of molecular variance for various Acris groupings ...... 122

Table 5.5. Pairwise sT for currently recognised taxa and phylogenetic clades of Acris ...... 123

Vlll

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 1 -

Chapter 1

General Introduction

Declining amphibian populations

Declining amphibian populations have received widespread attention since the

First World Congress of Herpetology in 1989, when researchers shared anecdotes of

disappearing amphibian populations around the globe (Barinaga 1990; Phillips 1990;

Wyman 1990). Initial skepticism suggested that these declines may simply be artifacts of

the short duration of most studies, during which natural population fluctuations were

mistaken to be catastrophic declines or extinctions (Pechmann et al. 1991; Pechmann and

Wilbur 1994; Alford and Richards 1999; Skelly et al. 2003). For example, many

amphibian species are thought to exist as metapopulations, in which local populations are

periodically extirpated but recolonised several years later (Blaustein et al. 1994; Hanski

1998; Alford and Richards 1999; Marsh and Trenham 2001). Some species will also

forego breeding in years with unfavourable climate; since most amphibian surveys focus

on reproductive habitat, researchers may wrongly conclude that a population has become

extirpated in such instances (Crump et al. 1992; Alford and Richards 1999; Skelly et al.

2003). However, numerous studies have since established the reality of global amphibian

declines. Models have suggested that the combination of events at some locations is

highly unlikely to have occurred by chance alone (Pounds et al. 1997, 1999; Alford and

Richards 1999; Houlahan et al. 2000). Continued monitoring has substantiated the

extinction of some species (Pounds et al. 1999; Meyer et al. 2004; Pounds and Savage

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 2 -

2004), while others have documented the decline of healthy populations in real time (Lips

et al. 2006). A recent survey of the IUCN Global Amphibian Assessment found that

more than 30% of the world's amphibians are globally threatened, far more than either

birds or (Stuart et al. 2004). We may be in the midst of a sixth mass extinction

(Wake and Vredenburg 2008), with the disappearance of amphibians possibly 45 000

times greater than the background rate of extinction (Mccallum 2007).

Numerous factors are considered to be involved in the decline of amphibian

populations around the world. Disappearances of many populations can be attributed to

"obvious" causes for which there is a good understanding of the underlying ecological

mechanisms; these include habitat alteration and fragmentation, introduced species that

prey upon and/or compete with native species, and over-exploitation for human

consumption, medicinal use, and the pet trade ("class I" hypotheses; Collins and Storfer

2003). However, many populations, primarily in Neotropical montane regions and

Australia, have declined despite their existence in protected and seemingly intact habitats

(Stuart et al. 2004). The underlying mechanisms of these "enigmatic" declines are poorly

understood, but likely involve global change ( encompassing climate and UV radiation),

chemicals, and emerging infectious diseases ("class II" hypotheses; Collins and Storfer

2003). Interaction among multiple stressors can compound the effects of these individual

factors and exacerbate declines of amphibian populations. For example, climate-induced

reductions in water depth resulted in increased exposure to UV-Band subsequent greater

infection by the fungus Saprolegniaferax in western toads (Bufo boreas) in the northwest

USA (Kiesecker et al. 2001). Similarly, climate warming has resulted in reduced mist

frequency in Neotropical,montane regions (Pounds et al. 1999, 2006), which has been

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 3 -

linked to outbreaks of the emergent pathogenic fungus Batrachochytrium dendrobatidis

that resulted in rapid, massive die-offs of aquatic amphibians in this area (Pounds and

Crump 1994; Burrowes et al. 2004; Lips et al. 2006; but see Rohr et al. 2008). Species

with restricted geographic ranges may be at greatest risk of extinction, as they typically

occupy highly specialised niches and are more vulnerable to stochastic events (Sodhi et

al. 2008). Pond-breeding anurans may be most affected by habitat fragmentation, as they

require more frequent dispersal due to large population fluctuations in an unstable habitat

(Green 2003). Larger body size, which is associated with lower reproductive rate and

reduced recovery potential, and pronounced seasonality, which increases susceptibility to

changes in climate, may also increase a species' extinction risk (Sodhi et al. 2008).

Conservation genetics

Addressing the environmental ( e.g. habitat destruction, contaminants, road

mortality) and demographic (e.g. stochastic variation in recruitment or sex structure)

issues that contribute to population declines is critical for the effective recovery of an

endangered species. However, it is the genetic variation within populations that will

determine their ability to adapt to changing environments, and thus their long-term

viability (Mace et al. 1996). The role of genetics in conservation has been recognised for

over 20 years (A vise 1989), but the field of "conservation genetics" has only emerged

during the past decade. A number of initiatives, including the journal Conservation

Genetics in 2000 and an introductory textbook (Frank.ham et al. 2002), have enabled this

field to flourish and become an integral component to the recovery of endangered

species. The primary genetic issues that should be considered within a conservation

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 4 -

framework include determining levels of genetic diversity, structure, and gene flow

among populations; resolving taxonomic uncertainties and defining management units;

evaluating the deleterious consequences of small population size and isolation; and the

application of management tools such as ex situ breeding programs and reintroductions.

Determining the geographic patterns of genetic diversity, structure, and gene flow

among populations, or phylogeography, is fundamental to understanding the biology and

evolutionary history of a species (A vise 2000). However, it also provides a framework

for developing appropriate and practical management strategies for endangered species.

For example, a phylogeographic study can determine whether isolated populations have

experienced recent gene flow, or if they represent independently evolving lineages

(James and Moritz 2000; Carpenter et al. 2001; Eizirik et al. 2001; Burns et al. 2007;

Amato et al. 2008). It may identify populations that are important to conserve because

they are reservoirs of genetic variation (Rowe et al. 1998; Tarr and Fleischer 1999), or

those with limited variation and thus reduced evolutionary potential that are at greatest

risk of extirpation (Bos and Sites 2001; Vianna et al. 2006). One important outcome of a

broad-scale phylogeographic study is the evaluation of existing taxonomy and the

identification of distinct population units. Phenotypic traits can be highly polymorphic

and may be influenced by environmental conditions; consequently, traditional taxonomic

assignments based on morphology may not precisely reflect patterns of genetic variation

(Burbrink et al. 2000; Culver et al. 2000; Nittinger et al. 2007; but see Avise and Walker

1999). This can waste available resources on groups that are not genetically distinct and

therefore have limited conservation value to the species (Avise and Nelson 1989; Zink et

al. 2000). Alternatively, a phylogeographic study may detect the presence of distinct

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 5 -

lineages that warrant separate management as species, subspecies, or evolutionary

significant units (Moritz 1999; Fraser and Bernatchez 2001 ). Early recognition of such

lineages is imperative, as they may be at greater extinction risk due to rarity (Daugherty

et al. 1990; Wang et al. 1999, 2008; Nielson et al. 2001; Vredenburg et al. 2007).

Small, isolated populations are known to be more vulnerable to stochastic

environmental and demographic events, such as hurricanes, droughts, low recruitment, or

skewed sex ratios. However, they also commonly suffer from genetic impoverishment.

The genetic bottleneck that may accompany a population crash can significantly erode

genetic variation, measured as heterozygosity and number of alleles, as only a portion of

the population's original genetic composition survives (Hoelzel et al. 1993; Whitehouse

and Harley 2001; Frankham et al. 2002; Larson et al. 2002). Genetic drift also has a

greater impact on rates of fixation in small populations, and can cause further loss of

diversity and increase the genetic distance between isolated populations as they become

fixed for different alleles ( Goodman et al. 2001; Hedrick 2001; Wilson et al. 2008). The

low diversity that results from genetic drift and inbreeding in small populations may be

associated with physical abnormalities and reduced fitness (Patenaude et al. 1994;

Westemeier et al. 1998; Rowe et al. 1999; Hedrick and Kalinowski 2000; Halverson et al.

2006; Siddle et al. 2007). The reduced genetic variation that often characterises small

populations may limit its ability to respond to changing environments, thereby increasing

its chance of extinction (Mehlman 1997; Bridges and Semlitsch 2001; Frankham et al.

2002). Populations at the periphery of a species' range are typically small and isolated,

experiencing reduced gene flow, increased genetic drift, and greater levels of inbreeding

(Lesica and Allendorf 1995). This may leave these populations genetically depauperate

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 6 -

and thus more susceptible to extirpation from stochastic events (Descimon and

Napolitano 1993; Rowe et al. 1999; Hutchison 2003; but see Lomolino and Channell

1995; Eckert et al. 2008). Conversely, peripheral populations may be a source for

speciation, as isolation and genetic drift facilitate divergence from core populations in

response to local selection pressures (Lesica and Allendorf 1995; Garcia-Ramos and

Kirkpatrick 1997). Furthermore, the highly variable conditions to which they are

subjected may enable them to maintain unique genetic diversity and evolve greater

resistance to environmental extremes and future changes (Volis et al. 1998; Eckstein et

al. 2006; Bohme et al. 2007). For these reasons, peripheral populations, even of

relatively widespread species, may be valuable and worthy of conservation efforts

(Hunter and Hutchinson 1994; Lesica and Allendorf 1995).

A thorough understanding of genetic diversity and structure of a species is crucial

to implementing efficient and optimal recovery activities, such as identifying populations

for use in translocation or reintroduction efforts (Shaffer et al. 2000, 2004a; Godoy et al.

2004; Wilson et al. 2008). Source populations should be genetically similar to the

recipient population to reduce outbreeding depression, increase survival potential, and

maintain evolutionary trajectories (Drew et al. 2003; Kraaijeveld-Smit et al. 2005), but

also genetically diverse to maximise variation, minimise swamping of native individuals,

and reduce current and future inbreeding depression (Madsen et al. 1999; Friar et al.

2000; Frankham et al. 2002; Land et al. 2004). A genetic survey can be critical to these

efforts, as geographically proximal populations are not necessarily the most genetically or

adaptively similar (Chek et al. 2001; Trapnell et al. 2007).

The effectiveness of reintroductions as a means of facilitating the recovery of

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 7 -

endangered amphibian populations is controversial (Burke 1991; Dodd and Seigel 1991;

Bloxam and Tonge 1995; Marsh and Trenham 2001; Trenham and Marsh 2002; Seigel

and Dodd 2002; Germano and Bishop 2009). Although many reintroduction attempts

have failed and the outcome of others remains unclear (Dodd and Seigel 1991 ), an

extensive recovery program involving habitat restoration, genetic analyses, and

reintroductions has successfully established at least six new populations of the natterjack

toad (Bufo calamita) in formerly occupied areas in Britain (Denton et al. 1997).

Furthermore, a recent review of amphibian translocations (Germano and Bishop 2009)

found a much greater success rate than was reported nearly 20 years ago (Dodd and

Seigel 1991). Most reintroduction programs for which long term data were available did

not incorporate genetic data into their design (Dodd and Seigel 1991); it is thus possible

that the failure of many of these programs was due in part to the lack of genetic

information prior to their initiation. Animals with limited dispersal capabilities, such as

amphibians, are particularly susceptible to increased genetic divergence between

populations, and substructuring may occur at very fine scales such as among individual

ponds (Shaffer et al. 2000). This suggests that the genetic composition of amphibian

populations is not trivial to the design of a recovery plan, and must be considered fully to

maximise the potential for successful restoration of endangered species.

Biology and conservation of Peltophryne lemur

The Puerto Rican crested toad (Peltophryne lemur) is the only toad endemic to

Puerto Rico. Peltophryne is a monophyletic genus consisting of ten species of bufonid

toads endemic to the West Indies (Schwartz and Henderson 1991 ), and has had a

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 8 -

tumultuous taxonomic history. The genus Peltophryne was first proposed by Fitzinger in

1843, but was later incorporated into Bufo (Gunther 1858; Stejneger 1902). The species

Peltaphryne lemur was first described, and misspelled (Smith and Goebel 1994), by Cope

(1868). In 1917 Barbour identified Bufo turpis on Virgin Gorda, which was later

considered to be Bufo lemur (Schwartz and Thomas 1975). Pregill (1981a) demonstrated

that all endemic West Indian bufonids formed a natural, monophyletic group distinct

from other members of Bufo based on the shared, unique feature of co-ossification of the

skull. He thus suggested that generic recognition of the West Indian bufonids was

appropriate, and resurrected the name Peltophryne. Further morphological and molecular

data also supported the monophyly of West Indian toads (Pramuk et al. 2001; Pramuk

2002). However, immunological distances (Hedges et al. 1992; Hass et al. 2001) and

mitochondrial DNA sequences (Graybeal 1997) showed that the genus Peltophryne was

nested within a paraphyletic Bufo; as such, Hedges (1996) stated that Peltophryne should

again be synonimised with Bufo. Most recently, Frost et al. (2006) proposed that the

large genus Bufo be partitioned into a number of distinct genera, including Peltophryne.

As this is the most recent treatment of West Indian toad taxonomy, I will refer to the

species group as Peltophryne.

Peltophryne lemur is a medium-sized toad, with a snout-vent length ranging from

72-120 mm (Rivero et al. 1980). It has a long, upturned snout, golden marbled eyes, and

prominent supraorbital crests (Figure 1.1) (Pregill 1981a). The toad's local name, sapo

concho, is derived from this latter feature and means shelled or crested toad. There is

conspicuous sexual dimorphism: breeding males are bright yellow, while females are

light to dark grey in colour, considerably larger, possess larger crests, and are wartier

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 9 -

Figure 1.1. Male (upper) and female P. lemur. Photo © Paddy Ryan/Ryan Photographic. Used with permission.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 10 -

(Figure 1.1) (Rivero et al. 1980). P. lemur has been collected from at least eight locations

across Puerto Rico and one on Virgin Gorda of the British Virgin Islands (Figure 2.1 ). It

generally occupies low-elevation(< 200 m), arid or semi-arid coastal regions and is often

associated with limestone karst formations (Schmidt 1928; Rivero et al. 1980; USFWS

1992). This rocky substrate provides abundant fissures and cavities into which it retreats

during the day, emerging late at night to feed and breed (Miller 1985). This may provide

an adaptive explanation for the co-ossification of their skulls: individuals have been

observed using their heads to block the entrance to their burrows (Rivero et al. 1980),

which may prevent desiccation (Seibert et al. 1974; Pregill 1981a). Breeding is sporadic

and occurs in temporary pools formed by heavy rains, with individuals likely breeding

only once per year (USFWS 1992). There appears to be high fidelity to breeding sites,

with individuals travelling up to four kilometres to reach a particular site (Miller 1985).

Although fossil specimens from caves suggest that P. lemur was once abundant

throughout Puerto Rico (Pregill 1981 b ), it has been scarce since the initiation of specimen

collection for natural history museums (Stejneger 1902). It has not been seen on Virgin

Gorda for over 30 years (USFWS 1992), and was thought to be extinct from 1931-1965

(Rivero 1978). In 1966, several individuals were found near the town of Isabela in

northwest Puerto Rico (Garcia-Diaz 1967), and another population was found in nearby

Quebradillas in 1974 (Rivero et al. 1980) (Figure 2.1). It is likely that the semi-fossorial

nature of P. lemur, combined with poor collecting techniques, contributed to this species'

apparent scarcity (Pregill 1981 b ). In 1984, a large breeding population of 1000-2000

toads was found in southwest Puerto Rico at the Guanica State Forest Reserve (Moreno

1985). Currently, it is believed that the Guanica population is the only remaining wild

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 11 -

population, as no toads have been seen in the north since 1988 (Barber 2007).

Peltophryne lemur was granted SSP (Species Survival Plan) designation by the

Association of Zoos and Aquariums in 1984, the first amphibian to be included in the

program (Maruska 1986). It was listed as threatened under the US Endangered Species

Act in 1987. The primary cause of decline is habitat loss and alteration from

anthropogenic factors. Breeding sites have been drained or filled in for construction,

agriculture, and mosquito control (USFWS 1992). The main breeding pond in Guanica is

a beach parking lot that was frequently drained for access, and the area around the

troughs where P. lemur was found breeding in Quebradillas is regularly sprayed with

pesticides (Johnson 1990). Predation by , cats, and introduced mongooses

(Herpestes auropunctatus) may also be a factor (USFWS 1992). The cane toad (Bufo

marinus) was introduced to Puerto Rico in the 1920s to assist in the control of sugar cane

grubs and has since become firmly established across the island (Rivero et al. 1980).

Because P. lemur was rare prior to the introduction of B. marinus, it cannot be solely

blamed for the species' current scarcity. However, predation by the larger B. marinus has

been known to occur in Guanica, and may be directed towards the much smaller males of

P. lemur (Rivero et al. 1980; USFWS 1992). Competition is likely also a factor: B.

marinus was twenty times more abundant at one site, and its earlier metamorphosis

enables it to exclude P. lemur toadlets from refuges, resulting in their death from

exposure (USFWS 1992). The small size and isolation of the only remaining population

in Guanica means that it is also vulnerable to stochastic environmental and demographic

events. For example, many breeding toads were washed out to sea when a pond at the

Guanica State Forest Reserve was inundated during a hurricane in 1985 (Johnson 1990).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 12 -

Recovery efforts for P. lemur have been extensive in Puerto Rico. The Guanica

breeding pond is closed during the three-week breeding season and is no longer drained

for parking (Johnson 1990). At Guanica National Forest, toads have been radiotracked to

monitor movement, the chemistry of the breeding ponds has been analysed, and habitat

preferences of toads have been evaluated (Johnson 1990; Matos-Torres 2006). Four

individuals from Quebradillas were captured and used to establish a captive· breeding

colony in 1980. In 1985, 20 individuals from the newly discovered Guanica population

were used to establish a separate breeding colony representing the southern lineage, with

12 additional individuals later incorporated. These captive populations are maintained

separately because preliminary protein electrophoresis data (Lacy and Foster 1987),

combined with the large geographic distance that separates them, suggest that they may

be divergent. To date, more than 110 000 tadpoles have been released into constructed

ponds in Guanica Forest (Barber 2007), and in 2003 and 2005, captive-born adult toads

returned to breed (PRCT SSP 2006). Before de-listing will be considered, at least three

disjunct, self-sustaining populations of 1500-2000 individuals must be maintained on

each of the northern and southern coasts of the island for 10 years (USFWS 1992). This

will limit the erosion of genetic variation and provide a source for recolonisation in the

event of extirpation of one population. To meet these requirements, several additional

ponds have been constructed in Guanica and Gabia in the south (Barber 2007). In the

north, efforts are underway to secure property in Quebradillas and Cambalache National

Forest, and in 2006 tadpoles were released at three newly constructed ponds in Arecibo

(Barber 2007). Due to the scarcity of P. lemur, Puerto Ricans have come to call the

introduced B. marinus by P. lemur's common name, sapo concho (Grant 1931; Rivero

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 13 -

Figure 1.2. Large costume character of P. lemur developed for public education in Puerto Rico. Photo taken at the Juan Rivero Zoo in Mayaguez with a group of school children. Photo© 2006 Eduardo Valdez. Used with permission.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 14 -

1978; pers. obs.). An island-wide program has since been developed to educate citizens

and instill a sense of pride and stewardship for their endemic toad. This has involved the

distribution of posters, bumper stickers, activity books for children, and the creation of

life-size models and a large costume character (Barber 2007) (Figure 1.2).

Biology and conservation ofAcris crepitans blanchardi

Acris are nonclimbing, semi-aquatic members of the treefrog family Hylidae.

This genus contains two species, northern (A. crepitans) and southern (A. gryllus) cricket

frogs, which are further subdivided into five currently recognised subspecies:

Blanchard's (A. c. blanchardi), northern (A. c. crepitans), coastal (A. c. paludicola),

southern (A. g. gryllus), and Florida (A. g. dorsalis) cricket frogs (Conant and Collins

1998) (Figure 1.3). A. c. blanchardi is the most western and northern form of the

complex, ranging throughout the south and central Great Plains and Midwest (Conant and

Collins 1998) (Figure 5.1). This small (1.6 to 3.8 cm SVL) anuranprimarily occupies

level banks or floating vegetation of slow-moving, permanent water bodies with low

canopy cover (Burkett 1984; Smith et al. 2003; Beasley et al. 2005; Lehtinen and Skinner

2006). Although it seldom ventures far from the water (Pyburn 1958; Blair 1961), it has

been found to disperse up to 1.3 km between ponds (Gray 1983). Calling may begin in

early February in southern areas (Pyburn 1958; Blair 1961) or April/May in northern

states, and lasts until June or July (Johnson and Christiansen 1976; Burkett 1984). Newly

metamorphosed frogs appear within five to ten weeks and are very abundant (Pyburn

1961; Burkett 1984 ). Typically, a single breeding event occurs and consists of one clutch

with up to 400 eggs (Gray et al. 2005). However, females may lay multiple clutches in

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 15 -

Figure 1.3. Currently accepted species and subspecies of Acris. Top left and right: A. c. blanchardi from Michigan(© 2005 Ariana Rickard); middle left: A. c. crepitans from Tennessee(© 2006 Liz Burton); middle right: A. c. paludicola (© Gary Nafis); bottom left: A. g. gryllus from South Carolina(© John Willson); bottom right: A. g. dorsalis from Florida(© 2006 James Harding). All photos used with permission.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 16 -

the south (Trauth et al. 2004), and a second breeding event by the young-of-the-year has

been observed in Texas (Pyburn 1961; Bayless 1969). Only about 5% of newly emerged

juveniles usually survive to the next breeding season, although evidence indicates that

much of the mortality occurs prior to overwintering (Gray 1983; Burkett 1984; but see

Irwin 2005). Activity declines towards the winter, and in some areas juveniles

overwinter in crayfish burrows and cracks in the pond bank (Gray 1971; Irwin et al.

1999). Adult life expectancy is four months, and most of the previous year's adults have

disappeared by October (Gray and Brown 2005). This results in complete population

turnover every 16 months (Burkett 1984; McCallum and Trauth 2004).

Acris c. blanchardi was once considered to be the most common amphibian in

many parts of the upper Midwest (Lannoo et al. 1994; Hay 1998; Mierzwa 1998; Gray

and Brown 2005f Although it continues to be abundant throughout much of the central

and southern Great Plains, populations have been in decline along the northern and

western edge of its range. It has disappeared from several historical sites in northern

Illinois (Mierzwa 1998) and Indiana (Minton 1998), Michigan (Lehtinen 2002), Iowa

(Lannoo et al. 1994; Hemesath 1998), Wisconsin (Mossman et al. 1998), Nebraska

(McLeod 2005), and eastern Ohio (Lehtinen and Skinner 2006). It is likely extirpated

from Minnesota (Moriarty 1998), Colorado (Hammerson and Livo 1999), and Ontario

(Kellar et al. 1997). A. c. blanchardi is now considered to be the "species of most

concern in the Midwest" (Lannoo 1998a, p. 430), and is listed as endangered or

threatened in several areas (e.g. Ontario, Kellar et al. 1997; Michigan, Eagle et al. 2005;

Wisconsin, WI DNR 2005). Declines in this subspecies are different from those of most

amphibians because it is relatively widespread and common in many areas. It has also

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 17 -

been considered somewhat of a habitat generalist (but see below), and is frequently found

in man-made habitats such as ditches, quarries, cattle tanks, and aquaculture ponds

(Bragg 1940; Lehtinen 2002; Lehtinen and Skinner 2006). Perhaps most disconcerting

about these declines is their lack of definitive causes. While habitat loss has likely

contributed, it does not appear sufficient to account for the magnitude and distribution of

the declines (Lannoo 1998b; Knutson et al. 2000). The conversion of the Midwest to

agriculture was complete by the early twentieth century, well before the reported declines

(Gray and Brown 2005). In fact, potential habitat has actually increased in many areas

due to the construction of rural ponds (Leja 1998). Habitat alteration, however, may have

greatly contributed to declines in some areas. The use of agricultural chemicals became

widespread after the 1950s, coinciding with the onset of A. c. blanchardi declines in the

Midwest (Knutson et al. 2000). The application of the herbicide copper sulfate extirpated

this amphibian from one pond in Illinois (Gray and Brown 2005), while levels of organic

pollutants, including PCBs and DDT, were found to be very high in frogs from the now­

extirpated population on Pelee Island, Ontario (Campbell 1978). Intersex gonads and

masculinisation occur in high frequency where exposure to PCBs and DDT is greatest,

and likely contributed to the disappearance of this subspecies from northeastern Illinois

(Reeder et al. 2005). Other industrial contaminants such as mercury (Diana and Beasley

1998) and perchlorate (Theodorakis et al. 2006) may also negatively affect individuals.

Tadpoles of the related subspecies A. c. crepitans had the lowest tolerance to acidified

water among eleven species (Gosner and Black 1957), and were significantly less

abundant in acidified macrocosms (Sparling et al. 1995). However, numerous studies

have found little association between declines and such environmental issues. Organic

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 18 -

pollutants were below toxic levels in several Ohio (Russell et al. 2002) and Illinois

(Beasley et al. 2005) ponds where declines occurred, and high concentrations of the

pesticide chlorpyrifos had little effect on tadpoles (Widder and Bidwell 2008). Some A.

c. crepitans individuals were found to develop a tolerance to lethal concentrations of

DDT (Boyd et al. 1963), demonstrating the potential for adaptation. Furthermore,

Lehtinen and Skinner (2006) found no effect of pH or acid neutralising capacity of ponds

on the distribution of extant and extirpated populations of A. c. blanchardi in Ohio, and it

was actually significantly more abundant in agriculturally-intensive parts of the state.

A variety of other factors have been implicated in declines of northern A. c.

blanchardi populations. Bullfrogs [Rana (Lithobates) catesbeiana] have been observed

to prey heavily on this small frog (Burkett 1984), and introductions significantly affected

amphibian populations in Iowa (Gray and Brown 2005). However, Beasley et al. (2005)

found that reproductive success in A. c. blanchardi was only marginally affected by the

number of bullfrogs. Morphological deformities within this anuran are generally less

than the baseline rate of 2% documented for amphibians (Gray 2000; Ouellet 2000;

Rickard and Sonntag 2006), although rates as high as 14.6% have been observed from

some localities and may be on the rise (McCallum and Trauth 2003; but see Gray 2000).

Some evidence also suggests that the incidence of undetected skeletal abnormalities may

be relatively high (Maglia et al. 2007). This may not be a contributing factor in Illinois,

however, as several malformed individuals were captured multiple times, and at least two

survived through the winter to the following breeding season (Gray 2000). Individuals

have been found that carry trematode (McCallum and Trauth, unpublished) and

protozoan (Jirku et al. 2006) parasites, but these studies did not report on pathological

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 19 -

symptoms. Furthermore, parasites are a common affliction of wild amphibians and are

rarely associated with clinical disease (Crawshaw 2000), although larval renal trematodes

appeared to influence survival of individuals in Illinois (Beasley et al. 2005). Chytrid

fungus (Batrachochytrium dendrobatidis) was detected in up to 15% of A. c. blanchardi

individuals across the Midwest and Great Plains, although no geographic trend was

apparent and both populations and individuals seemed healthy overall (Steiner and

Lehtinen 2008). Finally, the eggs of this anuran may be significantly affected by UV

light, perhaps because they are unprotected by large egg masses, egg sheaths, or

pigments, and are more exposed to solar radiation than some frogs as they are laid closer

to the water surface and later in the year (Van Gorp 2001 ).

Clearly, investigation into the cause of A. c. blanchardi declines has produced

ambiguous or conflicting results. It is probable that various factors have had a greater

impact on populations at different times and in different locations (Gray and Brown

2005). It is also likely that, in addition to the direct effect of each factor, they are acting

indirectly and/or synergistically to cause declines. For example, ponds in Illinois treated

with herbicides had the lowest numbers of metamorphosing A. c. blanchardi individuals

(Beasley et al. 2005). Rather than direct kills from chemicals, however, this appeared to

be caused by damage to that reduced food, cover, and dissolved oxygen for

tadpoles, while juvenile frogs were heavily infested by renal parasites and showed signs

of stress. Modelling of A. c. blanchardi populations in Illinois found that biocides and

predation were the most likely factors to cause extirpation (Veldman 1997).

Immunologically-challenged male A. c. blanchardi displayed reduced investment in

reproduction, suggesting that pathogens or toxins may diminish male fertility and

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 20 -

contribute to population declines (McCallum and Trauth 2007).

Global climate change has likely resulted in various modifications to the habitat

of A. c. blanchardi that may not be immediately apparent as potential contributors to

declines. Lannoo (1998b) describes a combination of three factors as the likely cause of

declines in the Prairie Pothole region of Iowa. Prolonged droughts, which occur on a

roughly ten-year cycle and may increase in severity and/or frequency with climate

change, can eliminate many preferred breeding sites (seasonal and semi-permanent

wetlands) for several consecutive years. The short lifespan of this anuran means that

adults must breed each year to maintain the population,· and therefore must utilise

permanent wetlands during droughts. However, many such sites have been rendered

unsuitable for amphibian breeding by aquaculture, agriculture, and fish and bullfrog

introductions. As well, movement to alternate sites or recolonisation of extirpated sites is

likely limited during droughts (Greenwell et al. 1996; Hay 1998), as this anuran is very

susceptible to desiccation (Ralin and Rogers 1972) and mainly disperses following heavy

rains (Pyburn 1958). This problem is exacerbated by the isolated nature of many

wetlands in the Midwest and anthropogenic habitat fragmentation (Green 2003). A. c.

blanchardi is thus faced with a "catch-22" situation during droughts in the upper

Midwest: either forego breeding and risk zero recruitment, or breed in highly degraded

habitats that have been modified by human use (Pechmann et al. 1991; Lannoo 1998b).

Irwin (2005) describes a similar situation in which the behaviour and physiology of A. c.

blanchardi, combined with climate change, may be contributing to declines. This frog

has a limited tolerance for water loss (Ralin and Rogers 1972) and freezing (Irwin et al.

1999). Throughout much of the Midwest, A. c. blanchardi overwinters in existing cracks

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 21 -

or crayfish burrows in the pond banks, where the moisture content of the soil provides a

thermal refuge (Gray 1971; Irwin et al. 1999). Dry climatic conditions would reduce the

moisture content of the soil, thereby diminishing the soil's capacity to buffer against

freezing temperatures. Indeed, a severe drought in the autumn of 1976 and extremely

cold winters in 1976 and 1977 coincided with a drastic decline in A. c. blanchardi

populations in both Wisconsin and on Pelee Island, ON (Irwin 2005). Winter mortality

may have been increased in these instances by pond dredging and clearing of shoreline

vegetation and debris; furthermore, the exotic rusty crayfish ( Orconectes rusticus), which

does not burrow, has displaced native burrowing crayfish and may also prey on A. c.

blanchardi (Irwin 2005).

Overall, it appears that declines of A. c. blanchardi populations may generally be

due to habitat alteration and climate change. Other insults ( e.g. environmental

contaminants, parasites, deformities, introduced species) may occur synergistically with

or independently of these factors, but likely exacerbate declines to varying degrees in

different locations. One factor that cannot be ignored regarding these declines is that

they are primarily occurring across the northern and western edge of this anuran's range.

Such peripheral populations are more likely to be isolated from central populations and

inhabit less suitable environments, and therefore tend to be smaller and more prone to

extirpation (Lesica and Allendorf 1995). Midwestern A. c. blanchardi populations exist

in marginal habitat and at the limits of their ecological tolerance; any slight perturbation

to their environment is much more likely to trigger population declines than if it occurred

at the centre of the range. This is supported by the greater abundance and stability of

populations in the central and southern Great Plains, despite a likely similar onslaught of

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 22 -

environmental disturbances (Lannoo 1998b; Gray and Brown 2005). Another proposal is

that the return of forested landscapes to the Midwest following widespread abandonment

of agriculture may be pushing A. c. blanchardi, which prefers aquatic habitats with open

canopy, out of the area (Lehtinen and Skinner 2006). Similarly, trampling by cattle at

one farm pond in Wisconsin produced mud banks with low vegetation that appeared to

favour A. c. blanchardi (Hay 1998). This large population disappeared once cattle

activity ceased and the banks were revegetated by tall grass. These observations imply

that A. c. blanchardi was only able to colonise the Midwest recently, following the

conversion of forests into agricultural land, and that current range contractions are either

the result of inevitable fluctuations of unstable peripheral populations, or the gradual

return to native conditions as cricket may not have inhabited these areas prior to

anthropogenic modification of the landscape (Lehtinen 2002; Lehtinen and Skinner

2006). However, a much clearer understanding of these declines, and the history of A. c.

blanchardi in this region, must be obtained before they can be deemed "natural".

Thesis objectives and overview

It is my view that genetic information is a vital component to the proper,

effective, and efficient conservation of endangered species. The objective of my research

is therefore to examine the utility and importance of genetic information to the

development of recovery strategies, using two endangered amphibians as case studies.

These species differ substantially in their life histories, ecologies, levels of endangerment,

and extent of recovery efforts to date. P. lemur is a medium-sized, tropical, terrestrial

toad with a very restricted distribution, few populations, and small population size. It has

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 23 -

an adult lifespan of several years in the wild, and breeds sporadically and explosively in

temporary pools. The causes of population declines are well characterised, and are

primarily related to habitat destruction and alteration. Conservation efforts have been

extensive for over 20 years, with toads bred in captivity at several institutions and large

numbers released to the wild. In contrast, A. c. blanchardi is a small, temperate,

semiaquatic treefrog with an extensive distribution in North America and many large

populations in some areas. It has an adult lifespan of only 4 months in the wild, and

breeds each spring in small ponds and wetlands. The causes of declines are not well

understood, and likely involve multiple factors, both direct and indirect. Very limited

conservation activity has occurred thus far, and attempts to breed this anuran in captivity

have not been successful (B. Johnson, pers. comm.). The commonality between these

two species is the lack of genetic information that existed before my research was

conducted: although some studies had been performed (Dessauer and Nevo 1969; Salthe

and Nevo 1969; Lacy and Foster 1987; Gorman and Gaines 1987; Ward et al. 1987;

Goebel 1996; Moore 1997; Estep 2000), they primarily used very conservative genetic

markers and/or limited sample sizes of restricted geographic distribution.

In this dissertation, I have used mitochondrial control region sequences and

several highly polymorphic microsatellite loci to characterise genetic variation and

structure within P. lemur and all Acris taxa. Mitochondrial DNA (mtDNA) evolves

relatively slowly, at a rate of 0.5-1 % per lineage per million years in primates (Moritz et

al. 1987), and has been used extensively to examine relationships among species (i.e.

phylogeny) and among populations within species (i.e. phylogeography) (Avise 1994,

2000). The various genes within mtDNA evolve at different rates; the most rapid is the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 24 -

control region, as it does not code for a functional product and thus experiences less

selective pressure (Moritz et al. 1987). Few amphibian studies have utilised the control

region, as it can be extremely divergent within a single species, and may suffer from

multiple mutations at individual sites (i.e. saturation). However, preliminary analyses of

both P. lemur and A. c. blanchardi indicated that the more conservative cytochrome b

gene possessed very limited variation within each taxon, and that the control region was

3 more appropriate. Microsatellite markers evolve very rapidly, around 10- events per

locus per generation in humans (Weber and Wong 1993), and are typically highly

polymorphic even among recently diverged groups. This affords a high resolution suited

to very fine-scale studies; as such, they are commonly used for individual identification,

parentage, population structure, gene flow, dispersal, and landscape genetics (Jehle and

Arntzen 2002; Tennessen and Zamudio 2003; Bums et al. 2004; Funk et al. 2005;

Kraaijeveld-Smit et al. 2005; Zamudio and Wieczorek 2007). However, they are also

finding use in large-scale phylogeographic studies (Zeisset and Beebee 2001; Rowe et al.

2006; Knopp and Merila 2008), particularly since concordance between mitochondrial

and nuclear markers provides evidence that genetic structure reflects historical patterns of

isolation rather than selection or the vagaries of a single stretch of DNA (A vise 2000).

The information obtained using mtDNA and microsatellite markers will be used

to address several specific issues for both P. lemur and A. c. blanchardi: 1) levels of

genetic diversity in populations across the range; 2) degree of population structure and

genetic differentiation; 3) whether current taxonomy reflects genealogical relationships;

and 4) identification of distinct lineages or populations. These results will enable me to

assess whether detailed genetic information would have been valuable at the initiation of

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 25 -

P. lemur recovery activities, and how useful it will be to the development of future

recovery efforts for both P. lemur and A. c. blanchardi. I will also make specific

recovery recommendations for each species using the newly acquired genetic data.

The following chapters describe the genetic analyses of P. lemur and A. c.

blanchardi (and related Acris taxa). In Chapter 2, I profile three populations of P. lemur

at six newly developed microsatellite loci and a fragment of the mitochondrial control

region. This includes the two captive populations at the Toronto Zoo, representing the

extirpated northern population (Quebradillas) and the remaining wild population in the

south (Guanica), as well as wild tadpoles from the Tamarindo breeding pond in Guanica.

I use this information to determine levels of genetic variation in each group, the genetic

differentiation between them, assess the appropriateness of the current breeding strategy,

and provide recommendations for future activities to maximise the probability that

individuals reintroduced to the wild will establish viable populations. Chapter 3

describes the development of 17 microsatellite loci for A. c. blanchardi, and their utility

in related Acris taxa. Individuals from two widely separated populations (Texas and

Illinois) are profiled at 16 of these loci to evaluate differences in genetic diversity

between the two regions. Chapter 4 is a detailed study of the phylogeographic patterns

and spatial structure of A. c. blanchardi across its range using mitochondrial control

region sequences. It demonstrates the influence of Pleistocene glaciations on genetic

structure of this anuran, and provides evidence that northern populations are a distinct

clade from those to the south. They are therefore valuable for conservation and should be

managed separately. In Chapter 5, I use mitochondrial control region sequences to

examine phylogeographic patterns and the current taxonomy of all A eris species and

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 26 -

subspecies. Based on this information, A. c. blanchardi appears to be a distinct species

from A. c. crepitans, while A. c. crepitans and A. gryllus contain several divergent

lineages. Major biogeographic patterns of this complex are described, and indicate that a

number of adjustments to currently accepted distributions are required. In Chapter 6, I

synthesise the results of chapters 3, 4 and 5 to provide a comprehensive evaluation of

genetic structure in A eris. I also discuss the importance of the genetic data to the

conservation of P. lemur, A. c. blanchardi, and amphibians in general.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 27 -

Chapter 2

Genetic Rescue of an Inbred Captive Population of the Critically Endangered Puerto Rican Crested Toad (Pe/tophryne lemur) by Mixing Lineages

This chapter has been published exactly as it occurs here as:

Beauclerc KB, Johnson B, White BN (2009) Genetic rescue of an inbred captive population of the critically endangered Puerto Rican crested toad (Peltophryne lemur) by mixing lineages. Conservation Genetics in press. doi: 10.1007/s 10592-008-9782-z

KB Beauclerc performed all laboratory work, data analysis, and wrote the paper. B Johnson provided valuable input and funding for the work. BN White is the principal investigator responsible for the work.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 28 -

Abstract

The Puerto Rican crested toad (Peltophryne lemur) is currently composed of a single wild

population on the south coast of Puerto Rico and two captive populations founded by

animals from the northern and southern coasts. The main factors contributing to its

decline are habitat loss, inundation of breeding ponds during storms, and impacts of

invasive species. Recovery efforts have been extensive, involving captive breeding and

reintroductions, habitat restoration, construction of breeding ponds, and public education.

To guide future conservation efforts, genetic variation and differentiation were assessed

for the two captive colonies and the remaining wild population using the mitochondrial

control region and six novel microsatellite loci. Only two moderately divergent

mitochondrial haplotypes were found, with one fixed in each of the southern and northern

lineages. Moderate genetic variation exists for microsatellite loci in all three groups. The

captive southern population has not diverged substantially from the wild population at

microsatellite loci (Fsr = 0.03), whereas there is little allelic overlap between the northern

and southern lineages at five of six loci (Fsr > 0.3). Despite this differentiation, they are

no more divergent than many populations of other amphibian species. As the northern

breeding colony may not remain viable due to its small size and inbred nature, it is

recommended that a third breeding colony be established in which northern and southern

individuals are combined. This will preserve any northern adaptive traits that may exist,

and provide animals for release in the event that the pure northern lineage becomes

extirpated.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 29 -

Introduction

The evaluation of current taxonomy and identification of genetic lineages within

threatened species are vital to developing efficient and optimal recovery programs. For

example, the use of morphological traits to designate taxonomic entities often does not

reflect patterns of genetic variation (Burbrink et al. 2000; Culver et al. 2000; Nittinger et

al. 2007), which can waste limited resources on groups that are not distinct and have

limited value to the conservation of the species as a whole (Avise and Nelson 1989; Zink

et al. 2000). A lack of distinct lineages can also facilitate recovery, as it may increase the

number of source populations for efforts such as captive breeding or translocation

between isolated populations (Eizirik et al. 2001; Godoy et al. 2004). In contrast, early

detection of multiple lineages may enable intervention before an evolutionarily

significant lineage becomes irreversibly imperiled (Nielson et al. 2001; Shaffer et al.

2004a; Vredenburg et al. 2007), as occurred for the tuatara (Daugherty et al. 1990).

Knowledge of the genetic structure of a species can identify the most closely related

population or subspecies for use in reintroductions following extirpation (Drew et al.

2003; Kraaijeveld-Smit et al. 2005), while individuals from divergent populations may be

incorporated to maximise genetic variation (Madsen et al. 1999; Land et al. 2004).

Finally, assessment of the genetic composition of existing captive programs can

determine whether animals are suitable for further breeding (Brock and White 1992;

Ruokonen et al. 2000) or reintroduction (Garcia-Moreno et al. 1996; Wyner et al. 1999).

The Puerto Rican crested toad (Peltophryne lemur) is a highly endangered

bufonid for which two separate lineages are currently maintained in captivity. Captive

breeding was initiated prior to the frequent use of molecular markers to identify genetic

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 30 -

structure within species. In this study, mitochondrial DNA (mtDNA) sequences and six

novel microsatellite loci were used to investigate the genetic variation and differentiation

of the two P. lemur captive breeding colonies, as well as the sole remaining wild

population. This information will be used to determine the optimal future management

actions to ensure the persistence of P. lemur in the wild.

Biology and conservation history of Peltophryne lemur

The only toad endemic to Puerto Rico, P. lemur has been collected from at least

eight locations across Puerto Rico and one on Virgin Gorda of the British Virgin Islands

(Figure 2.1 ). It generally occupies low-elevation ( < 200m), arid or semi-arid coastal

regions and is often associated with limestone karst formations that provide fissures into

which it retreats during the day (Rivero et al. 1980; Miller 1985). Breeding is sporadic,

and occurs in temporary pools formed by heavy rains (USFWS 1992). P. lemur has not

been seen on Virgin Gorda for over 40 years, and was thought to be extinct on Puerto

Rico until 1965 (USFWS 1992). Populations have since been found on both the northern

(Quebradillas and Isabela) and southern (Guanica State Forest Reserve) coasts of Puerto

Rico (Rivero et al. 1980; USFWS 1992), although no toads have been seen in the north

since 1988 (Barber 2007). It is therefore probable that Guanica, which may have

numbered as few as 80 individuals in 2003 (Hedges et al. 2004 ), is the only remaining

wild population. P. lemur is listed as critically endangered by the International Union for

Conservation of Nature. The primary cause of decline is habitat loss and alteration,

including drainage of breeding sites and spraying of surrounding areas with pesticides

(Johnson 1990; USFWS 1992). Predation and/or competition by introduced species (e.g.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Figure regions (after captive Monroe 2.1. zoo of Puerto lsabela programs Map 1976). Rico of Puerto ( * are ). 0 indicated: San Rico Juan and Mountainous (@) the British the I' is karst the interior 50 belt capital Virgin in Coastal the of Islands Puerto northeast, plain showing Rico 100 the and mountainous historical is shown Atlantic Caribbean ~ locations for 150 interior, reference Kilometres Ocean ~~ of Sea and P. only. lemur the discontinuous The ( •) A N three and r:;p·ir sites ~;n\\s"'\J\~ main coastal represented Virgin physiographic ~~) Gorda plain by - 31 -

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 32 -

cane toad, Bufo marinus; mongoose, Herpestesjavanicus; feral dogs and cats) may have

contributed to declines (Rivero et al. 1980; USFWS 1992). The small size and isolation

of the remaining population in Guanica indicates that it is also vulnerable to stochastic

environmental ( e.g., hurricanes; Johnson 1990) and demographic events.

Recovery efforts for P. lemur have been extensive. The Guanica breeding pond is

closed for three weeks during the breeding season and is no longer drained for parking

(Johnson 1990). Toads have been radiotracked to monitor movements, and habitat

characteristics have been evaluated (Johnson 1990; Matos-Torres 2006). Wild

individuals were collected during the 1980s to establish two breeding colonies

representing the northern (four founders) and southern (32 founders) populations. To

date, more than 110 000 southern tadpoles and toads have been released into constructed

ponds in Guanica Forest, and releases of northern animals began at Arecibo in 2006

(Barber 2007). In 2003 and 2005, captive-born adult toads returned to the southern

release site to breed (PRCT SSP 2006). Additional property is being secured and several

ponds constructed in both regions, while public outreach and education are a continuing

component of the recovery program (Barber 2007).

In order to maintain the optimal recovery strategy for P. lemur, several important

issues must be resolved using genetic data. On average, a founding population of 20

individuals will capture> 95%ofheterozygosity and approximately 87% of allelic

diversity (Frankham et al. 2002). While the southern captive colony may have sampled

sufficient genetic variation from the wild to maintain a healthy population for many

generations, the northern colony likely had low variation from its initiation. The

potential for increased homozygosity due to inbreeding and loss of genetic variation

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 33 -

through drift are factors that must also be considered in the breeding populations. This is

especially applicable to the northern colony, as the 123 individuals currently in captivity

and used for breeding are the product of four siblings that themselves are descended from

two first cousins. In addition, the isolation of the captive colonies from each other and

the wild population may have led to genetic or morphological divergence between them

(Rowe et al. 1998; Shuster et al. 2005; Hakansson et al. 2007). The degree of genetic

differentiation between the northern and southern populations also requires assessment.

Fossil evidence suggests that P. lemur was widespread throughout Puerto Rico during its

history (Pre gill 1981 b ); however, it is currently restricted to low-elevation areas on

opposite coasts, separated by the Cordillera Central mountain range. Previous allozyme

assays suggested that the two populations may be divergent (Lacy and Foster 1987). It

thus remains unclear whether the northern and southern populations have been separated

throughout their history, or if population declines and habitat fragmentation have isolated

them more recently.

The genetic data obtained here will be used to address three main questions: (1)

do the captive populations suffer reduced genetic diversity relative to the wild population;

(2) has the captive southern population diverged genetically from the wild southern

population; and (3) what is the level of differentiation between the populations from the

northern and southern coasts? Given the precarious state of the northern breeding colony

and its probable extirpation in the near future, the long-term reintroduction of purely

northern individuals is not realistic: The information from this study will thus be used to

discriminate between the few remaining options for this lineage: ( 1) continue breeding

the northern and southern populations separately. If the northern colony becomes

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 34 -

extirpated, terminate reintroductions in this region; (2) continue breeding the northern

and southern populations separately. If the northern colony becomes extirpated,

reintroduce individuals from the southern lineage to the northern region; or (3) create an

additional breeding colony in which northern and southern individuals are mixed. If

adaptive differences exist between the northern and southern populations, this approach

will facilitate the preservation of these traits. In the event that the pure northern colony

becomes extirpated, these mixed individuals can be reintroduced to the north and may

experience greater survival over purely southern animals due to the presence of some

northern adaptations.

Materials and Methods

Sample collection and DNA extraction

Muscle or organs (heart, kidney, liver) were collected post-mortem from

individuals of the northern and southern breeding colonies at the Toronto Zoo and

Audubon Nature Institute in New Orleans. In total, four individuals from the northern

colony and 58 from the southern colony were sampled. During a breeding of wild toads

in 2003 at Tamarindo Pond in Guanica State Forest, Puerto Rico, an additional 43

tadpoles that had died were collected opportunistically. The southern breeding colony

was founded by individuals collected from several breeding ponds, including Tamarindo,

during the 1980s. Differential mortality of tadpoles may reflect the genotype at the major

histocompatibility complex (Barribeau et al. 2008), suggesting that the targeted collection

of dead individuals may not constitute a random genetic sample. However, the perilous

state of P. lemur in the wild mandates that sampling impacts be minimised. Furthermore,

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 35 -

the high mortality rate of amphibian larvae, particularly in temporary ponds such as those

used by P. lemur (Duellman and Trueb 1986), indicates that assumptions of genotype

interactions may be premature without supporting information. Hereafter, the southern

breeding colony will be called Guanica, the northern breeding colony will be called

Quebradillas, and the wild southern population will be called Tamarindo.

All 105 samples were genotyped at the six microsatellite loci to ensure that no

rare alleles were missed. However, maternal inheritance of the mitochondrial genome

means that all offspring of a female possess the same haplotype; thus, mtDNA was

sequenced for only a subset of samples from the breeding colonies. According to the P.

lemur studbook maintained by the Toronto Zoo, the 58 Guanica samples employed in this

study trace their lineage to six wild females and 14 wild males. Thirty captive-bred

individuals representing these six matrilineal lineages were sequenced, as were five male

founders that could potentially harbour different mitochondrial lineages. Due to the

small sample size for Quebradillas, all samples were sequenced. Because the genealogy

of the Tamarindo individuals is unknown, all 44 samples were sequenced.

Although the number of samples used for the northern colony is small, it must be

emphasised that a very limited gene pool is available for profiling. Only four wild

individuals (two male and two female) were used to establish the colony, and all current

breeding individuals are descendants of two cousins. Two of the samples profiled here

are direct offspring of these cousins, and the other two are the product of a mating

between two of these offspring (i.e. between full siblings). Tissue samples of earlier

generations are not available. Thus, a maximum of only one mitochondrial lineage and

four alleles for each microsatellite locus are possible in any combination of samples from

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 36 -

this population. Additional sampling would therefore not be informative, and the four

samples analysed here are representative of the variation present in the northern colony.

Samples were prepared for extraction by dissolving 10 mg of tissue in 500 µL of

1 X lysis buffer (2 M urea, 0.1 M NaCl, 0.25% n-lauroyl sarcosine, 5 mM CDTA, 0.05 M

Tris-HCl pH 8), and digesting with 50 µL of proteinase K. Total genomic DNA was

extracted with the DNeasy Tissue Kit (Qiagen Inc.) and quantified with the PicoGreen

Quantitation Kit (Molecular Probes) on a FLUOStar Galaxy (BMG Labtech).

Mitochondrial control region sequencing

A fragment of the control region was amplified with the primers CytbA-L and

ControlP-H (Goebel et al. 1999) in a 25 µL reaction containing 1-5 ng DNA, 1 X PCR

buffer, 2 mM MgCh, 0.2 mM dNTPs, 0.1 mg/mL BSA, 0.2 µM each primer, and 2 U

Taq polymerase (Invitrogen). Thermal cycling used an MJ Research PTC-225 thermal

cycler with the following conditions: denaturation at 94°C for 5 min, 30 cycles of

denaturation at 94°C for 30 sec, annealing at 55°C for 30 sec, and extension at 72°C for

45 sec, with a final extension of 72°C for 10 min. PCR product was visualised with

ethidium bromide on 1 %agarose gels, and 5 µL was purified with ExoSAP-IT (USB).

Both strands were sequenced using the PCR primers with the DYEnamic ET-Terminator

cycle sequencing kit on a MegaBACE 1000 DNA Analysis System (GE Healthcare).

Microsatellite development and profiling

Tetranucleotide microsatellites were isolated according to the protocol of

Beauclerc et al. (2007) (see also Chapter 3). In total, 115 clones were sequenced, of

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 37 -

which 83 were positive for a tetranucleotide repeat, with five duplicate clones found.

Primer pairs were designed manually for 20 positive clones that possessed sufficient

flanking sequence. Individual loci were amplified in 10 µL volumes containing 5 ng

genomic DNA, 1 X PCR buffer, 1.5 mM MgCh, 0.2 mM each dNTP, 0.3 µMeach

primer, and 0.5 U Taq polymerase using the following thermal cycling conditions:

denaturation at 94°C for 5 min, 30 cycles of denaturation at 94°C for 30 sec, annealing at

the optimal temperature (see Table 2.1) for 1 min, and extension at 72°C for 1 min, with a

final extension of 60°C for 45 min. Nine and three individuals from Guanica and

Quebradillas, respectively, were screened for polymorphism by visualising PCR products

on 4% agarose gels stained with SYBR Green I (Molecular Probes). Six loci amplified

reliably and were polymorphic. To profile all samples, each locus was amplified

individually and pooled into two multiplex reactions prior to desalting (Table 2.1 ). All

PCR product was desalted using Sephadex G50, profiled on a MegaBACE 1000 DNA

Analysis System, and scored using GENETIC PROFILER v.2.0 (GE Healthcare).

Data analysis

Sequences were aligned in CLUSTALX v.1.81 (Thompson et al. 1997) and edited

by eye using BIOEDIT v.7.0.533 (Hall 1999). Pairwise sequence divergence was

calculated using MEGA v.3.1 (Kumar et al. 2004). Microsatellite diversity indices

(number and frequencies of alleles, observed and expected heterozygosity) were

calculated using GENEPOP v .3 .4 (Raymond and Rousset 1995), and conformation to

Hardy-Weinberg equilibrium was tested by estimation of exact P-values with a Markov

chain method (Guo and Thompson 1992). Allelic richness (number of alleles per

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Table and

a

b c motif.

d

gels

Fluorescently

Annealing

Repeat

Primer

Multiplex

Multiplex

PleMs-14

PleMs-10 pooled

PleMs-11

PleMs-4

PleMs-8 PleMs-3

stained

Locus

2.1.

See

concentration

motif

into

Primers

GenBank

temperature

with

1 2

is

(I'a

(I'a

labeled

multiplexes

F: R:CCTGAAAAAAACTGAGAGATGG

R:CCCAGGGTACTGCAACTCG F: R: F: F: R: R:CAGGTTTTGAGAAGAGTTCC F:

F:

R:

based

SYBR

=

=

TGCCACTGAGAAAGATTTGG

ATGGGTGAATAAAGACCTCC

CGTACCAGAAACTAATCTCAACTGG

GACTATGTATGTGTGTGTAGC

GGGAAACTGGAGCAAATACC

TCCATTACCTTCTCAGTGTTGC

and

AGTTGTGACTGCTGTGACC TCAGTTCCT

TTCTGT

55°C)

55°C)

for

is

primers

on

amplification

Green

is

complete

for

sequence

the

prior

both

AAGGTCTGGCTGC

range

I.

are

~(_5'-3'l

Primer

to

the

ATGCACTGAGC

sequences.

indicated

from

desalting.

of

forward

conditions

sequence

temperatures

the

by

original

and reverse

the

for

letters

(H)

six

at

clone.

(H)

(H)

(F)

which

(H)

microsatellite

(F)

primers

Hand

"Imperfect"

the

F,

single

(GATA)

(GATA)13 (GATA)1s

(GATA) in

(GATA)12

imperfect

(GATA)9

imperfect

(CATA)

which

Repeat

motifb

the

loci

locus

reaction.

indicates

refer

isolated

36

12

3

amplified successfullyamplified

to

Ta

55-60

50-60

55-65 55-60

HEX 55-60

from

that

55

(°Ct

one

and

P.

lemur.

or

6-F

more

Primer

(µMt

AM,

0.6 0.2

0.3 0.3

0.6

0.3

Loci

base

respectively.

and

were

pairs interrupted

cleanly,

Pooling

volume

2.5

10

10

10

amplified

5

5

µL

µL

µL

µL

µL

µL

based

individually

accession#

on

EU149940

EU149942 EU149941 EU149944

EU149945

EU149943

GenBank

the

agarose

repeat

-

-

38

-

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 39 -

population) and private allelic richness were calculated by standardisation to the smallest

sample size (N = 4) using rarefaction in the program HP-RARE v.1.0 (Kalinowski 2005).

Population differentiation was estimated between all pairs of populations for all

loci with Nei's genetic distance Ds (1972) and the genotype likelihood ratio distance DLR

(Paetkau et al. 1997) using software from http://www2.biology.ualberta.ca/jbrzusto/ (J.

Brzustowski, unpublished). These distance measures are calculated in very different

manners, and Paetkau et al. ( 1997) recommended their use to provide independent

estimates of genetic distance between populations. Population structure was evaluated

with 0w (Weir and Cockerham 1984), the unbiased estimator of Wright's (1951) FsT,

using ARLEQUIN v.3.1 (Excoffier et al. 2005). Significance was determined by a

permutation test of 1000 iterations. Genetic differentiation was further evaluated using

the clustering program STRUCTURE v.2.0 (Pritchard et al. 2000). The likelihood that the

samples represented between one and five genetic clusters (K = 1 to K = 5) was

determined, using the following model: admixture, correlated allele frequencies, bum-in

of 500 000 steps, and 1 000 000 iterations of the Markov Chain Monte Carlo algorithm.

The model was run four times for each value of K to ensure consistency between runs.

The largest value of !::.K (Evanno et al. 2005) was used to determine the most likely

number of populations.

Results

Mitochondrial sequencing

Quality sequence was obtained for 963 bp of the mitochondrial control region,

yielding two haplotypes (GenBank accession numbers EU149946 and EU149947):

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 40 -

PleCR-1 occurred only in Guanica and Tamarindo, and PleCR-2 was exclusive to

Quebradillas. Sixteen variable sites were found: 13 transitions, one transversion, and two

indels of one base pair. Pairwise sequence divergence between the haplotypes was 0.015.

Microsatellite profiling

The total number of alleles for the six microsatellite loci was 46, with a range

from four to ten and an average of 7. 7 alleles per locus (Table 2.2). Guanica possessed

the largest number of alleles; however, the disparate sample size between groups clearly

contributed to this outcome. When standardised by rarefaction, allelic richness was

nearly equal for all three populations (Table 2.2). Each population also had several

private alleles (Figure 2.2, Table 2.2): Guanica possessed six, Tamarindo three, and

Quebradillas twelve. The sampling of amphibian tadpoles from a single location is

typically discouraged as it may consist of a group of siblings and yield biased data.

However, the observation of seven alleles at locus PleMs-3 in the Tamarindo samples

indicates that this group is descended from at least four parents, and individuals are

therefore not all full siblings.

Observed and expected heterozygosities ranged from 0.41 to 1.00 and 0.43 to

0.79, respectively (Table 2.3). Following sequential Bonferroni correction, all loci were

in Hardy-Weinberg equilibrium and linkage equilibrium in Quebradillas and Tamarindo.

In Guanica, however, loci PleMs-8 and PleMs-14 continued to deviate significantly from

Hardy-Weinberg equilibrium due to heterozygote excess, and significant linkage

disequilibrium remained between loci PleMs-14/PleMs-l 0, PleMs-14/PleMs-3, PleMs-

10/PleMs-11 and PleMs-10/PleMs-3. It is probable that the disequilibrium in Guanica is

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. populations Table

PleMs-11

PleMs-10

PleMs-14

PleMs-3 PleMs-4

PleMs-8 Average

Locus

Total

2.2.

Number

of

Allele

P.

Min

282

311 143

129

147

87

lemur.

of

size

alleles,

Max

Corresponding

201 426

298

173 128

157

(bp)

allelic

richness

'fotal

7.67

46

10

3 6 3 4 9

8 9

indices

No.

when

alleles

Guanica

for

31

7

() 6(0) 7(1)

5(2) 6(1)

5.17

(2)

(0) (0)

private

(6)

rarefied

(no.

private

alleles

28

to

Tam.

4(1) 7 3

3

5(0)

4.67

(2) (0)

(0)

N

(3)

=

alleles)

are

4,

and

shown

.

17

Queb.

2(1)

3 4(3) 3 2 3(3)

2.83

(1)

(1) (3)

(12)

size

in

range

parentheses.

of

19.26

4.28 2.23

2.65 (0.92) 3.08 3.65 3.37

5.51

Allelic

Guanica

alleles

(1.73) (0.13)

(1.23) (0.07)

(0.89)

(1.42)

(4.97)

richness

for

six

17.37

3.48

2.11 2.59 2.76 3.14

3.29 4.96

microsatellite

(private

Tam.

(0.88) (0.49) (0.60)

(1.00) (0.03)

(1.00) (0.52)

(3.52)

allelic

loci

4.00 3.00 2.00 3.00 2.00 3.00

4.85

17

richness)

Queb.

(12.27)

in

(1.01) (3.00) (3.00) (3.00) (1.00) (1.26)

(3.51)

three

-

41

-

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 42 -

0.70 0.50 0.60 PleMs-3 0.40 0.50 0.40 0.30 0.30 0.20 0.20 0.10 0.10 0.00 0.00 143 145 147 149 155 159 163 167 171 173 311 315 319 323 327 398 402 406 426 * * * * * * * * * 0.80 0.80 0.70 PleMs-4 0.70 PleMs-11 >, (J 0.60 0.60 C 0.50 0.50 Q) ::J 0.40 0.40 C'" ....Q) 0.30 0.30 0.20 0.20 Q) - 0.10 0.10 Q) 0.00 0.00 <( 87 95 99 103 120 128 129 133 137 141 145 149 153 157 * * * * * 0.60 0.80 PleMs-8 0.70 PleMs-14 0.50 0.60 0.40 0.50 0.30 0.40 0.30 0.20 0.20 0.10 0.10 0.00 0.00 147 172 176 180 184 188 193 197 201 282 286 290 298 * * * * * * * Allele

Figure 2.2. Allele frequencies at six microsatellite loci for three populations of P. lemur: Guanica (shaded), Tamarindo (hatched) and Quebradillas (solid). Private alleles are indicated by *.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 43 -

Table 2.3. Observed heterozygosity (Ho), expected heterozygosity (HE), and exact P- values of Hardy-Weinberg equilibrium for six micro satellite loci in three populations of P. lemur.

Guanica Tamarindo Quebradillas Locus HE Ho p HE Ho p HE Ho p PleMs-3 0.79 0.81 0.13 0.73 0.84 0.44 0.61 0.50 0.43 PleMs-4 0.46 0.41 0.23 0.60 0.49 0.34 0.68 1.00 0.31 PleMs-8 0.66 0.71 0.0002 0.62 0.56 0.18 0.68 0.75 1.00 PleMs-10 0.73 0.72 0.06 0.69 0.67 0.94 0.79 1.00 1.00 PleMs-11 0.60 0.71 0.16 0.59 0.60 0.80 0.43 0.50 1.00 PleMs-14 0.59 0.78 0.02 0.43 0.51 0.25 0.54 0.75 1.00 Overall 0.64 0 0.62 0.57 0.61 0.98

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. -44 -

due to the managed breeding of this closed population. The fact that all loci are in

equilibrium in the wild population, which presumably experiences random mating,

suggests that the loci are not linked or under selection and are therefore suitable for use in

further analyses.

Guanica and Tamarindo shared the greatest number of microsatellite alleles, 25

out of 46 total alleles (54%; Figure 2.2). When excluding alleles found only in

Quebradillas, Guanica and Tamarindo shared 74% of their alleles (25 out of 34). Very

few alleles were common between Quebradillas and the southern populations: none were

shared for loci PleMs-4 and PleMs-8, and only one was shared at loci PleMs-10, PleMs-

11, and PleMs-14. In total, Quebradillas shared only five alleles (11 %) with both

Guanica and Tamarindo; its remaining 12 alleles were unique. Both Guanica and

Tamarindo are much more distant from Quebradillas than they are from each other based

on the pairwise genetic distance measures Ds and DLR (Table 2.4). FsT values were

significant at P = 0.0l between all pairs of populations, but values were much larger

between either of the southern populations and Quebradillas (Table 2.4).

The average log probability of K [ln Pr(XJK)] increased with the number of

clusters (data not shown). However, the rate of change of the log probability of K (t:J()

was largest for K= 3 (tiK2 = 2.6; M 3 = 551.5; M 4 = 117.1). This was-therefore chosen

as the number of groups most likely represented by the microsatellite data. Guanica and

Tamarindo individuals were both assigned in similar proportions and nearly equally to

clusters 1 and 2, with no individuals assigned to cluster 3 at> 2% of their genome ( data

not shown). All Quebradillas individuals were strongly assigned to cluster 3 (> 98% of

each genome).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 45 -

Table 2.4. Pairwise values of FsT (calculated as 0w), Nei's genetic distance (Ds), and the genotype likelihood ratio distance (DLR) using six rnicrosatellite loci for three populations of P. lemur. All 0w values were significant at P < 0.01.

Population pair 0w Ds Gminica/Tarnarindo 0.03 0.06 1.07 Guanica/Quebradillas 0.31 1.65 16.02 Tarnarindo/Quebradillas 0.34 1.76 16.16

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 46 -

Discussion

Genetic diversity within populations

The control region is the most variable section of the mitochondrial genome

(Moritz et al. 1987), making the very low haplotypic diversity of P. lemur somewhat

surprising. This bufonid has one of the lowest mitochondrial diversities among a number

of amphibian species: multiple haplotypes typically exist for this and less variable

regions (Bos and Sites 2001; Austin et al. 2002; Smith and Green 2004), and often over

relatively small scales (Shaffer et al. 2004b; Dever 2007). However, Rowe et al. (2006)

found that all British and Irish individuals of Bufo calamita possessed a single control

region haplotype.

Peltophryne lemur is moderately variable at the six microsatellite loci. Within­

population genetic diversity (HE) is similar to or greater than that found for microsatellite

loci in several other amphibians (Rowe et al. 1998; Palo et al. 2003; Brede and Beebee

2004; Martinez-Solano et al. 2005; Arens et al. 2006), while only a few have been found

to have substantially greater HE (Austin et al. 2003a; Bums et al. 2004; Hoffman et al.

2004). Allelic diversity for P. lemur is comparable to several species examined over

similar and larger geographical scales, despite much larger sample sizes for some (Rowe

et al. 1998; Palo et al. 2003; Brede and Beebee 2004; Kraaijeveld-Smit et al. 2005; Arens

et al. 2006). This diversity is surprising, considering the small sample sizes, relatedness

of individuals, and limited variation in mtDNA. It thus appears that neither the captive

nor wild populations of P. lemur have suffered a severe reduction in heterozygosity due

to small population sizes and prolonged isolation. In fact, the Guanica captive colony

had greater heterozygosity than expected at two loci. This likely reflects the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 47 -

minimisation of kinship management system used to designate pairs for breeding

(Johnson 1990; Ballou and Lacy 1995), which reduces the tendency to homozygosity via

inbreeding and the loss of alleles through genetic drift (Montgomery et al. 1997). By

comparison, random mating in the wild Tamarindo population, combined with a possible

bottleneck of 80 individuals (Hedges et al. 2004), has allowed these processes to reduce

heterozygosity below that of Guanica for several loci.

Comparison ofpopulations

There appears to be very little difference in the amount of genetic variation found

in Guanica and Tamarindo. Both populations possess similar numbers of alleles and HE

estimates for most loci. There is also little divergence between them: a single

mitochondrial haplotype exists in all individuals, there are very few unshared alleles, and

the measures of genetic distance and differentiation are very low. Comparable estimates

of these measures have been found over similar distances in several other amphibian

species (Newman and Squire 2001; Brede and Beebee 2004; Bums et al. 2004; Funk et

al. 2005; Arens et al. 2006). In addition, the STRUCTURE algorithm could not distinguish

individuals from these two groups. The genetic similarity of these populations is not

surprising, as the Guanica captive colony is recently descended from the wild Tamarindo

population. Although gene flow between captive-bred individuals released into Guanica

State Forest and wild individuals located at Tamarindo pond cannot be completely

excluded, it is highly unlikely as P. lemur has been shown to move only up to 2 km

(Johnson 1999) and these locations are separated by 20 km of migratory distance.

Despite the fact that only a small number of related, inbred individuals were

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 48 -

sampled, moderate genetic variability exists at nuclear loci in Quebradillas. It is probable

that most alleles present in the northern population were detected during this analysis, as

50-100% of the maximum number of possible alleles (four) was observed for all loci.

Thus, more extensive sampling would likely have revealed little additional variation.

Although it possesses fewer alleles than Gminica or Tamarindo based on direct counts, it

is nearly equal when all groups are standardised to N = 4. As well, HE is very similar to

both southern populations. Quebradillas is moderately divergent from the southern

populations at mtDNA, with 1.5% sequence divergence between the two haplotypes. A

comparison with other amphibian mitochondrial studies shows that many species possess

deeper divergences in the control region (Shaffer et al. 2004b) and more conserved

regions (Lougheed et al. 1999; Austin et al. 2002; Vredenburg et al. 2007). In contrast to

the mtDNA data, Quebradillas is quite differentiated from the southern populations at

nuclear microsatellite loci. It possesses far more private alleles than either Guanica or

Tamarindo, and shares more than a single allele with them at only one of the six loci.

This abundance of private alleles in Quebradillas is likely not an artifact of sample size,

as one would expect the populations with larger sample sizes (i.e. Guanica or Tamarindo)

to detect a greater number of rare or unique alleles. Furthermore, STRUCTURE clearly

distinguished the northern individuals, and the measures of genetic distance and

differentiation are much greater between Quebradillas and both southern populations than

between the two southern populations. These Fs1 values of> 0.3 are quite large

compared to some other amphibian microsatellite studies (Newman and Squire 2001;

Brede and Beebee 2004; Johansson et al. 2005), particularly given the small geographic

distance separating the two populations (approximately 55 km). However, some studies

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 49 -

have found similar or greater values over very short geographic distances (Kraaijeveld­

Smit et al. 2005; Arens et al. 2006), and several report values> 0.4 across the distribution

of a single species (Rowe et al. 1998; Bums et al. 2004; Funk et al. 2005).

For small animals with low vagility such as amphibians, it is often not the

geographic distance that determines levels of gene flow between populations, but rather

the ecological and/or geographical barriers that separate them such as mountain ridges

(Funk et al. 2005; Kraaijeveld-Smit et al. 2005; Martinez-Solano et al. 2005), unsuitable

habitat (Johansson et al. 2005), or even ancient geological features that are no longer

apparent (Lougheed et al. 1999). The Cordillera Central mountain range, with a

maximum elevation> 1300 m (Pico 1974), extends across the interior of Puerto Rico in

an east-west direction (Figure 2.1). The strong differentiation between northern and

southern populations of P. lemur is thus not unexpected, as this species is restricted to

low-elevation coastal areas (Rivero et al. 1980). The genetic data, combined with P.

lemur's historical distribution and probable inability to traverse the mountainous interior,

suggests that northern and southern populations have likely not experienced gene flow for

a substantial time.

Conservation and management recommendations

An important finding of this study is that the breeding of P. lemur in AZA

(Association of Zoos and Aquariums) institutions has been very successful in terms of

genetic management: the captive Guanica population has not suffered reduced genetic

diversity relative to the wild Tamarindo population, nor has it diverged substantially

during its isolation. Although only neutral molecular markers were evaluated, this

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 50 -

suggests that adaptation to captivity has been minimal (e.g. Ford 2002) and that the

captive southern population is suitable for continued reintroduction to the south coast of

Puerto Rico. Additional supplementation of the captive colony by wild individuals does

not appear to be immediately necessary for genetic purposes.

The future of P. lemur in northern Puerto Rico, however, is in jeopardy.

Although the Quebradillas breeding colony possesses adequate variation at this time, it

likely cannot be sustained much longer as the 123 individuals currently in captivity are

descended from four inbred siblings. It was previously believed that the colony had

reached reproductive senility as no clutches had been obtained for some time (Bloxam

and Tonge 1995). While this is likely in part due to the age of the individuals, it is also

possible that inbreeding avoidance (Waldman et al. 1992) and/or depression have

contributed to lower fitness. Successful breeding of northern animals has occurred

recently, with releases of tadpoles on the north coast of Puerto Rico. However, it is

unlikely that individuals have become established since it required nearly 20 years and

over 100 000 released individuals before breeding animals were observed in the south.

This indicates that there will be limited opportunities to restock the captive northern

colony should it become depleted. The prospects for the northern lineage of P. lemur are

therefore dire and it is no longer sufficient to maintain separate populations in captivity;

in order to ensure the persistence of P. lemur in the north, less conventional management

options must be explored.

The presence of an exclusive mitochondrial haplotype and many unique

microsatellite alleles in Quebradillas suggests that the northern and southern populations

of P. lemur are quite distinct and have not experienced gene flow for some time. This

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 51 -

isolation may have enabled unique, adaptive characteristics to evolve in the north,

facilitated by differences in habitat and climatic conditions: Guanica is located in an arid

coastal plain, receiving ~ 3 0-3 5 inches of rainfall annually and possessing dry limestone

vegetation, while Quebradillas receives 40-60 inches of annual rainfall and consists of

limestone hills with humid limestone vegetation (Pico 1974). Any unique adaptations

that northern individuals may possess have the potential to greatly assist with the

establishment of reintroduced individuals to this region and should be preserved. It is

therefore recommended that whilst the Quebradillas breeding colony is successful in

producing offspring for release, the Guanica and Quebradillas colonies should continue to

be maintained separately and animals released into their respective regions. In order to

ensure the continuation of the northern lineage, however, a third breeding colony

consisting of both northern and southern individuals should be created. This will increase

the genetic diversity and number of breeding individuals for the Quebradillas colony,

thereby increasing its health and longevity. Initially, this would serve as an experimental

colony to determine whether mixed individuals possess greater or diminished fitness

prior to their release into the wild. In the event that the pure Quebradillas lineage

becomes extirpated, these animals- could then be used for release to the northern coast of

Puerto Rico. This strategy would preserve the most overall genetic diversity, including

any unique genetic adaptations that may be present in the northern lineage. The

alternative option to release pure Guanica individuals into the north should the

Quebradillas colony become extirpated is less desirable, as they would lack any

potentially adaptive traits that may increase their survival in the conditions found in

northern Puerto Rico.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 52 -

The mixing of distinct lineages is generally avoided because of the possibility of

reduced fitness due to outbreeding depression. However, it "should not be dismissed

lightly as a management option, especially where there is evidence that remnant

populations have reduced viability because of inbreeding depression" (Moritz 1999).

Inbreeding depression is generally considered a more serious problem than outbreeding

depression, and the negative impacts of outbreeding depression may be overemphasised

while the benefits of mixing lineages are overlooked (Frankham et al. 2002). "Genetic

rescue" (e.g., Tallmon et al. 2004) by mixing distinct lineages has successfully reduced

inbreeding depression and increased fitness for several threatened species (Westemeier et

al. 1998; Madsen et al. 1999; Land et al. 2004) and continues to be recommended (Soltis

and Gitzendanner 1999; Tarr and Fleischer 1999; Wyner et al. 1999; Godoy et al. 2004).

Experimentally inbred populations of plants (Richards 2000; Newman and Tallmon

2001) and animals (Bryant et al. 1999; Ball et al. 2000) also experienced increased fitness

when immigrants were introduced. Very low levels of migration (e.g., one individual per

generation) may be sufficient for successful genetic rescue, which minimises genetic

swamping and allows local adaptation to proceed (Hedrick 1995; Newman and Tallmon

2001; Palo et al. 2003; Vila et al. 2003 ). Despite the considerable divergence of northern

and southern P. lemur populations, they do not appear to warrant classification as

separate subspecies. Many other amphibian species are much more differentiated at

mtDNA, and the Fsr estimates fall within those for several other amphibians.

Furthermore, husbandry protocols are the same for the northern and southern captive

colonies, suggesting that critical reproductive and rearing aspects do not differ (Lentini

2000). Although the potential for reduced fitness due to outbreeding depression cannot

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 53 -

be completely excluded, these factors indicate that it should be minimal. In contrast, it is

known that the northern colony of P. lemur is highly inbred. While the extent of

inbreeding depression has not been thoroughly evaluated and can vary within a single

lineage (Pray and Goodnight 1995), it was suggested over a decade ago that the northern

colony may be suffering from reduced fitness (Bloxam and Tonge 1995). Given the

available evidence, inbreeding depression is a much greater concern for the future of the

northern lineage than should be outbreeding depression in a mixed population.

The establishment of a third breeding colony of mixed northern and southern P.

lemur lineages can be viewed as insurance for securing some northern traits, as it is

probable that the pure northern colony will become extirpated due to demographic

stochasticity or inbreeding depression. Because outbreeding depression is due primarily

to the breakup of coadapted gene complexes between divergent lineages, it is probable

that negative effects would not manifest until the F 3 generation when such complexes

have undergone recombination (Tallmon et al. 2004). The mixed colony should therefore

be created immediately so that the fitness of mixed individuals can be monitored for

several generations prior to release into the wild ( e.g. Moritz 1999). Although fitness in

captivity cannot necessarily be extrapolated to the wild, monitoring should provide an

indication of the overall health and reproductive capacity of the interbred colony relative

to the pure northern colony. Sufficient individuals are generally produced in captivity

that these efforts should not affect the existing breeding colonies or release efforts. Its

initiation would also not require a large investment of resources, as several facilities

currently house P. lemur and much is known about its husbandry (Miller 1985; Lentini

2000). Future work should develop models to determine the minimum number of

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 54 -

southern individuals necessary to effect genetic rescue of the northern population while

minimising genetic swamping, as was done for the Florida panther (Hedrick 1995).

Not only will a mixed breeding colony facilitate the establishment of P. lemur

populations in northern Puerto Rico by preserving adaptive traits, its experimental nature

enables these recovery efforts to contribute to the development of reintroduction biology

as a science (sensu Seddon et al. 2007). It can also provide insight into the actual

repercussions of outbreeding depression, allowing future conservation biologists to assess

the potential for adverse effects on a given species. Finally, this study demonstrates that

it is not always possible to follow generalised conservation guidelines for all species;

rather, each species must be evaluated individually and decisions made to maximise the

retention of genetic diversity, thereby maximising the species' prospects for recovery.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 55 -

Chapter 3

Characterisation, Multiplex Conditions, and Cross-Species Utility of Tetranucleotide Microsatellite Loci for Blanchard's Cricket Frog (Acris crepitans b/anchardt)

This chapter has been published with minor modifications as:

Beauclerc KB, Johnson B, White BN (2007) Characterization, multiplex conditions, and cross-species utility of tetranucleotide microsatellite loci for Blanchard's cricket frog (Acris crepitans blanchardi). Molecular Ecology Notes 7: 1338-1341. doi: 10.l 111/j.1471-8286.2007.01874.x

KB Beauclerc performed all laboratory work, data analysis, and wrote the paper. B Johnson provided valuable input and funding for the work. BN White is the principal investigator responsible for the work.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 56 -

Abstract

Seventeen tetranucleotide microsatellite were isolated and characterised for Blanchard's

cricket frog (Acris crepitans blanchardi), an anuran common to the central USA.

Conditions for four multiplex reactions, using 16 of the loci, are also described. These

loci were highly polymorphic when screened in two geographically distant populations,

with 11 to 48 alleles per locus (average= 24.8) and observed and expected

heterozygosities of 0.18-0.97 and 0.17-0.96, respectively. The population from Illinois

had significantly lower genetic diversity than the Texan population. Genetic

differentiation between the populations, as measured by FsT, was moderate, but several

private alleles were found in each. Nine loci were also polymorphic in the related taxon

A. c. crepitans, and seven were polymorphic in A. gryllus. Five loci amplified in all three

taxa. These genetic markers will be useful for population- and species-level

investigations of this widespread group.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 57 -

Introduction

Cricket frogs (genus Acris) are small, non-climbing, aquatic hylids that occur

throughout the central and eastern United States (Conant and Collins 1998). Two species

are recognised, the northern (A. crepitans) and southern (A. gryllus) cricket frog, with

these each further divi.ded into subspecies. Blanchard's cricket frog (A. c. blanchardi) is

the most widely distributed member, ranging from Texas through the central states and

into extreme southwestern Ontario (Conant and Collins 1998). Once extremely abundant

throughout its range, many northern populations of A. c. blanchardi are in decline and it

is now rare or extirpated from several localities (Kellar et al. 1997; Lannoo 1998b;

Hammerson and Livo 1999; Lehtinen 2002). As part of the recovery program for A. c.

blanchardi in Canada, I characterised 17 polymorphic tetranucleotide mircrosatellite loci

for A. c. blanchardi, developed multiplex conditions for them, and screened several

individuals from two widely separated populations. I also tested for cross-amplification

ability in related A eris subspecies and species. These genetic markers will be valuable

for future studies to investigate questions such as the evolutionary relationships among

Acris, hybridisation among sympatric species and subspecies, levels of genetic variation,

structure, and gene flow, and fine-scale landscape analyses.

Materials and Methods

Microsatellite sequence isolation utilised a biotin-enrichment protocol similar to

Hamilton et al. (1999). Briefly, DNA was extracted from two individuals by a standard

phenol-chloroform method. Five micrograms of genomic DNA from each individual

were pooled and digested with HaeIII, of which 500 ng was ligated to SNX linkers. One

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 58 -

hundred nanograms oflinker-ligated DNA was enriched with a biotin-labelled probe

containing the repeat motif (GATA)8 and isolated by binding to streptavidin-coated

magnetic beads (Dynal Biotech). Enriched DNA was amplified via the polymerase chain

reaction (PCR) using SNX linkers as primers to double strand the DNA and add an 'A'

overhang. This product was ligated into the pCR2.1-TOPO vector and transformed into

TOPl0 chemically competent cells following the manufacturer's instructions (TOPO TA

cloning kit, Invitrogen).

Colonies with an insert were isolated, amplified using the forward and reverse

M13 primers, and screened on 1 %agarose gels stained with ethidium bromide. Five

microlitres of amplified product was purified using ExoSAP-IT (USB) and sequenced in

the forward direction using the DYEnamic ET-Terminator cycle sequencing kit on a

MegaBACE 1000 DNA Analysis System (Amersham Biosciences). In total, 139 clones

were sequenced, of which 90 contained a tetranucleotide repeat, with only one duplicate.

Primer pairs were designed manually for 37 clones, and FASTPCR v.3.6.31

(Kalendar 2005) was used to check primer pairs for secondary structure and primer dimer

formation. PCR mixtures contained 5 ng genomic DNA, 1 X PCR buffer, 1.5 mM

MgCh, 0.2 mM each dNTP, 0.3 µMeach primer, and 0.5 U Taq polymerase (Invitrogen)

in 10 µL total volumes. Thermal cycling conditions consisted of 30 cycles of

denaturation at 94°C for 30 sec, annealing at the optimal temperature (see Table 3.1) for 1

min, and extension at 72°C for 1 min, with a final extension of 60°C for 45 min, using an

MJ Research PTC-225 thermal cycler. Loci were screened for polymorphism in 4

individuals from each of three regions (Ohio, Illinois and Texas) by visualising PCR

products on 4% agarose gels stained with SYBR Green I (Molecular Probes). Sixteen

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Table concentration

Multiplex

Multiplex

AcrMs-28

AcrMs-14

AcrMs-34

AcrMs-29

AcrMs-36

AcrMs-17

AcrMs-3

AcrMs-2

Locus

3.1.

Characteristics

1

2

is

(Fa

(Fa=

for

F:GCGGTGCTCCCTGTGAG R:ATGCTGCTATTCTTGGATGAC(H)

R:AATGTTTCAGTGTGAGATGTTC F: R:CCGTAATATCCTAACACGCTGG F:ACATTGCCTGTGATATTTGG(H) R:CCCACTCTTCATGTATCCTAGG

F:AGAATAGTCATTGAGATCTCTTAGC(H) F: F: R:GGACGTCACAAGATATTTCCAGC

R:AGAAGAGCCAAATGATAACATCC

F: R:GGAGGACGTAACACATAACTTCC F: R:ATTTGGGTGCTCAGAACAATGG

=

TACAGTGCCAATGTGGGC

AAGTTGCCTGTCACTTTAAA

TCAGGCAGTTGTCAGAGTC

TGTTCCAGGA

GAGTGCTGCAGTGATTATTTGC

the

55

55

concentration

°C)

°C)

of

tetranucleotide

Primer

TCCCGTTCG

{5'-3't

of

sequence

both

microsatellite

forward

(F)

(N)

(H)

TGG

and

(F)

reverse

(F)

loci

and

primers

multiplex

(GATA)16(GACA)1s

(CAGA)s(GATA)10

(GACA)14(GATA)1 (CAGA)

(GACA)3(GATA)14

(GACA)7(GATA)29

(GGCA)2

Re~eat

(GATA)33

(GATA)16

imperfect (GATA)7

imperfect

imperfect

in

the

3

amplification

imperfect motif

imperfect

multiplex

b

reaction.

345-634

297-458 134-258

520-616 225-302

112-219

conditions 183-683

102-375

(bp)

Size

50-65

50-55

50-65 60-65

55-65

for 50-65

(°Ct

60

Ta

50

A.

c.

Primer

blanchardi.

(µM)

0.1

0.1

0.4

0.6 0.5

0.1

0.3

0.2

Accession EF450231

EF450240

EF450230

EF450236

EF450244

EF450241

EF450237

EF450246

GenBank

Primer

-

59

-

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

-

60 60

-

agarose agarose

repeat repeat

EF450239 EF450239

EF450234 EF450234

EF450245 EF450245

EF450235 EF450235

EF450242 EF450242

EF450232 EF450232

EF450243 EF450243

EF450233 EF450233

EF450238 EF450238

on on

the the

respectively. respectively.

based based

0.5 0.5

0.4 0.4

0.2 0.2

0.6 0.6

0.6 0.6

0.2 0.2

0.3 0.3

0.25 0.25

NIA NIA

NED, NED,

interrupted interrupted

cleanly, cleanly,

and and

50 50

65 65

55-60 55-60

55-60 55-60

45-60 45-60

55-65 55-65

55-65 55-65

55-65 55-65

55-60 55-60

pairs pairs

and and

6-FAM 6-FAM

base base

148-249 148-249

245-259 245-259

228-522 228-522

282-506 282-506 194-590 194-590

178-248 178-248

206-299 206-299 154-228 154-228

308-361 308-361

HEX, HEX,

more more

successfully successfully

to to

or or

one one

refer refer

that that

amplified amplified

which which

A)11(GACA)z2 A)11(GACA)z2

im2erfect im2erfect

imperfect imperfect

(GATA)9 (GATA)9

imperfect imperfect

(GATA)9 (GATA)9

imperfect imperfect

imperfect imperfect

(GATA)22 (GATA)22

(GATA)3s (GATA)3s

(GATA)23 (GATA)23

(GATA)16 (GATA)16

(GATA)12 (GATA)12

(GATA)31 (GATA)31

locus locus

indicates indicates

(GATA)s(GACA)3 (GATA)s(GACA)3

(GAT (GAT

single single

parentheses, parentheses,

the the

(F) (F)

(F) (F)

(F) (F)

(F) (F)

Nin Nin

"Imperfect" "Imperfect"

(N) (N)

which which

(N) (N)

(N) (N)

and and

at at

F, F,

clone. clone.

H, H,

by by

ATCAATTGC ATCAATTGC

original original

temperatures temperatures

sequences. sequences.

TCATTCGCCC TCATTCGCCC

of of

from from

indicated indicated

TGTACATACAGTGC TGTACATACAGTGC

are are

I. I.

TCTGACACTTACGG TCTGACACTTACGG

range range

the the

sequence sequence

multiplexes multiplexes

Green Green

°C) °C)

°C) °C)

complete complete for

primers primers

on on

in in

55 55

55 55

TCCATCTCTA TCCATCTCTA

TCCTTACAA TCCTTACAA

TCAGTTGATCTGGTTGAATGTCG TCAGTTGATCTGGTTGAATGTCG

TCTCAGACAGACTTTGACTGG TCTCAGACAGACTTTGACTGG

CCCACAA CCCACAA

ACAAACAATAGGAGTCTATGGAGG ACAAACAATAGGAGTCTATGGAGG

= =

A A

TCTCATCGGACTTGT TCTCATCGGACTTGT

= =

not not

YBR YBR

S S

based based

R:AATTCAATAATGGGTGCATCATGC R:AATTCAATAATGGGTGCATCATGC

R:CTCAAATCACACCATACCTGCC R:CTCAAATCACACCATACCTGCC

F: F:

F: F:

F: F:

R:CCTGGATCTTACTAAGGAGTGC R:CCTGGATCTTACTAAGGAGTGC

R: R:

F:AATCCTGATGAGCAAGAACC(H) F:AATCCTGATGAGCAAGAACC(H)

R:ACTCTACAACCAGCAATCAG R:ACTCTACAACCAGCAATCAG F:GACAAAGTGAAGATGGTAAACTACC F:GACAAAGTGAAGATGGTAAACTACC

F: F: R:TAAGAAGACCTCCAGCCACC R:TAAGAAGACCTCCAGCCACC

R: R:

F:GAAGGTCAGTTGTGTAGACTAGC F:GAAGGTCAGTTGTGTAGACTAGC

R:TCCGTGGACTGGCTCATGG R:TCCGTGGACTGGCTCATGG

R:CGGGGCGATCATTGGATAAATAC R:CGGGGCGATCATTGGATAAATAC

F: F:

F:TCACCTTTCTTTTGGAGTTCTGC(H) F:TCACCTTTCTTTTGGAGTTCTGC(H)

labelled labelled

(Ta (Ta

(Ta (Ta

is is

loci loci

GenBank GenBank

4 4

3 3

with with

is is temperature

See See

motif motif

stained stained

AcrMs-9 AcrMs-9

AcrMs-4 AcrMs-4

AcrMs-8 AcrMs-8

AcrMs-26 AcrMs-26

AcrMs-12 AcrMs-12

AcrMs-35 AcrMs-35

AcrMs-31 AcrMs-31

AcrMs-32 AcrMs-32

Additional Additional

Multiplex Multiplex

AcrMs-24 AcrMs-24

Multiplex Multiplex

Repeat Repeat

Annealing Annealing

gels gels

c c

b b

aFluorescently aFluorescently motif(s). motif(s).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 61 -

loci amplified reliably and were found to be polymorphic (Table 3.2). One additional

locus (AcrMs-26) amplified well, but was not used in further analyses due to low

polymorphism and difficulty incorporating into multiplex reactions. Sequences for

clones are deposited in GenBank under accession numbers EF450230 to EF450246.

The 16 loci were organised into four multiplex reactions of four loci each (Table

3.1 ). Initial primer concentrations were 0.3 µM for all loci; these were subsequently

adjusted to balance peak heights. Interaction between primers for AcrMs-1 7 and AcrMs-

34 in Multiplex #2 produced strong nonspecific peaks that interfered with scoring; as

such, this multiplex was amplified as two separate reactions consisting of AcrMs-

29/AcrMs-17 / AcrMs-36 and AcrMs-34, which were pooled in the ratio 2:1 prior to

desalting. In Multiplex #3, AcrMs-32 and AcrMs-4 amplified weakly; this multiplex was

therefore also amplified as two reactions consisting of AcrMs-8/AcrMs-24 and AcrMs-

32/AcrMs-4, which were pooled in the ratio 1:2 prior to desalting. PCR products were

desalted using Sephadex G50 (Sigma Aldrich), profiled on a MegaBACE 1000 DNA

Analysis System, and analysed using GENETIC PROFILER v.2.2 (Amersham Biosciences).

Twenty-two individuals from a single pond in Champaign County, Illinois and 33

individuals from a single pond in Kendall County, Texas were screened at the 16 loci

using the multiplexes described in Table 3.1. All summary statistics and molecular

indices were calculated using GENEPOP v.3.4 (Raymond and Rousset 1995). Because of

the large geographic distance between the two populations and the likelihood that they

represent independent breeding populations, statistics were calculated for each region

separately. Genetic differentiation was assessed using the number of private alleles and

FsT according to Weir and Cockerham (1984) in GENEPOP.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 62 -

Table 3.2. Genetic diversity indices for 16 tetranucleotide microsatellite loci in two populations of A. c. blanchardi. Number of alleles (Na) and expected (HE) and observed heterozygosities (Ho) were determined using 22 individuals from Illinois and 33 from Texas. Deviation from Hardy-Weinberg equilibrium is indicated by P < 0 05. Allelic richness (ag) was determined for the Texas population by standardisation to N = 22 using rarefaction.

Illinois Texas Locus Na HE Ho p Na HE Ho p ag Total Na 3 7 0.73 0.73 0.67 19 0.93 0.91 0.35 16.63 20 28 13 0.91 0.91 0.02 23 0.96 0.94 0.19 23.34 34 2 11 0.86 0.77 0.79 17 0.94 0.52 0 16.00 20 14 8 0.81 0.77 0.44 33 0.96 0.79 0 25.12 32 34 4 0.40 0.41 1 10 0.81 0.36 0 9.31 11 29 8 0.83 0.36 0 16 0.92 0.81 0.06 14.99 18 17 13 0.87 0.91 0.94 30 0.96 0.85 0 24.23 37 36 8 0.77 0.86 1 24 0.96 0.88 0.27 21.03 27 8 3 0.41 0.41 0.58 15 0.92 0.91 0.77 14.22 15 24 3 0.50 0.36 0.08 18 0.91 0.94 0.85 15.78 18 32 7 0.77 0.59 0.03 19 0.92 0.55 0 16.58 21 4 20 0.95 1 1 29 0.96 0.94 0.67 23.71 48 31 10 0.85 0.86 0.15 17 0.93 0.94 0.37 15.19 20 35 8 0.76 0.82 0.50 16 0.90 0.91 0.6 14.02 16 9 12 0.87 0.37 0 31 0.96 0.97 0.63 24.91 38 12 2 0.17 0.18 1 22 0.93 0.52 0 18.38 22 Total 137 339 293.41 399 Average 8.56 0.72 0.65 21.19 0.93 0.79 18.34 24.81

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 63 -

Results and Discussion

For Illinois, 137 alleles were found (allelic diversity= 8.56) and the number of

alleles for each locus ranged from two to 20 (Table 3.2). Observed and expected

heterozygosities ranged from 0.18 to 0.91, and 0.17 to 0.95, respectively. Texas was

substantially more diverse, with 339 total alleles (allelic diversity= 21.19), and 10 to 33

alleles per locus (Table 3.2). Observed and expected heterozygosities were also generally

higher than Illinois, ranging from 0.52 to 0.94 and 0.82 to 0.96, respectively. Although

the discrepancy in sample size between these populations is not great, it could contribute

to the differences in diversity. To determine whether genetic diversity truly varied

between these populations, I calculated allelic richness (ag) for Texas using rarefaction in

the program HP-RARE v.1.0 (Kalinowski 2005). This value is the expected number of

alleles when the sample size is standardised to be equal to that of the Illinois population

(N = 22) (Kalinowski 2004). Results showed that Texas still possessed substantially

more alleles at each locus than Illinois (Table 3.2), and the null hypotheses that the

number of alleles (ag for Texas) and HE were equal between the two populations were

both rejected using a 2-tailed Wilcoxon signed-rank test in SPSS v.10 (P < 0.001).

Genetic differentiation, as measured by total FsT, was moderate between Texas

and Illinois (Table 3.3). However, all loci in the Texas population contained private

alleles, which is not surprising given its high polymorphism, and even the much less

diverse Illinois population possessed private alleles at 75% ofloci (Table 3.3). Private

allelic richness (ng) was also much greater in Texas following standardisation to N = 22

using rarefaction in HP-RARE (Table 3.3).

Fisher's exact test showed that loci AcrMs-9, AcrMs-28 and AcrMs-29 departed

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 64 -

Table 3.3. Number of private alleles [p(l)] and FsT values for 16 microsatellite loci in two populations of A. c. blanchardi. Private allelic richness (ng) was determined for the Texas population by standardisation to N = 22 using rarefaction.

Illinois Texas Locus p(l) p(l) 7l'g FsT 3 1 13 11.44 0.13 28 5 21 16.04 0.04 2 3 9 8.19 0.04 14 1 24 19.05 0.07 34 1 7 6.45 0.34 29 2 10 9.11 0.08 17 7 24 19.70 0.07 36 3 19 16.84 0.12 8 0 12 11.69 0.29 24 0 15 12.89 0.21 32 2 14 12.02 0.10 4 19 28 23.04 0.04 31 3 10 8.64 0.06 35 0 9 7.84 0.10 9 8 26 21.05 0.07 12 0 20 16.71 0.32 Total 55 261 Average 3.44 21.19 13.79 0.13

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 65 -

significantly from Hardy-Weinberg equilibrium in Illinois, as did loci AcrMs-2, AcrMs-

12, AcrMs-14, AcrMs-17, AcrMs-32 and AcrMs-34 in Texas (Table 3.2). All were due

to heterozygote deficiency. This may result from a number of factors, including

population substructure (Wahlund effect), inbreeding in small populations, selection

against heterozygotes, and null alleles (Frankham et al. 2002). Given that samples were

collected from single ponds, it seems unlikely that such large deficiencies could be the

result of substructure within each population; moreover, a deficiency of heterozygotes

would be expected at all loci. It is possible that highly related individuals (i.e. family

groups) were sampled, essentially creating very fine-scaled structure. However, the

occurrence of up to 20 alleles for a single locus in Illinois and 33 in Texas indicates that

at least 10 and 17 parents, respectively, are responsible for producing the samples

profiled here. This also argues against the effects of small population size, as does the

observation that populations in these regions can have > 1000 individuals at some ponds

(Pyburn 1958; Nevo 1973b; Gray 1983, 1984). It also seems unlikely that selection

against heterozygotes is causing the deficiency, as microsatellites are generally not

known to produce a functional protein product. While it is possible that the affected loci

are linked to genes that may be under selection, it seems improbable that this would be

the case for 25-38% ofloci (Table 3.2). The high variability of these markers and the

difficulty in designing primers that successfully amplified individuals from both regions

(K. Beauclerc, unpublished data) suggest that mutations in priming sites are very likely to

occur in some individuals, and that null alleles are therefore the probable cause of the

heterozygote deficiency. After the sequential Bonferroni correction was applied, locus

pairs AcrMs-31/AcrMs-35 and AcrMs-2/AcrMs-12 continued to show linkage

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 66 -

disequilibrium in Texas. While this may indicate that these loci are physically linked, the

fact that they were in equilibrium in Illinois suggests that they are not.

The loci developed for A. c. blanchardi were screened for amplification ability in

related species and subspecies: the northern cricket frog (A. c. crepitans), the southern

cricket frog (A. gryllus gryllus), and the Florida cricket frog (A. g. dorsalis). When

screened on agarose gels, the following loci either amplified poorly, not at all, or with

several nonspecific bands: AcrMs-2, AcrMs-9, AcrMs-12, AcrMs-14, AcrMs-31 (all

three taxa), and AcrMs-3, AcrMs-4, AcrMs-24, AcrMs-35 (A. gryllus). AcrMs-36

amplified very weakly in A. g. gryllus, and appeared to have null alleles in A. g. dorsalis.

AcrMs-4 and AcrMs-28 amplified well in A. c. crepitans, but had alleles> 900 bp and

thus were discarded from use. AcrMs-32 generally amplified well in A. c. crepitans, but

was also discarded due to suspected null alleles. In summary, only a subset of the loci

amplified reliably in each taxa (Table 3.4): nine loci in A. c. crepitans and seven loci in

A. gryllus. Five loci amplified in all four Acris subspecies examined here. All loci that

amplified were polymorphic in seven (A. c. crepitans) or nine (A. gryllus) individuals.

The highly polymorphic nature of these microsatellite loci in A. c. blanchardi, and

their utility in related taxa, will make them valuable for a variety of applications. These

preliminary analyses have already demonstrated genetic differentiation between northern

and southern regions, as well as substantial differences in genetic diversity, which is

concordant with mitochondrial data (Chapter 4) and colour polymorphisms (Nevo 1973b;

Gray 1983; Gorman 1986). The presence and frequency of null alleles is an issue that

must be thoroughly investigated, as statistical adjustments may be required to account for

them (Chapuis and Estoup 2006; Van Oosterhout et al. 2006; Wagner et al. 2006).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Table

and agarose

individuals.

species.

3.4.

gels

AcrMs-17 AcrMs-24

AcrMs-29

AcrMs-26 AcrMs-28

AcrMs-32 AcrMs-34

AcrMs-36

AcrMs-35

AcrMs-3

AcrMs-8

Cross-amplification

Locus

stained

Annealing

with

temperature

SYBR

Ta

45-60

45-55 55-60

45-50 55-65

50-60

50-60

50-60

50

(°C)

ability

Green

A.

c.

(Ta)

Size

crep_itans

245-260

375-591

229-262

171-217 180-281 120-690 119-158

132-225

165-193

of

I.

is

tetranucleotide

Number

(bQ)

the

range

of

Na

12

4

6 5 3 8

6 9

5

of

allele~

temperatures

microsatellite

(Na)

Ta

45-50 50-60 45-55

50-60 50-65 50-65 50-55

was

(°C)

at

calculated

A.

loci

which

g.

Size

243-500 245-258 506-606

224-360 302-382 111-206

154-209

isolated

gryllus

the

(bQ)

using

locus

from

nine

Na

amplified

11

11

12

11

6

2

5

A

(A.

c.

blanchardi

gryllus)

Ta

successfully

45-50

45-50

50-65 55-65 50-60 50-60

50

(°C)

and

A.

in

seven

related

g.

Size

244-258

232-615 225-686 520-600 308-399 102-260

122-202

dorsalis

and

(A. (bQ)

A

cleanly,

eris

c.

crepitans)

subspecies

Na

12

10

14

11 10

11

3 based

-

on

67

-

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 68 -

Chapter 4

Phylogeography and Spatial Structure of Blanchard's Cricket frog (Acris crepitans blanchardl): Conservation Issues at the Northern Edge of its Range

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 69 -

Abstract

Blanchard's cricket frog (Acris crepitans blanchardi) is relatively widespread and

abundant throughout the central and southern United States, but enigmatic declines across

the north of its range have stimulated much interest in this small treefrog. To determine

the distinctiveness and conservation value of the northern populations and evaluate

potential management actions, I investigated its spatial genetic structure and

phylogeography using mitochondrial control region sequences. Analysis of 479

individuals found 101 haplotypes, but relatively low nucleotide diversity. Two

moderately divergent clades were present. One was restricted to the southwest region,

which was likely a refugium during the Pleistocene as it has retained much genetic

diversity and appears to be at demographic equilibrium. The other occurred primarily

across the north, where low genetic diversity, unimodal mismatch distributions, starburst

genealogies, and significant neutrality tests provide evidence of recent postglacial

colonisation. Spatial autocorrelation showed that significant genetic structure occurs up

to a distance of 450 km. The divergence of northern populations indicates that

conservation efforts are warranted in this region, and, when combined with known life

history traits, suggests that adaptive differences may exist. For this reason, any future

supplementations or reintroductions should occur within the geographic boundaries of the

observed genetic structure.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 70 -

Introduction

Global declines in amphibian populations have been at the forefront of the

scientific and public media for almost two decades (Barinaga 1990; Houlahan et al. 2000;

Collins and Storfer 2003; Stuart et al. 2004; Sodhi et al. 2008). Causes of these declines

are numerous, and include obvious factors such as overexploitation, introduced species,

and habitat degradation, but also subtle ones such as climate change, UV radiation,

chemicals, and infectious diseases (Collins and Storfer 2003). It is probable that the

precise cause of declines varies between locations, and that multiple stressors are

interacting in most situations (Collins and Storfer 2003). Much attention has focused on

declines in tropical regions, where there is more amphibian biodiversity and species are

more narrowly distributed (Duellman and Trueb 1986; Myers et al. 2000), where there is

a greater variety and complexity of life histories (Duellman and Trueb 1986), and where

significantly more declining species exist and causes of declines have been particularly

enigmatic (Stuart et al. 2004).

Unlike many of these threatened amphibian species, Blarichard's cricket frog

(A eris crepitans blanchardi) has a broad North American distribution (Figures 4.1, 5 .1 ),

is abundant throughout much of the Great Plains, and has an unremarkable life history.

However, populations have been in decline along the northern and western edge of its

range (Lannoo et al. 1994; Hay 1998; Mierzwa 1998; Minton 1998; Lehtinen 2002;

McLeod 2005; Lehtinen and Skinner 2006), and it is likely extirpated from Canada and

several northern US states (Kellar et al. 1997; Moriarty 1998; Hammerson and Livo

1999). Once considered to be the most common amphibian in many parts of the upper

Midwest (Lannoo et al. 1994; Hay 1998; Mierzwa 1998; Gray and Brown 2005), it is

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 71 -

lu !!!I

!!II IC!

!ill • ll'l

O Clade I El Clade II A Clade I & II

N 0 500 1000 Kilometres A

Figure 4.1. Location of sampling sites for A. c. blanchardi. The generalised glacial limits for the pre-Illinoian (dotted line) and Wisconsinan glaciations (solid line) are also shown (Reed and Bush 2005).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 72 -

now the "species of most concern" (Lannoo 1998a, p. 430). Like many tropical species,

however, the specific causes of declines in this anuran are unclear. While habitat loss has

likely contributed, it does not appear sufficient to account for the magnitude and

distribution of the declines (Lannoo 1998b; Knutson et al. 2000; Gray and Brown 2005).

Several additional hypotheses have also been postulated, including agricultural chemicals

(Knutson et al. 2000; Gray and Brown 2005; Reeder et al. 2005), habitat acidification

(Gosner and Black 1957; Sparling et al. 1995), heavy metal contamination (Diana and

Beasley 1998), introduced bullfrogs [Rana (Lithobates) catesbeiana] (Burkett 1984; Gray

and Brown 2005), morphological deformities (Mccallum and Trauth 2003), and

infectious diseases or parasites (Beasley et al. 2005; Steiner and Lehtinen 2008).

However, other studies have found these individual factors to be relatively insignificant

and potentially conflicting (Boyd et al. 1963; Gray 2000; Russell et al. 2002; Beasley et

al. 2005; Lehtinen and Skinner 2006). It is probable that the precise cause of decline

differs among locations, and that multiple threats interact in a complex manner (for

details on such scenarios, see Lannoo 1998b; Beasley et al. 2005; Irwin 2005).

Concern for the fate of A. c. blanchardi has recently increased, and has prompted

a variety of ecological (McCallum and Trauth 2003; Gray and Brown 2005; Irwin 2005;

Lehtinen and Skinner 2006; Steiner and Lehtinen 2008) and taxonomic (McCallum and

Trauth 2006; Rose et al. 2006; Gamble et al. 2008) investigations. However, little is

known about the distinctiveness and conservation value of the declining northern

populations. The large range of A. c. blanchardi and highly structured nature of many

amphibian species (Shaffer et al. 2000; Palo et al. 2004; Zeisset and Beebee 2008)

suggest that cryptic genetic diversity is likely to be found for this anuran. Indeed, several

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 73 -

North American amphibians with similarly large ranges have been shown to contain

multiple genetic lineages (Austin et al. 2002; Church et al. 2003; Zamudio and Savage

2003; Hoffman and Blouin 2004; Lee-Yaw et al. 2008). Furthermore, peripheral

populations such as those at the northern edge may be a source for speciation, as isolation

and genetic drift can facilitate divergence from the core populations in response to local

selection pressures (Lesica and Allendorf 1995; Garcia-Ramos and Kirkpatrick 1997).

The fluctuating conditions to which they are subjected may also allow them to evolve

greater resistance to environmental extremes, enabling them to endure changing climates

(Lesica and Allendorf 1995; Volis et al. 1998). Molecular phylogenetic analyses have

also provided support for recognition of A. c. blanchardi as a distinct species (Chapter 5;

Gamble et al. 2008), which means these northern populations may be even more critical

to preserving the overall genetic diversity of this anuran.

In addition to evaluating the taxonomy and distinctiveness of endangered

populations or species (e.g. Wang et al. 1999, 2008; Shaffer et al. 2004a,b), genetic

information has been widely used in conservation. It can provide details on population

connectivity and structure that is crucial to proper management (Bums et al. 2007; Amato

et al. 2008). It may identify populations with limited genetic variation, and thus reduced

evolutionary potential, that are at greatest risk of extirpation from anthropogenic

disturbances (Bos and Sites 2001; Vianna et al. 2006). Genetic surveys can identify the

populations most suitable for use in translocation or reintroduction efforts (Shaffer et al.

2000, 2004a; Wilson et al. 2008). Source individuals should be genetically and

adaptively similar to reduce outbreeding depression and increase survival potential

(Frankham et al. 2002), yet genetically diverse to minimise swamping of native

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 74 -

individuals (Friar et al. 2000) and future inbreeding depression (Frankham et al. 2002).

The limited resources available for conservation means that actions to protect

globally common but locally rare species such as A. c. blanchardi are often questioned

(Hunter and Hutchison 1994). As part of the recovery program for A. c. blanchardi in

Canada, I have investigated the phylogeography and spatial genetic structure of this

subspecies across its range using mitochondrial DNA (mtDNA) sequences. I will use this

information to determine if the declining northern populations are sufficiently distinct to

warrant conservation efforts, identify potential factors that may be contributing to

declines, and provide a framework for recovery recommendations for populations in

Canada and other regions of concern across its northern periphery.

Materials and Methods

Sample collection and molecular methods

Samples were collected from across the entire range ofA. c. blanchardi. Counties

were considered to be sampling sites; the majority(> 75%) of sites consisted of a single

location, and nearly all (95%) comprised no more than two locations. In total, 479

individuals were sampled from 107 sites in 13 states (Figure 4.1; Table 4.1). The number

of samples per site ranged from one to 30, with an average of 4.5. Geographic locations

of the sites were taken as the latitude and longitude for the county in which the pond or

stream was located, and were obtained from the U.S. Census Bureau Census 2000

Gazeteer (http://www.census.gov/; Appendix 1). Tissue samples were collected from

adults between 2001 and 2007 except Roger Mills, Cherokee, Wagoner, Custer, Logan,

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 75 -

Table 4.1. Summary of sampling collection sites for A. c. blanchardi. Additional details on sampling sites and haplotypes can be found in Appendices 1 and 2.

State N No. sites No. haplotypes Arkansas 22 8 19 Illinois 65 5 15 Indiana 31 7 9 Iowa 31 7 4 Kansas 47 23 20 Kentucky 3 2 2 Michigan 27 2 1 Missouri 41 11 14 Nebraska 6 6 2 Ohio 91 14 3 Oklahoma 42 12 22 South Dakota 20 2 2 Texas 53 8 23

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 76 -

and Canadian, OK, which were collected in 1998 as part of a previous study. All tissue

samples consisted of either toe clips collected in the field or muscle from vouchered

specimens, and were preserved in 95% ethanol or IX lysis buffer (2 M urea, 0.1 M NaCl,

0.25% n-lauroyl sarcosine, 5 mM CDTA, 0.05 M Tris-HCl pH 8).

Approximately 10 mg of tissue were dissolved in 500 µL of IX lysis buffer and

digested with 50 µL of proteinase K. Total genomic DNA was extracted with the

DNeasy Tissue Kit (Qiagen Inc.) or an automated magnetic bead procedure, and

quantified with the PicoGreen Quantitation Kit (Molecular Probes) on a FLUOStar

Galaxy (BMG Labtech). A fragment of the mitochondrial control region was amplified

using the primers ControlWrev-L/ControlP-H (Goebel et al. 1999) in a 25 µL reaction

containing 1-5 ng DNA, IX PCR buffer, 2 mM MgCh, 0.2 mM dNTPs, 0.1 mg/mL BSA,

0.2 µM each primer, and 1.25 U Taq polymerase (Invitrogen). Thermal cycling used an

MJ Research PTC-225 thermal cycler with the following conditions: initial denaturation

at 94°C for 5min, 30 cycles of denaturation at 94°C for 30 sec, annealing at 55°C for 30

sec, and extension at 72°C for 45 sec, with a final extension of 72°C for 1O min. PCR

product was visualised on 1 % agarose gels stained with ethidium bromide, and 5 µL of

product was purified with ExoSAP-IT (USB). Cycle sequencing was performed with the

primer ControlP-H using the DYEnamic ET-Terminator kit on a MegaBACE 1000 DNA

Analysis System (GE Healthcare).

Data analysis

Sequences were aligned with CLUSTALX v.1.81 (Thompson et al. 1997) and edited

in BIOEDITV.7.0.5.3 (Hall 1999). A comparison oflog likelihood scores using the Akaike

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 77 -

Information Criterion in MODEL TEST v .3. 7 (Posada and Crandall 1998) found the

Tamura-Nei model with invariable sites and gamma distribution (TN+I+f) to be the most

likely model of sequence evolution for the data (AIC = 3082.09, lnL = -1534.05). The

parameter estimates from this model were: base frequencies: A = 0.31, C = 0.12, G =

0.25, T = 0.33; substitution rate matrix: A-G = 13.75, C-T = 19.90, all others= 1; ratio of

transitions to transversions = 7.42; proportion of invariant sites (I)= 0.74; and gamma

distribution shape parameter (a)= 0.90. This model was used to calculate diversity

indices at the nucleotide level for the entire dataset and for all groups identified during

subsequent analyses. ARLEQUIN v.3.11 (Excoffier et al. 2005) was used to determine the

number of variable sites and haplotype frequencies, and to calculate pairwise sequence

divergences, nucleotide diversity (7rn; average number of pairwise differences), and

haplotype diversity (h; probability that two randomly chosen haplotypes are different).

Nucleotide diversity was calculated without a model of nucleotide substitution for each

site with> 2 samples, and was regressed against latitude and longitude in SPSS v.10 to

evaluate geographical patterns of genetic diversity. The significance of the slope of the

regression line was determined with analysis of variance.

Several features inherent to intraspecific genetic datasets frequently violate the

assumptions of traditional interspecific phylogenetic methods (Posada and Crandall 2001;

Huson and Bryant 2005); network methods are thus often preferred to incorporate the

additional information present in such datasets into meaningful genealogies. As such, a

median-joining network was created in the program NETWORK v. 4.5 (Bandelt et al. 1999)

using the default parameters to explore the relationship among A. c. blanchardi control

region haplotypes.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 78 -

The relatively continuous distribution of A. c. blanchardi and lack of distinct

phenotypic differences (McCall um and Trauth 2006) makes a priori separation of the

sample sites into populations somewhat arbitrary. The genetic data were therefore

analysed with spatial techniques to describe the genetic structure and identify the best

grouping of sites (Diniz-Filho and Telles 2002). Spatial analysis of molecular variance

(SAMOVA) groups geographically adjacent sites into a number of user-defined clusters (K)

for which the proportion of total genetic variance due to differences between groups (FCT

index) is maximised. All 107 sample sites were analysed in the program SAMOVA v.l

(Dupanloup et al. 2002) for clusters of K = 2 to K = 20 (the maximum number of clusters

possible in SAMOV A), with 10 000 iterations of the simulated annealing steps and 100

repetitions of the entire process. Significance of the FCT index was tested by comparison

with the null distribution produced by 1000 simulations. The extent of genetic structure

was also examined using spatial autocorrelation analyses in GENALEX v .6.1 (Peakall and

Smouse 2006). Squared genetic distances (calculated as sequence divergences) and

geographic distances between all samples were calculated, and the correlation coefficient

was determined for distance classes of 50 km. The significance of the correlation

coefficient for each distance class was evaluated using 9999 permutations of the data, and

95% confidence intervals were obtained using 9999 bootstraps. Correlograms were

generated for the entire sample set as well as Groups 1 through 5 identified during

SAMOVA analysis (Groups 6 and 7 consisted of only two locations each and were not

analysed; see Results). The entire correlogram was considered significant if at least one

coefficient remained significant following Bonferroni's correction based on the number

of distance classes (Oden 1984 ).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 79 -

The demographic history of A. c. blanchardi was explored by examining the

frequency of pairwise differences between all samples (mismatch distribution) for the

entire sample set and each group identified by SAMOVA analysis using ARLEQUIN. The

null hypothesis that each observed data set conformed to the expected distribution under

the sudden expansion model was tested using 5000 bootstrap replications. Each group

was further evaluated for conformation to the neutral model of evolution using Tajima's

(1989) D and Fu's (1997) Fs statistics in ARLEQUIN, which are more powerful than

mismatch distributions for detecting population size changes (Ramos-Onsins and Rozas

2002). Coalescent simulations were used to generate 10 000 random samples of each

statistic under the hypothesis of selective neutrality and population equilibrium. The

proportion of random D or Fs statistics less than or equal to the observed value was taken

as the P-value of the statistic. Fu (1997) showed that for the empirical distribution of the

Fs statistic, the critical points at the 5% significance level correspond to the second

percentile; thus, Fs should only be considered significant when P < 0.02.

Results

Quality sequence was obtained for 485 bp of the control region in 4 79 individuals,

yielding 101 haplotypes (Appendix 2). Sequences are deposited in GenBank under

accession numbers EU828245 to EU828345. Seventy-three sites were variable, with 58

transitions, 17 transversions, and six indels. Maximum sequence divergence was 0.036,

and occurred between haplotypes AcrCR-75 (Texas) and AcrCR-95 (Oklahoma).

Nucleotide diversity (nn) for the entire sample set was 0.0089 (0-0.0247 for sample sites

with> 1 individual), and haplotype diversity (h) was 0.77. Nucleotide diversity for each

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 80 -

site decreased significantly in a south to north and west to east direction (Figure 4.2).

Haplotype frequencies ranged from to 0.0021 (singletons) to 0.47 in AcrCR-1. The next

most frequent haplotypes were AcrCR-19 (0.042) and AcrCR-22 (0.031). Many low­

frequency haplotypes were present: 74% occurred once or twice in the sample set.

AcrCR-1 was the most widespread, and was found in ten of 13 states and 53 of 107 sites.

The median-joining network identified two clades that were separated by a

minimum of five mutations (Figure 4.3). These roughly corresponded to the northern

(Clade I) and southwestern (Clade II) portion of A. c. blanchardi' s range (Figure 4.1 ).

Clade I was more widespread, being found at 91 locations in every state sampled. It also

contained the majority of individuals (391) and haplotypes (72), but had relatively low

nucleotide (0.0042) and haplotype diversity (0.66). In contrast, Clade II occurred at 30

locations primarily in Kansas, Oklahoma, and Texas, contained far fewer individuals (88)

and haplotypes (29), but had much greater nucleotide (0.0113) and haplotype diversity

(0.93). The average sequence divergence between the groups was 0.021. Overlap of

haplotypes belonging to the two clades occurred at 14 locations primarily in Kansas and

Oklahoma, but also in Arkansas, Missouri and Texas (Figure 4.1 ). A large "starburst"

with haplotype AcrCR-1 as its centre was present in Clade I of the median joining

network. All but one of these haplotypes differed from AcrCR-1 by a single mutation.

Clustering of the 107 sample sites into 20 groups (K = 20) yielded the largest F CT

value during SAMOVA analysis (Figure 4.4). However, FcT is expected to increase with

increasing K (Dupanloup et al. 2002). F CT appeared to reach a plateau after K = 4,

suggesting that there is no apparent difference between greater values of K. As well, Fsc,

the variation within populations relative to the groups, declined abruptly until K = 7

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 81 -

0.030 y =-0.0013x + 0.0567 2 0.025 • R =0.31 • F =34.52, p < 0.01

0.020 • • • 0.015 - • ·- • • ••• • • • 0.010 • • 0.005 • • • _.>, • "cii.... 0.000 -~Q) 28 30 32 34 36 38 40 42 44 ""C Latitude Q) .::""C 0.030 0 Q) y =-0.0006x- 0.0527 u 2 :::, 0.025 • • R =0.23 z F =23.36, p < 0.01

0.020 • • • • • •• • 0.015 • • • •• •• • • • 0.010

0.005

0.000 +----,---...... ----.----~ .... --~----,--41>---,----_...,.-----r------t -102 -100 -98 -96 -94 -92 -90 -88 -86 -84 -82 Longitude

Figure 4.2. Linear regression of nucleotide diversity with latitude (top) and longitude (bottom) for 79 sample sites across the range A. c. blanchardi. Only sites where two or more samples were profiled are included.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 82 -

I' Clade I I I I I I' I I I I I ' I' ' I' I ' I' I Clade II

I I I I ' I' I I' I' I' I I I I I I'

Figure 4.3. Median joining network of 101 control region haplotypes found in A. c. blanchardi. Open circles are haplotypes, with size of the circle proportional to the frequency of the haplotype in 4 79 individuals. Small black circles are median vectors. All nodes are connected by single mutations, unless otherwise specified by the number of hatch marks. The groups ofhaplotypes comprising the two major clades, designated Clades I and II, are separated by the dashed line. For clarity, only haplotype AcrCR-1 is labelled.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 83 -

0.8

I I 0.7 ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ ■ • 0.6 ■

A 0.5 (.) ~en ~ro 0.4 A -en A A A LL 0.3

A A A 0.2 I A I A A A A A A I A A A A 0.1

0 0 5 10 15 20 25 K

Figure 4.4. The variation observed among groups (FcT; squares) and variation among sample sites within groups (Fsc; triangles) when 107 A. c. blanchardi sample sites are partitioned into K = 2 to 20 groups using spatial analysis of molecular variance. The vertical dashed line indicates K = 7, the optimal number of groups for the data.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 84 -

(Figure 4.4). This value was therefore chosen as the most likely number of groups, since

it maximises the variation among groups (FcT) while minimising the variation among

sites within groups (Fsc) (Dupanloup et al. 2002). Seventy-two of the 107 sample sites

were included in a single group, with the remaining 35 divided among six additional

groups (Table 4.2, Appendix 1). The FsT for K= 7 was 0.75 (P < 0.01), and 68.01% of

the variance was accounted for by the diversity among the seven groups (

6.93% by the variation among sites within these groups (

variation within each of the sites (

widespread, but was the least diverse in terms of Jrn and h (Table 4.2). In contrast, Group

6 contained the fewest individuals and was localised to two counties in south-central

Texas, but possessed the highest 7rn and second highest h. The sites that comprise

SAMOVA Groups 1 and 2 consisted mainly of Clade I haplotypes, while sites belonging to

Groups 3, 4, 6, and 7 were primarily composed of Clade II haplotypes. Group 5

contained a mixture of haplotypes from both Clades I and II, although Clade II

haplotypes were slightly more numerous.

A profile of declining spatial autocorrelation with distance class was obtained for

the entire sample set and Group 1, with significant spatial autocorrelation up to a distance

of approximately 750 km and 450 km, respectively (Figure 4.5). Both correlograms were

significant overall using Bonferroni' s correction. Groups 2 through 5 appeared to have

little overall pattern and the autocorrelation coefficients rarely differed significantly from

random expectations (Figure 4.5).

Significantly small values for neutrality tests in the entire data set are indicative of

a population expansion in A. c. blanchardi' s demographic history (Table 4.3) (Tajima

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Table diversity

Group

Group

Group Group

Group Grour_ Group

4.2.

4 2 3

6

5

1 7

(1rn)

350

Characteristics

20 30

23

31

12

13 N

and

locations

haplotype

No.

72

10

1 2 2 1 8

2 5 2 8

of

diversity

states

o No. No.

groups

12

3

9 1

haploty2es

identified during

(h).

48

22

10

15

9 6

OK:

KS:

Sebastian;

AR:

spatial

KS:

Gove,

Canadian,

Benton,

Cowley,

analysis

Harvey,

OK:

Sequoyah;

Columbia,

Kiowa,

Atoka,

TX:

TX:

Wallace,

Kiowa,

All

of

Jefferson,

McMullen,

molecular

Kendall,

remaining

Locations

Logan,

Leflore;

TX:

Pawnee;

Franklin,

Washington

Hood,

Morris,

Menard

Muskogee,

variance

sites

Travis

TX:

OK:

Garland,

Wise

Anderson,

Osage,

Custer,

for

Roger

Lafayette,

A.

Shawnee,

Ellis

c.

Bowie

Mills,

blanchardi,

0.0094 0.0029

0.0127

0.0064

0.0126

0.0157 0.0044

including

7l"n

nucleotide

0.91 0.58

0.98

0.72

0.91

0.95 0.82

H

-

85

-

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. permutations Figure

blanchardi

4.5.

·u :.:; :E -

(.)

C Q)

Q) 0 m 0 .__ .__ Q)

0 C

0

samples

-0.2

-0.4

-0.1

-0.2

-0.2 -0.4 -0.6 -0.8

0.8

0.6

0.4

0.2 Correlograms

0.3

0.2

0.1

0.8 0.6 0.4

0.2

1.2

(dashed

0

0

1

0

25

25

25

~

·--

225

and

75

lines),

SAMOVA

425

75

------

of

125

and

625

the

95%

correlation

175

Groups

825

125

confidence

1025-

225

1

to

1225

coefficient

175

275

5.

45 1625 1425

Also

intervals

325

shown

All

225

for

Group

Group

Distance

samples

375

1825

around

genetic

....

2

4

is

-

.

--

2025

the

425

275

..

class

the

and

distribution

correlation

-0.1 -0.2

-0.2 -0.4

-0.6

-0.2

-0.3

-0.1

0.4

0.3

0.2

0.1

0.8 0.6 0.4

1.2

0.2

0.2

0.3

0.1

geographic

(km)

0

0

0

25

,------,

25

,------,

25

_,__

-t.·•

L

,------,

-,

1=:i:::1::i::1::T:::1::~::!::=::::1::!:;:~::[::1:_:j

r ·

______

·,

·-.

______

..

125

..

-

···,,.

.... •

_.,,

of

75

75

225

coefficient -

...

..

the

•_•

distance

.....

325

correlation

125

125

,I

425

......

at

175

based

525

distance

175

I ·

......

625

coefficient

225

----···

on

725

225

..

9999

classes

275

825

---·

bootstraps.

based

925

275

2 375 325

of

1025

50

Group

on

Group

Group

325

1125

km

9999

1

(error

3

5

1225

for

___.

__,

425

375

all

bars).

A.

-

c.

86

-

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Table

and of

squared

parameters

4.3.

deviation)

Measures

Group Group Group Group Group

GrouE

Group

of

All

the

4

3 2

7 5

6

1

values

of

mismatch

selective

-1.6656 -2.3030 -0.8969 -1.3321 -0.2639

-0.2696

-0.0902

.02 0.5889 0.1042

were

D

distribution

neutrality

significant

0.0127

0.0849 0.4385 0.2021 0.4398

0.5034

(Tajima's

Neutrality

for

p

0

at

P

the

=

entire

0.05.

D

-24.8904

-27.7578 -11.3716 Tests

-1.0866

-3.6179 -1.1808 -1.9933

0.7917

and

A.

Fs

c.

Fu's

blanchardi

Fs),

0.3129 0.6736

0.1468

0.0713 0.2426

their

0.002

p

0 0

data

significance

set

and

4.7

4.7

3.4 6.6

7.1

2.1

5.9

1.5

x

seven

(var.)

based

(13.4) (17.4)

(16.1) (15.6) (12.6)

( (5.7) (2.3)

4.7)

Mismatch

SAMOVA

on

10

0.0202

0.0290

0.0015 0.0050 0.0097 0.0627 0.0359 0.0191

SSD 000

groups.

Distribution

simulations

None

10.1621 10.0898

0.0801 6.3965

9.2188 9.9551

10.416

2.875

't

of

of

the

the

SSD

data

(P),

(sum

-

87

-

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 88 -

1989; Fu 1997). The bimodal mismatch distribution of pairwise sequence differences

suggests that at least two expansions have occurred, one very recently and one earlier

(Rogers and Harpending 1992). The unimodal mismatch distributions for Groups 1 and 2

indicate these are the groups contributing the signature of demographic expansion to the

entire data set (Figure 4.6). The expansion occurred more recently in Group 1, as

indicated by the smaller mean number of pairwise differences and significant values for

both Tajima's D and Fu's Fs, while the older expansion in Group 2 was detectable only

by the more sensitive Fu's Fs test (Table 4.3). Mismatch distributions for Groups 3 to 7

were somewhat multimodal with little apparent trend and large variances (Figure 4.6;

Table 4.3), suggesting demographic stability (Rogers and Harpending 1992). This was

also supported by the neutrality tests, which were not significant (Table 4.3). None of the

mismatch distributions were significantly different from a model of demographic

expansion based on the sum of squared deviations (Table 4.3), which conflicts with the

shape of the distributions and neutrality tests for Groups 3 to 7. However, a recent study

with a similarly large number of haplotypes and relatively shallow population history also

found discordant results among these parameters (Debes et al. 2008). Such discrepancies

may reflect sensitivity to violation of the infinite sites model by a high frequency of

multiple hits, which is especially common in the control region (Schneider and Excoffier

1999; Debes et al. 2008). Furthermore, statistics based on the mismatch distribution are

less powerful than neutrality tests and may display some unexpected behaviour such as

reduced power with larger sample sizes (Ramos-Onsins and Rozas 2002). Due to these

potential issues, the results of the model of demographic expansion may be unreliable and

only the neutrality tests and graphical mismatch distributions will be considered further.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 89 -

0.3 0.5 All samples 0.4 Group 1 0.2 0.3

0.1 0.2 0.1

0 0 0 2 4 6 8 10 12 14 16 0 2 4 6 8 10 12 14

0.2 0.2 Group 2 Group 3

0.1 0.1

0 0 >, (.) 0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 C: Q) ::::, 0.4 0.3 C" Q) Group 4 Group 5 .... 0.3 LL 0.2 0.2 0.1 0.1

0 0 0 2 4 6 8 10 12 0 2 4 6 8 10 12 14

0.3 0.3 Group 6 Group 7

0.2 0.2

0.1 0.1

0 0 0 2 4 6 8 10 12 0 2 3 4 5 6 7

Number of differences

Figure 4.6. Pairwise mismatch distributions for all A. c. blanchardi samples and the groups identified using spatial analysis of molecular variance. Heavy lines are the observed distribution of pairwise differences, and light lines are simulated distributions using the same parameters.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 90 -

Discussion

Spatial genetic structure

Acris c. blanchardi lacks the deep divergences characteristic of many North

American amphibians (Shaffer et al. 2000; Austin et al. 2002; Monsen and Blouin 2003;

Zamudio and Savage 2003; Hoffman and Blouin 2004; Lee-Yaw et al. 2008).

Nonetheless, it appears to have significant genetic structure. The two clades identified by

the median-joining network principally occur in the northern and southwestern portion of

the range, with only limited overlap in the west-central and southern regions. Sixty-eight

percent of the genetic variation was attributed to grouping of the sample sites into seven

relatively geographically distinct clusters by spatial analysis of molecular variance.

Finally, genetic and geographic distances are significantly and positively correlated up to

a distance of approximately 750 km in the entire sample set, and up to approximately 450

km in Group 1.

The correlograms showed that there is little genetic structure within the groups

located in the southern portion of the range (Groups 2-5; Figure 4.5). This is in

accordance with the limits of genetic structure found in Group 1, as these groups do not

have samples separated by> 450 km. The absence of genetic structure at this spatial

scale could be due to high migration rates within each group making them effectively a

single genetic unit; however, this is not consistent with the widely held view that

amphibians are highly structured over even small distances due to low dispersal

capabilities (Shaffer et al. 2000; Palo et al. 2004; Zeisset and Beebee 2008; but see

Austin et al. 2004; Smith and Green 2005; Bums et al. 2007). Mark-recapture studies

suggest that A. c. blanchardi has limited dispersal capacity (Pyburn 1958; Burkett 1984),

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 91 -

and Ward et al. (1987) found that sites along a single creek were strongly differentiated.

This indicates that the lack of genetic structure may instead be due to very low migration

rates isolating sites from one another, making random genetic drift the primary

determinant of relationships between sites (Ward et al. 1987; Wilson et al. 2008).

However, overlap ofhaplotypes between sites within groups, as well as between groups,

may be evidence that gene flow is occurring to a much greater extent than expected ( e.g.

Smith and Green 2006). Further studies using markers with greater resolution ( e.g.

microsatellites) will help to clarify the extent of gene flow and the cause of limited

genetic structure at these smaller distances.

Phylogeography

The large-scale pattern of mtDNA variation in A. c. blanchardi consists of two

major groups: a northern clade that is characterised by low genetic diversity and a broad,

relatively homogeneous distribution (Clade I), and a southwestern clade with high genetic

diversity, restricted distribution, and finer subdivision (Clade II). Colour morphs (Nevo

1973b), some enzymes (Dessauer and Nevo 1969; Salthe and Nevo 1969), and nuclear

microsatellite markers (Chapter 3) also reflect this pattern of "southern richness-northern

purity", which is most likely the result of rapid colonisation of regions that were glaciated

during the Pleistocene era. Simulations show that exponential reproduction by a few

long-distance dispersants from the populations' edge can rapidly fill available habitat,

while individuals behind the advancing front are limited to logistic growth and contribute

little to the gene pool (Hewitt 1996). This form of colonisation entails a series of

bottlenecks in the direction of the dispersion, generating a gradient of decreasing genetic

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 92 -

variation such as that seen in A. c. blanchardi (Hewitt 1996). Numerous plants and

animals in North America and Europe display similar patterns of genetic variation that

have been shown to be primarily structured by climatic changes during the Pleistocene

(Hewitt 1996, 2004; Lessa al. 2003), and leading-edge dispersal from refugia may be

responsible for the north-south phylogeographic breaks seen near the southern edge of the

ice sheets in some North American species (Hewitt 2000; Swenson and Howard 2005).

Models of long-distance dispersal also show a clustering of genotypes and gradients of

correlation coefficients with distance (Ibrahim et al. 1996), and similar profiles have been

attributed to colonisation in other species (Easteal 1985; Bertorelle and Barbujani 1995;

Williams et al. 2007).

Multiple lines of evidence suggest that Clade II is an old, demographically stable

group, and that the southwestern portion of the range (Texas, Oklahoma and Kansas)

served as a refugium for A. c. blanchardi during the fluctuating climate of the

Pleistocene. Greater genetic diversity and subdivision are generally characteristic of

older, refugial populations (Hewitt 1996; Bernatchez and Wilson 1998; Hoffman and

Blouin 2004; Lee-Yaw et al. 2008), and the observed mismatch distributions and

neutrality tests indicate that no population expansions have occurred recently in Groups 3

to 7, which consist largely of Clade II haplotypes. The lack of spatial genetic structure in

this region also suggests an absence of recent demographic events such as bottlenecks or

expansions. In addition, A. crepitans fossils of various ages have been found from Texas

to southern Kansas since 3.4 Ma (Chantell 1966; Holman 1995), demonstrating that this

region has been continuously inhabited by cricket frogs (likely A. c. blanchardi) for a

substantial time. This is also supported by habitat reconstructions showing that the Great

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 93 -

Plains have consisted of essentially modern grasslands since ~ 1.6 Ma, even at the peak of

climatic deteriorations (Graham 1999). Furthermore, the Great Plains has been identified

as a putative Pleistocene refugium for several amphibian and reptiles (Masta et al. 2003;

Austin et al. 2004; Hoffman and Blouin 2004; Howes et al. 2006; Fontanella et al. 2008).

In contrast to the southwest, the northern region appears to be young and

dominated by recent population expansions. Large "star phylogenies" or starbursts, such

as that found in Clade I, occur when many low-frequency haplotypes differ from the

central, ancestral haplotype (AcrCR-1) by only a single mutation, and are usually

attributed to population expansions (Avise 2000). Significant neutrality tests and a wave­

like mismatch distribution with small mean for Group 1, which largely corresponds with

this starburst, are also indicative of a very recent demographic expansion in this region

(Rogers and Harpending 1992; Fu 1997). It likely required several hundred thousand

years to generate the amount of variation seen in the starburst (maximum distance=

0.0043), suggesting that it was created prior to the last glaciation (the Wisconsinan, 122

Ka to 10 Ka BP; Richmond and Fullerton 1986). The most severe glaciations occurred in

the pre-Illinoian stage (ca. 550 Ka to 860 Ka BP), during which the ice sheet extended as

far south as northeastern Kansas and covered Iowa and Illinois, northern Missouri, and

eastern Nebraska and South Dakota (Richmond and Fullerton 1986; Figure 4.1 ). This

would have eliminated most or all populations north of southern Kansas, placing an upper

age of~ 550 Ka on the starburst. A plausible scenario is that an expansion from the

southwest into northern Kansas, western Missouri, and southern Iowa and Nebraska

occurred during the interglacial preceding, or one within, the Illinoian stage (302 Ka to

132 Ka BP). Glacial ice covered nearly all of Illinois during this stage and would also

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 94 -

have eliminated many populations, but it did not penetrate further west into Iowa and

Missouri (Richmond and Fullerton 1986; Stiff and Hansel 2004). Thus, some of these

newly established populations were likely not eliminated by subsequent glaciations, and

generated the starburst pattern evident today.

Group 2 also shows evidence of a population expansion based on neutrality tests

and mismatch distributions. However, the greater mean of the distribution, greater

nucleotide and haplotype diversity, and its southern distribution indicate that the

expansion occurred much earlier in the history of this species. It is possible that this

group reflects the remnants of a dispersal event that occurred in the early Pleistocene, but

whose northern populations were later eliminated by the very severe glaciations that

occurred during the pre-Illinoian stage.

The most northeastern states have clearly been colonised recently, as Ohio

possesses only three closely related haplotypes and Michigan is fixed for a single

haplotype. In addition, no Pleistocene fossils of Acris exist north of central Kansas

(Holman 1995). Although the Laurentide glacier began its major retreat 20 Ka BP, it

experienced several minor advances and ice surges that could have prevented populations

from expanding into the Great Lakes area (Richmond and Fullerton 1986). Holman

(1995) classifies A. crepitans as a secondary invader, indicating that it would not have

begun colonising southern Michigan until at least 10 600 BP. However, this anuran is

primarily a grassland species that prefers open vegetation (Beasley et al. 2005; Lehtinen

and Skinner 2006). At 10 Ka BP, temperatures were estimated to be 2-4 °C cooler than

present and the southern Great Lakes area consisted of closed canopy, mixed conifer­

northern hardwoods forest (Delcourt and Delcourt 1980; Graham 1999). A. c. blanchardi

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 95 -

may therefore have been excluded from this region due to habitat and perhaps

temperature limitations. By 6000 BP, temperatures were near present values and an

essentially contemporary distribution of vegetation had been established, with more open

oak-hickory and mixed hardwood forest dominating the southern Great Lakes region

(Delcourt and Delcourt 1980; Graham 1999). A. c. blanchardi would then have been able

to penetrate further north and east to occupy western Ohio and southern Michigan, and

therefore likely colonised this region only after ~6000 BP.

Conservation implications

Although the low mitochondrial variation in northern populations of A. c.

blanchardi likely reflects recent colonisation of this region, it may have been exacerbated

in recent years by declining populations and habitat fragmentation (Frankham et al.

2002). While it is not likely to be the primary cause of declines, this limited variation

could be a contributing factor when combined with myriad other assaults on these

populations. For example, amphibian populations have much natural variation in

resistance to environmental contaminants (Bridges and Semlitsch 2001 ), and some

southern populations of cricket frogs apparently have developed resistance to the lethal

pesticide DDT (Boyd et al. 1963). However, the reduced genetic diversity present in

northern A. c. blanchardi populations could prevent them from adapting to similar habitat

alterations (Frankham et al. 2002).

The data presented here suggest that the principal cause of the genetic

discontinuity between the northeast and southwest regions is historical range adjustments

in response to changing conditions during the Pleistocene. Initially, southern individuals

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 96 -

were likely prevented from penetrating into the north simply due to carrying capacity

(Hewitt 1996). They may also have been at a selective disadvantage, however, as the

repeated bottlenecks experienced during colonisation can release additive genetic

variation and enable rapid adaptation to novel environments (Hewitt and Ibrahim 2001 ).

The peripheral nature of the most northern populations may further intensify adaptive

differences, as the isolation and more variable environment they encounter may enable

them to generate unique and potentially vital characteristics for survival (Lesica and

Allendorf 1995; Garcia-Ramos and Kirkpatrick 1997). Indeed, southern populations of

A. c. blanchardi may be active year round and experience two breeding peaks (Blair

1961; Pyburn 1961; Gray 1971; Burkett 1984 ), while northern populations have a single,

short breeding season and have developed overwintering strategies ( Gray 1971; Irwin

2005). There may also be adaptive differences in size, colour, and resistance to

desiccation (Nevo 1973a, b). Northern populations of the common frog Rana temporaria

were found to have significantly faster development when examined over a similar

latitudinal gradient in Europe (Palo et al. 2003). It is therefore possible that the

phylogeographic break observed here is maintained by an extrinsic barrier to gene flow.

The potential for adaptive differences in northern populations suggests that,

despite its abundance in southern areas, conservation efforts are warranted for A. c.

blanchardi in this region ( e.g. Lesica and Allendorf 1995). Furthermore, future climate

changes may continue to negatively impact these peripheral populations (Mehlman

1997), which are already susceptible to declines. Future management actions for this

species must take into account the observed spatial genetic structure and phylogeographic

patterns, and the data presented here can provide guidelines. For example, the x-intercept

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 97 -

of the correlograms for genetic and geographic distance (Figure 4.5) defines the

geographic range of independent samples (Diniz-Filho and Telles 2002); that is, samples

separated by less than this distance are statistically correlated and can be considered a

single genetic unit, whereas samples at greater distances are genetically independent. It

appears that for A. c. blanchardi, samples separated by< 450 km are genetically similar;

movement of animals to extirpated or declining regions should therefore not occur

between sites separated by greater distances. It might be argued that the cause of this

uniformity in Group 1, located in the northern part of the range, is recent shared history

due to postglacial expansion from the southern refuge. However, the fact that the

southern groups did not show significant structure at smaller distances suggests that this

may be the real limit to genetic structure in this anuran. Any movement of animals

should also occur within the same clade, which will minimise the potential for

outbreeding depression between adaptively divergent groups, maintain the evolutionary

trajectory of the northern populations, and increase the survival potential oftranslocated

animals as they are likely to be more genetically and behaviourally suited to the

environment (Moritz 1999). However, source populations with the greatest genetic

variation possible should be used for recovery efforts, which will minimise impacts from

inbreeding and provide adaptive potential for the newly established populations (Moritz

1999). Future work should evaluate the concordance of the mtDNA data with both

nuclear data and external factors ( e.g. elevation, climate, or habitat), as these three

components are not necessarily correlated (Irwin 2002; McKay and Latta 2002; Monsen

and Blouin 2003). Their congruence, however, would suggest the presence of adaptive

differentiation mediated by extrinsic barriers to gene flow, and provide support for the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 98 -

possible taxonomic distinctiveness of northern populations.

Conservation frequently consists of ad hoc, emergency efforts that are executed

when few options are available, and amphibian reintroductions in particular have a

relatively low success rate (Dodd and Seigel 1991; but see Germano and Bishop 2009).

Genetic information is one element that may increase the success of reintroductions and

translocations, and is increasingly being used to guide such activities ( e.g. Wilson et al.

2008). The fact that A. c. blanchardi is not yet perilously close to extinction provides the

opportunity to execute a carefully planned program that can generate robust data

regarding the causes of success or failure of certain approaches; it is essentially an

experiment in conservation (Hunter and Hutchison 1994; Moritz 1999; Seddon et al.

2007). I therefore recommend that recovery efforts for this species proceed across the

northern extent of its range, using the guidelines established by this study. In addition to

preserving a potentially distinct taxonomic entity, these efforts will generate valuable

data that can be applied to other species for which little time remains.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 99 -

Chapter 5

Molecular Phylogenetics of the North American Cricket Frog Complex (Genus Acris) Across its Range

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 100 -

Abstract

The accurate delineation of species and their distributions is essential to the conservation

and management of wildlife. Currently, the North American cricket frog complex (genus

Acris) is considered to consist of two species and five subspecies. Although this group's

taxonomy has been investigated using morphology, ecology, genetic compatibility, and

molecular markers, several issues remain. To determine if genetic structure within Acris

is concordant with existing taxonomy and distributions, I sequenced a segment of the

mitochondrial control region in 843 individuals, encompassing the entire geographic

distribution and all currently recognised species and subspecies. Three major clades were

found, corresponding to A. blanchardi, a western grassland form, A. crepitans, an eastern

upland form, and A. gryllus, a southern lowland form. The distributions of these clades

were primarily defined by major rivers and the elevation change of the Fall Line. Large

sequence divergences from other A. gryllus indicate that the subspecies A. g. dorsalis is

valid, while A. c. paludicola was nested within A. blanchardi and is not. Few sympatric

sites were observed which, when combined with differences in microhabitat use and

vocalisation, suggests that hybridisation between species is likely infrequent. Moderate

structure was found within the southern range of A. blanchardi, while A. crepitans and A.

gryllus were strongly subdivided into two or more clades that likely reflect allopatric

separation in glacial refugia during the Pleistocene. Overall, Acris conforms to the

longitudinal pattern of diversity seen in many North American amphibians and reptiles.

This revised classification of Acris taxa and their distributions is timely, as declining

populations have been reported for all three species and the information can be used to

develop efficient and relevant conservation recommendations.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 101 -

Introduction

Historically, species and subspecies designations were based primarily on

morphological characteristics, with additional attributes such as vocalisation employed

when relevant. However, many species are known to be highly polymorphic for traits

such as colour or markings (Galeotti et al. 2003; Bond 2007; Forsman et al. 2008), and

some morphometric measures may be influenced by the environment (Lynch et al. 1999;

Lada et al. 2008; Teplitsky et al. 2008). Phylogenetic analyses using molecular markers

can reconstruct the genealogical relationships of groups to uncover species boundaries

without the influence of environmental plasticity or subjective interpretation of

phenotypic traits (Hebert et al. 2003; Tautz et al. 2003). Although such analyses may be

nearly concordant with traditional species classifications (Avise and Walker 1999), some

groups have been artificially subdivided based on phenotypic traits that do not reflect

genetic partitioning (Burbrink et al. 2000; Culver et al. 2000; Nittinger et al. 2007;

Cullingham et al. 2008). In contrast, morphological conservatism can result in cryptic

species being unidentified by traditional means (Bickford et al. 2007), a problem that

may be particularly pervasive in amphibians (Stuart et al. 2006; Vredenburg et al. 2007;

Angulo and Reichle 2008; Elmer and Cannatella 2008; Padial and de la Riva 2009).

The delimitation of genetic lineages within a group, their distributions, and

phylogenetic relationships is crucial to a variety of applications. It can identify the

historical events that shaped the evolution of a particular group (Austin et al. 2003b;

Noonan and Chippindale 2006; Lemmon et al. 2007a; Beamer and Lamb 2008; Koscinski

et al. 2008), or ongoing contemporary processes such as hybridisation (Good et al. 2003;

Schelly et al. 2006). Competing hypotheses regarding biogeography, taxonomic origin,

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 102 -

and patterns of diversity can be tested explicitly within a robust phylogenetic framework

(Evans et al. 2003; Pauly et al. 2004; Wiens et al. 2006; Lemmon et al. 2007a), and it

may identify the key life history traits, historical events, or ecological changes that

promoted adaptive radiations (Kozak et al. 2005; Vieites et al. 2007; Parent et al. 2008;

Givnish et al. 2009). Perhaps most importantly in this era of rapid anthropogenic

modification of the environment, phylogenetic analysis of taxa facilitates the

management and conservation of wildlife. It can be used to identify hybrid populations

that were created inadvertently and should be avoided for management purposes

(Galbreath et al. 2001; Vianna et al. 2006), or demonstrate that a taxon is a valid species

rather than a hybrid that formed as a result of anthropogenic modification of its habitat

(Wilson et al. 2000). Invasive species (Darling et al. 2008; Dubey and Shine 2008; Rius

et al. 2008) or diseases (Jancovich et al. 2005; Cullingham et al. 2008) may be more

effectively controlled by determining the origin and dispersal dynamics of animals and/or

pathogens using phylogenetics. It can also assist law enforcement through the forensic

identification of harvested animals (Wu et al. 2005; Kyle and Wilson 2007; Petersen et al.

2007). A thorough understanding of taxonomy and phylogenetic relationships is crucial

to the recovery of endangered species. It can identify cryptic species (Wang et al. 1999,

2008; Nielson et al. 2001) or genetically distinct populations (Tarr and Fleischer 1999;

Shaffer et al. 2004a) that are at greater risk of extinction due to rarity. Conversely, it may

determine that wild or captive populations are not distinct or consist of hybrids and are

therefore of little conservation worth (Avise and Nelson 1989; Ruokonen et al. 2000;

Zink et al. 2000). Numerous authors have advocated the incorporation of molecular

markers to set conservation priorities, rather than focusing solely on phenotypic variants

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 103 -

(Moritz 1999; Fraser and Bernatchez 2001). It can also be used to inform decisions

regarding source populations for use in captive breeding, reintroduction, or translocation

efforts (Wyner et al. 1999; Drew et al. 2003; Land et al. 2004; Beauclerc et al. in press).

Taxonomic overview ofAcris

The taxonomy of cricket frogs (genus A eris) has been repeatedly debated and

altered. The first specimens described, from Georgia, were classified within the true

frogs as Rana gryllus and R. nigrita (LeConte 1825). A distinct genus, Acris, was soon

recognised (Dumeril and Bibron 1841 ), although A. nigrita was later considered to

belong to the chorus frogs (Pseudacris; Fitzinger 1843). A northern form, A. crepitans,

was described shortly thereafter (Baird 1854). Although many authors acknowledged the

presence of two species (Viosca 1923, 1944; Bushnell et al. 1939), most western and

northern specimens continued to be classified as A. gryllus (Harper 1947; Burger et al.

1949) and the validity of A. crepitans was revisited several times (Dunn 1938; Mittleman

1947; Neill 1950). By the mid-1950's, however, the existence of two distinct species was

generally accepted (Blair 1958). Several subspecies were also distinguished during this

time: A. c. blanchardi, primarily west of the Mississippi River (Harper 194 7); A. c.

paludicola in coastal eastern Texas (Burger et al. 1949); and A. g. dorsalis from

peninsular Florida (Netting and Goin 1945). The currently accepted classification for this

genus includes three subspecies of northern cricket frog (northern cricket frog, A. c.

crepitans; Blanchard's cricket frog, A. c. blanchardi; coastal cricket frog, A. c.

paludicola), and two subspecies of southern cricket frog (southern cricket frog, A. g.

gryllus; Florida cricket frog, A. g. dorsalis) (Figure 5 .1; Conant and Collins 1998).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 104 -

~ A. c. crepitans • A. c. blanchardi • A. c. pa/udicola - - - A. Q. gry//us • •...... • A Q. dorsalis

Figure 5.1. Distribution of currently accepted Acris taxa. A. crepitans subspecies are indicated with shading and A. gryllus subspecies are bound by dashed lines. Modified from Conant and Collins (1998).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 105 -

As with most classifications, Acris taxa were originally defined using phenotypic

characters, including, among others: size and weight, colouration, foot length, extent of

toe webbing, size of toe discs, and the presence or absence of warts, areola, and pectoral

folds (LeConte 1825; Baird 1854; Netting and Goin 1945; Harper 1947; Burger et al.

1949). Call characteristics reportedly also distinguish species and subspecies (Nevo and

Capranica 1985), and ecological (Viosca 1944; Blair 1958; Bayless 1969) and

behavioural differences have been noted (Netting and Goin 1945; Neill 1950). A range­

wide survey of variation in several proteins was consistent with the reproductive isolation

of A. crepitans and A. gryllus, and appeared to support three distinct subspecies within A.

crepitans (Dessauer and Nevo 1969). However, Acris is the least differentiated

osteologically of North American hylids (Chantell 1968), and morphologically the two

species are highly variable and overlap at several characteristics. A number of specimens

of an intermediate nature or possessing traits of one type but located well outside their

accepted range have been reported (Netting and Goin 1945; Mittleman 1947; Neill 1950;

Mecham 1964; Mount 1975; McCallum and Trauth 2006; J. Lee, pers. comm.). Several

morphological traits, including size, appear to vary clinally across large areas rather than

clearly defining species or subspecies (Nevo 1973a; McCallum and Trauth 2006). The

colour of the vertebral stripe may reflect selection for cryptic colouration on local

substrates, gradients in aridity, or random fluctua!ions, and these factors may differ in

influence across the range (Pyburn 1961; Nevo 1973b; Gray 1983, 1984). It has also

been observed to undergo metachrosis within a single individual in response to breeding

activity, substrate, and moisture level (Pyburn 1961; Gray 1972). Furthermore, very little

ecological divergence has been observed between A. crepitans and A. gryllus in some

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 106 -

areas, and cross-fertilisation experiments found a high degree of developmental

compatibility (Mecham 1964). One study using mitochondrial DNA sequences was

unable to distinguish A. crepitans from A. gryllus and recommended that they be

synonimised (Moore 1997). In addition, there are few obvious geographic boundaries

that correspond to the current distributional limits of A. crepitans subspecies ( e.g. Rose et

al. 2006; see Figure 5.1). The reliability of many of the phenotypic traits is therefore

doubtful, and the validity of currently recognised taxa has been questioned repeatedly.

Recently, several studies have investigated phylogenetic relationships within

A eris. McCall um and Trauth (2006) conducted a detailed morphological investigation of

A. c. blanchardi and A. c. crepitans. They found little support for A. c. blanchardi as a

separate subspecies, and recommended that it be subsumed within A. crepitans. Rose et

al. (2006) attempted to determine the phylogenetic affinity of A. c. paludicola. Although

several morphological and behavioural traits were shared with A. gryllus, cytochrome b

sequences placed it with A. c. blanchardi, but without strong support for subspecific

status. Finally, a molecular analysis of all Acris species and subspecies indicated that the

phylogenetic relationships of these taxa and their distributions require substantial revision

(Gamble et al. 2008). Although this study provided strong evidence, it was based on a

small number of samples (64) and did not cover the full geographic spread of all groups.

My objective in this chapter is therefore to use mitochondrial DNA (mtDNA) sequences

from all species and subspecies of Acris to determine their phylogenetic relationships,

and to use this information to assess the validity of the currently recognised groups. I

will use samples from across the entire geographic range of this complex and numerous

samples per site. This detailed sampling will increase the likelihood of detecting

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 107 -

geographically restricted lineages, provide clearer delineation of geographic boundaries

for each lineage, and identify the biogeographic barriers that were important in shaping

the genetic structure of this group. It will also provide information on the extent of

sympatry between related forms, and thus the potential for hybridisation. In light of

recent conservation concerns for A. c. blanchardi (Gray and Brown 2005; also see

Chapter 4), A. c. crepitans (Gray and Brown 2005; NYSDEC 2007), and A. g. gryllus

(Jensen 2005; Micancin 2008), it is imperative that the species and subspecies status of

Acris taxa be clarified and their distributions described explicitly.

The most appropriate philosophical framework through which species should be

designated, known as a "species concept", continues to be debated among biologists.

Two of the most common are the biological species concept (BSC), which states that a

species is a group of naturally interbreeding populations that are reproductively isolated

from other such groups (Mayr 1963), and the phylogenetic species concept (PSC), in

which a species is defined as the smallest group of populations that is diagnosable by

unique characters and shares a parental pattern of common ancestry and descent (Cracraft

1983). Numerous criticisms have been made of both these concepts (see Avise 2000 for

a review). To incorporate the strengths of both the BSC and the PSC, Avise and Ball

(1990) proposed the "principles of genealogical concordance" as a means of recognising

species. In this concept, common patterns in the variation at two independent characters

(such as multiple genetic loci, morphology, or biogeography) indicate shared

evolutionary history and thus reflect true phylogenetic relationships. These principles

can be used to identify both species ( a concordant group that can interbreed and is

reproductively isolated from other such groups) and subspecies (phylogenetically distinct

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 108 -

groups that are reproductively compatible but geographically isolated from other such

groups) (Avise 1994). The inherent appeal of this concept is its incorporation of multiple

lines of evidence that increase the likelihood that a given genealogical tree is

representative of the true species phylogeny. This concept will be employed to the extent

possible, given the available data, when reviewing the taxonomic status of Acris.

Materials and Methods

Sample collection and molecular methods

Tissue samples of all subspecies of A. crepitans and A. gryllus were collected

from several locations across their range (Figure 5 .2). Counties were considered to be

sampling sites, and the majority (75%) consisted of a single pond. Geographical

locations were taken as the latitude and longitude for the county in which a pond or

stream was located, and were obtained from the U.S. Census Bureau Census 2000

Gazeteer (http:/ /www.census.gov/; Appendix 1). The number of samples per site ranged

from one to 30, with an average of 4.6. In total, 843 samples were genetically profiled

from 185 sites in 22 states (Table 5.1, Appendix 1). Tissue samples were collected from

adults between 2001 and 2007 except the Oklahoma sites of Roger Mills, Cherokee,

Wagoner, Custer, Logan, and Canadian, which were collected in 1998 as part of a

previous study. All tissue samples consisted of either toe clips collected in the field or

muscle from vouchered specimens, and were preserved in 95% ethanol or 1 X lysis

buffer (2 M urea, 0.1 M NaCl, 0.25% n-lauroyl sarcosine, 5 mM CDTA, 0.05 M Tris-HCl

pH 8). Samples were collected and identified to species by biologists located across the

distribution of Acris. The highly variable morphology of these anurans means that the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

......

®

e

®

@

--""_

0

0

®

@

...

<0

,

.....

-·'

~/

li':l1ll

.'@ Ji @

Q

(rf:•0

0 /

I

/

\

0

0

l

,'

0

·'

,--~--

,G

\~

0

0

0

0

0

@

J

0

......

"'

L

~0

O! ol

0

@\

_

1

----

200

0

0

0

.....

A.

G 02

0 195

00

,

0

0

-

.....

0

,.,'

4

3 1

NoClade

0

blanchardi

'

,---'

0

0

400

,,

\

,'

!

)/

:

0

.,

,

600

A.

1111

Western

Eastern Oro

crepitans

800

~;c

b,. A.

.A A

A

A

1000

Western

Florida

NC-02 Noclade

NC-01

gryl/us

Kilometres

A

N

"l,

-

109

-

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Figure gryllus.

phylogenetic

rivers as Charles

follows:

that

5.2.

City

Different

are

A.

Distribution

VA

reconstruction

discussed

blanchardi/A.

(4);

colours correspond

A.

crepitans/A.

in

of

the

Acris

of

gryllus

text.

control

sample collection

The

in

gryllus

to

region

St.

subclades

dashed

Tammany

in

haplotypes:

Gates

line

identified

sites.

represents

NC LA

(1);

(5),

Symbols

green

within

A.

Wayne

the

blanchardi/A.

circles,

Fall

correspond

each

NC

Line

A.

major

(6),

blanchardi;

(modified

Durham

crepitans

to

clade.

the

three

NC

Sites

from

orange

in

(7),

major

Henderson

containing

Fenneman

and

squares,

clades

Gadsden

KY

A.

multiple

and

identified

crepitans;

(2),

Johnson

FL

Adams

(8).

species

during

Dark

blue

1946).

OH

are

triangles,

lines

(3),

numbered

and

are

-

110

A.

-

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 111 -

Table 5.1. Summary of samples collected for five A eris subspecies used in this study.

Taxon State Site # sites N Benton, Franklin, Fulton, Garland, AR 6 18 Sebastian, Stone IL All 5 65 IN All 8 36 IA All 7 31 KS All 23 47 KY All 3 6 A. c. blanchardi MI All 2 27 MO All 11 41 NE All 6 6 OH All 14 92 OK All 12 42 SD All 2 20 TN Montgomery 1 2 TX Hood, Kendall, McMullen, Menard, Travis, Wise 6 45 AR Columbia, Lafayette 2 4 DE All 2 6 GA Greene, Monroe 2 6 all except Calcasieu, Cameron, East Feliciana, LA Livingston, St. Helena, St. Martin, St. Tammany, 20 66 Tangipahoa, Vermillion, Washington MD All 3 6 A. c. crepitans MS Carroll, Grenada, Tate 3 11 NC Gaston, Mecklenburg, Durham 3 10 NJ All 1 2 SC Pickens, Abbeville, Darlington 3 15 TN Cumberland, McNairy 2 6 VA All 1 8 TX Anderson, Bowie 2 8 A. c. paludicola LA Vermillion, Cameron, Calcasieu, St. Martin 4 16 AL Covington, Tuscaloosa 2 19 GA Taylor 1 4 Ascension, East Feliciana, Livingston, Pointe LA Coupee, St. Helena, St. Tammany, Tangipahoa, 8 39 A. g. gryllus Washington Forrest, Greene, Hancock, Harrison, Perry, MS 7 30 Tishomingo, Wilkinson NC Moore, New Hanover, Washington, Wayne 4 19 SC Barnwell, Berkeley, Darlington, Marion 4 23 FL All 5 39 A. g. dorsalis GA Liberty 1 10 Unknown AL Tallapoosa 1 3 crel!._itans/ gryllus NC Gates, Wayne 2 15

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 112 -

characteristics used to diagnose subspecies are at best ambiguous, making further

identification of specimens difficult and potentially unreliable (Mittleman 1947; Neill

1950; Chantell 1968; Nevo 1973a; McCallum and Trauth 2006; Rose et al. 2006;

Micancin 2008). Subspecies designations were therefore determined based on the

location of the specimen within the range distributions of Conant and Collins ( 1998).

While this does not allow an evaluation of the suitability of the morphological traits, it

does enable an approximate assessment of these currently accepted distributions relative

to any genetic lineages that are found. In total, 478 samples were classified as A. c.

blanchardi, 148 as A. c. crepitans, 16 as A. c. paludicola, 134 as A. g. gryllus, and 49 as

A. g. dorsalis. Eighteen samples from an area where both A. crepitans and A. gryllus are

believed to exist in sympatry were not identified to species (Table 5.1 ). The majority of

A. c. blanchardi samples were previously described and profiled in Chapter 4.

Approximately 10 mg of tissue was dissolved in 500 µL of IX lysis buffer and

digested with 50 µL of proteinase K. Total genomic DNA was extracted with the

DNeasy Tissue Kit (Qiagen Inc.) or an automated magnetic bead procedure, and

quantified with the PicoGreen Quantitation Kit (Molecular Probes) on a FLUOStar

Galaxy (BMG Labtech). A fragment of the control region was amplified using the

primers ControlWrev-L/ControlP-H (Goebel et al. 1999) in a 25 µL reaction containing

1-5 ng DNA, 1 X PCR buffer, 2 mM MgCb, 0.2 mM dNTPs, 0.1 mg/mL BSA, 0.2 µM

each primer, and 1.25 U Taq polymerase (Invitrogen). DNA amplification was

performed in an MJ Research PTC-225 thermal cycler with the following conditions:

initial denaturation at 94°C for 5min, 35 cycles of denaturation at 94°C for 30 sec,

annealing at 40°C for 30 sec, and extension at 72°C for 45 sec, with a final extension of

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 113 -

72°C for 10 min. PCR product was visualised on 1% agarose gels stained with ethidium

bromide, and 5 µL of product was purified with ExoSAP-IT (USB). Cycle sequencing

was performed with the primer ControlP-H using the DYEnamic ET-Terminator kit on a

MegaBACE 1000 DNA Analysis System (GE Healthcare).

Phylogenetic analyses

Sequences were aligned with CLUSTALX v.1.81 (Thompson et al. 1997) and edited

in BIOEDIT v.7.0.5.3 (Hall 1999). ARLEQUIN v. 3.11 (Excoffier et al. 2005) was used to

determine the number of variable sites and haplotype frequencies. A comparison of log

likelihood scores for 56 models using the Akaike Information Criterion 2 in

MODELGENERATOR V.0.85 (Keane et al. 2007) found the HKY model with a proportion of

invariable sites and gamma distribution to be the most likely model of sequence evolution

for the data (AIC = 13010.44, lnL = -4812.72). The parameter estimates from this model

were: base frequencies: A= 0.33, C = 0.11, G = 0.21, T = 0.6; ratio of transitions to

transversions = 5.23; proportion of invariant sites (I)= 0.37; and gamma distribution

shape parameter (a)= 0.55. This model was used in ARLEQUIN to calculate pairwise

sequence divergences, nucleotide diversity (7rn; average number of pairwise differences),

and haplotype diversity (h; probability that two randomly chosen haplotypes are

different) for the entire dataset and all major groups identified during phylogenetic

analysis. However, ARELQUIN does not provide the option to use the HKY model, so the

Tamura-Nei model was used instead, as it encompasses the same parameters (Tamura

and Nei 1993). Nucleotide diversity was also calculated for each site with two or more

samples without assuming a model of nucleotide substitution or gamma distribution.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 114 -

Phylogenetic inference of all haplotypes employed two approaches: distance and

Bayesian. An unrooted neighbour-joining tree was constructed in MEGA v.4.0 (Kumar et

al. 2004). Since MEGA also does not support the HKY model, the Tamura-Nei model was

used with a= 0.55. One thousand bootstrap replications were used to determine support

for each node. Bayesian analysis was conducted in MRBAYES v.3.1 (Ronquist and

Huelsenbeck 2003). MRBA YES uses a Markov chain-Monte Carlo algorithm to sample the

posterior distribution of the parameters, and Metropolis coupling to accelerate the

convergence of the Markov chain. Eight Markov chains were run simultaneously for two

million generations, with two random chain swaps each generation. Sampling occurred

every 100th generation with a bum-in of 5000. The heating parameter was reduced to 0.1

to enable the cold chain to swap states with heated chains more frequently, which

increases the probability of finding isolated peaks in the data. Two independent runs

were conducted simultaneously and used to generate a consensus tree. Parameters for the

HKY+l+f model were estimated from the dataset.

Evaluation ofcurrently recognised species and subspecies

Analysis of molecular variance (AMOVA) examines the genetic variation at each of

three levels of a hierarchical grouping: within individual sample sites, among sites within

groups, and among groups (Excoffier et al. 1992). If grouping specimens into species

and subspecies using currently accepted morphological criteria and distributions (Table

5.1) is concordant with the clades obtained during phylogenetic analysis, AMOVA should

recover a similar amount of genetic variation within and between groups for both

methods. Samples were grouped into all five recognised subspecies, as well as only three

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 115 -

groups to enable direct comparison to the three main clades identified during

phylogenetic analysis (see Results): A. c. blanchardi, A. c. crepitans/A. c. paludicola,

and A. g. gryllus!A. g. dorsalis. The 18 unidentified samples were excluded from this

analysis. AMOVA was carried out in ARLEQUIN using the Tamura-Nei model of sequence

evolution with a= 0.55. Pairwise sT values were also calculated in Arlequin for the

three-group system and the phylogenetic clades to determine which method accounted for

a greater degree of structure. This analogue of FsT incorporates nucleotide variation as

well as haplotype frequencies (Excoffier et al. 1992). Significance at a = 0.05 was

determined using 1000 permutations of the data.

Genetic structure within clades

The presence of genetic structure among sample sites within each of the major

clades identified during phylogenetic reconstruction was assessed using AMOV A. Due to

the large number of haplotypes, individual phylogenetic trees were generated for each of

the major clades. Both neighbour-joining and Bayesian methods were used, and settings

were identical to those for the entire dataset, with the exception that Bayesian analysis

used six chains instead of eight. To obtain outgroups for each tree, the most common

haplotype of one or both other clades was used: AcrCR-1 (A. c. blanchardi) and AcrCR-

172 (A. c. crepitans) were used for A. gryllus, and AcrCR-172 and AcrCR-202 (A. g.

gryllus) were used for A. c. blanchardi. Only AcrCR-1 was used as an outgroup for A. c.

crepitans, as the very large divergences between A. c. blanchardi and A. gryllus

sequences were problematic (data not shown).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 116 -

Barriers to gene flow

BARRIER v.2.2 (Manni et al. 2004) was employed to determine the existence of

historical barriers to gene flow by defining zones of maximum genetic change among the

185 sample sites. BARRIER uses Delauney triangulation to connect a set of points (sample

sites), and the genetic distance between the two points is associated with the edges of the

triangles. A barrier is then constructed using Monmonier' s maximum difference

algorithm, which identifies the edge with the largest genetic distance. A line is traced

perpendicular to this edge, and is extended along adjacent edges of decreasing genetic

distance until the boundary reaches a previously computed boundary, the limits of the

map, or closes on itself. Multiple boundaries can be constructed by progressively

locating the next largest genetic distance. Geographic coordinates of each sample site

were used (Appendix 1), and genetic distances were computed in ARLEQUIN as the

corrected average pairwise sequence difference between populations. All negative values

were assumed to be 0. Ten barriers were constructed; however, the algorithm will

generate as many barriers as requested regardless of their biological significance. Thus,

all barriers that were initiated with a genetic distance of< 0.1, the maximum sequence

divergence found within a single major clade (see Results), were disregarded.

Results

Phylogenetic reconstruction

Analysis of 485 bp of sequence identified 337 haplotypes (Appendix 2), with 215

variable sites consisting of 182 transitions, 93 transversions, and 10 one base pair indels.

Overall nucleotide diversity was 0.13 and haplotype diversity was 0.93. Maximum

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 117 -

sequence divergence was 0.37 between two groups of haplotypes: AcrCR-24 (Kiowa/

Pawnee KS) vs. AgrCR-313 (Wayne NC); and AgrCR-277 (Livingston LA) vs. AcrCR-

173 (New Castle DE), AcrCR-184 (Gaston/Mecklenburg NC), and AcrCR-186 (Cape

May NJ). For sample sites with two or more samples, nucleotide diversity ranged from 0

to 0.099 (Durham NC), and the number ofhaplotypes ranged from one to 17 (Gadsden

FL). As with the detailed analysis of A. c. blanchardi (Chapter 4), a large number of

low-frequency haplotypes were present: 86% occurred only once or twice in the sample

set. The haplotypes that were most frequent also did not change when incorporating

additional species: AcrCR-1 was most common (0.27), followed by AcrCR-19 (0.02)

and AcrCR-22 (0.02). Of the new haplotypes described, two were relatively common:

AcrCR-172 and AgrCR-202, both of which occurred at a frequency of 0.02 (Appendix 2).

The unrooted phylogenetic trees generated using neighbour-joining and Bayesian

methods produced concordant topologies: three well-supported clades that

approximately correspond to the current distributions of A. c. blanchardi, A. c. crepitans,

and A. gryllus (Figure 5.3). The majority of samples and haplotypes belonged to A. c.

blanchardi (Table 5.2). However, nucleotide and haplotype diversity and maximum

pairwise sequence divergence were lowest for this lineage, and greatest in A. gryllus

(Tables 5.2 and 5.3). The A. c. blanchardi and A. c. crepitans clades clustered together at

a large distance from the A. gryllus clade. Analysis of molecular variance showed that>

90% of the genetic variation was attributed to differences among the three clades, and

very little was due to differences among sample sites within clades or within sites (Table

5.4). Pairwise sr values between clades were overall very large and highly significant,

with the largest value between A. c. blanchardi and A. gryllus, while A. c. crepitans was

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 118 -

A. gryllus

A. c. crepitans

Eastern

Western A. c. blanchardi

0.9

Figure 5.3. Unrooted Bayesian tree for 337 control region haplotypes from the Acris species complex. Major lineages and probability values (x 100) are indicated. Haplotype labels are omitted for clarity.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 119 -

Table 5.2. Summary statistics of three clades identified during phylogenetic analysis of Acris haplotypes.

Statistic A. c. blanchardi A. c. crepitans A. gryllus No. individuals 573 72 198 No. sample sites 135 25 34 No. haplotypes 171 30 136 Most frequent haplotype 1 (0.269) 172 (0.017) 202 (0.017) (frequency)

,r0 (average no. differences) 0.0109 (5.3) 0.0277 (13.3) 0.0414 (19.9) H 0.84 0.95 0.99

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 120 -

Table 5.3. Average, maximum and minimum sequence divergences between several clades of A eris identified by phylogenetic analysis. The putative subspecies A. c. paludicola is also included, and was classified based on location of the samples according to Conant and Collins (1998). The putative subspecies A. g. dorsalis consists of the three divergent haplotypes identified as the Florida clade during phylogenetic analysis.

Clade Average Min Max Max between A. crepitans 0.043 0 0.141 120 and 1851190 A. c. blanchardi 0.017 0 0.045 160 and 95 A. c. paludicola 0.018 0.002 0.035 103 and 149 blanchardi - paludicola 0.018 0 0.042 95 and 1451149 A. c. crepitans 0.036 0 0.091 183 and 1781180 1941195 and eastern 0.010 0.004 0.015 1781181 western 0.008 0 0.018 190 and 183 eastern - western 0.082 0.070 0.091 183 and 1781180 A. gryllus 0.042 0.002 0.102 276 and 245 A. g. gryllus 0.041 0.002 0.093 297 and 302 eastern 0.012 0.002 0.032 321 and 3351337 western 0.015 0.002 0.038 256 and 289 eastern - western 0.060 0.042 0.080 321 and 2151266 NC 01 0.007 0.002 0.013 313 and 3101315 NC02 0.018 NIA NIA NIA NC 01-NC 02 0.059 0.052 0.069 302 and 313 eastern - all NC 0.045 0.027 0.074 308 and 337 western - all NC 0.071 0.053 0.093 297 and 302 A. g. dorsalis 0.021 0.013 0.025 245 and 2241246 dorsalis - gryllus 0.075 0.050 0.102 276 and 245 dorsalis clade - traditional a 0.070 0.053 0.082 245 and 2051226 blanchardib - crepitans 0.118 0.097 0.141 120 and 1851190 blanchardl - gryllusc 0.294 0.24 0.370 313 and 24 277 and - gryllusc 0.322 0.278 0.370 crepitans 17311841186 a includes all A. g. dorsalis individuals classified according to Conant and Collins (1998) b includes putative A. c. paludicola individuals c includes both A. g. gryllus and A. g. dorsalis

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 121 -

equally distant from A. c. blanchardi and A. gryllus (Table 5 .5). A comparison of

pairwise sequence divergences between haplotypes among these clades suggested a

slightly different pattern (Table 5.3): values were smallest between A. c. blanchardi and

A. c. crepitans, while values were much larger and similar between A. gryllus and both A.

c. blanchardi and A. c. crepitans.

The distribution of the A. c. blanchardi clade was similar to that currently

recognised for this subspecies (Conant and Collins 1998), consisting primarily of the

Great Plains region. However, it was also found across Arkansas and Texas in their

entirety, Louisiana west of the Mississippi River, and seven sites east of the Mississippi

River in Louisiana and Mississippi, while no haplotypes were found in its previously

recognised range of northern and western Tennessee (Figure 5.2). Two sites along the

Ohio River in northern Kentucky harboured A. c. blanchardi haplotypes, although limited

sampling prevents drawing conclusions about the remainder of the state. Several samples

also occurred substantially outside this distribution in St. Mary's MD and Charles City

VA. The A. c. crepitans clade occurred exclusively east of the Alabama and Coosa

Rivers, with no haplotypes found in its previously recognised range in Texas, Louisiana,

Mississippi, and western Alabama (Figure 5.2). Its northwestern distribution was bound

by the Ohio River in Kentucky, although six samples occurred north of this in

Washington IN and Adams OH. It also occurred throughout the southern Appalachian

Highlands and into the Atlantic Coastal Plain from northern North Carolina to New

Jersey. A single sample was also found in Gadsden FL. The distribution of the A. gryllus

clade corresponded well with its previously recognised range: it occurred primarily

throughout the Atlantic Coastal Plain, exclusively east of the Mississippi River and south

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Talble

5.4.

a

call b d

Analysis

classified

A.

original

A.

g.

A.

A.

A.

A.

A. A.

A. c.

5

dorsalis

g.

A.

c.

g.

g.

A. subspeciesb

paludicola

3

g.

g.

3

blanchardi

of

dorsalis

5

Group

crepitans

blanchardi

groupsa

gryllus

gryllus

clades

according to

gryllus

dorsalii

dorsalisc

subspecies

molecular

individuals

vs.

vs.

individuals

vs.

variance

as

Conant

classified

identified

classified

for

and

650

644

438 644

183 190

438 181

133

164

163 134

164

df 48 32

33 23

33

2 4 2

1

1 1

several

according

Collins

as

a

according

distinct

groupings

(1998)

Among

Among

to

Among Among

Among Among

Variance

Conant

clade

to

but

Among

Among

Among

Among

Among

Among

Conant

Among

Among

Among

Within

Within

Within Within

Within

Within

Within

Within

Within

of

sites

sites

sites sites

sites

sites

A.

during

Acris

and

components

c.

within

within

within within

within within

species

species

subsp. clades

subsp.

subsp.

blanchardi,

Collins

sites

and sites

sites sites

sites sites

sites

sites

sites

sites

sites

sites

haplotypes.

phylogenetic

Collins

species

species

clades

subsp.

subsp.

subsp.

(1998)

A.

(1998)

c.

analysis

crepitans,

%

of

91.93 79.92

47.42 79.62

44.00 46.96 40.70 60.55

48.39 20.85 14.73 24.65

75.35 14.26

51.61 24.17 18.60

75.83

12.34

2.84

5.23

5.65

5.82

8.58

variation

and

A.

gryllus

< < < < < < < <

< < < < < < < < < <

<

P-value

0.1154

0.0606

0.0001 0.0001 0.0001 0.0001 0.0001 0.0001 0.0001 0.0001

0.0001

0.0001 0.0001 0.0001 0.0001 0.0001 0.0001 0.0001 0.0001

0.0001

0.0001

as

separate

groups

-

122

-

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 123 -

Table 5.5. Pairwise sr values obtained when Acris samples are classified according to currently accepted species designations (above diagonal) and the phylogenetic clades observed here (below diagonal). All values were significant at P = 0.05 based on 1000 permutations.

A. c. blanchardi A. c. crepitans A. gryllus A. c. blanchardi 0.50 0.83 A. c. crepitans 0.88 0.81 A. gryllus 0.93 0.88

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 124 -

of the Tennessee River, throughout Florida and along the Atlantic coast to northeastern

North Carolina (Figure 5.2). Of the 185 sites sampled, only eight harboured haplotypes

belonging to two different clades: Adams OH, Henderson KY, and Charles City VA

contained A. c. blanchardi and A. c. crepitans haplotypes; Gadsden FL, and Gates,

Wayne, and Durham NC contained A. c. crepitans and A. gryllus haplotypes; and both A.

c. blanchardi and A. gryllus haplotypes were found in St. Tammany LA (Figure 5.2).

Testing existing species classification

When samples were classified according to the currently accepted species and

subspecies (as listed in Table 5.1), pairwise

(Table 5.5). However, these values were substantially smaller than those obtained for

groups based on the mitochondrial clades, particularly for the A. c. blanchardi-A. c.

crepitans comparison. AMOV A also found that less variation was due to differences

among these currently recognised forms, whether grouped into three or five taxa, than the

clades identified here (Table 5.4).

The subspecies A. c. paludicola and A. g. dorsalis, as defined by Conant and

Collins ( 1998), did not appear as distinct clades within the phylogenetic trees and were

therefore investigated further. Although both groups possessed some unique haplotypes,

most were either unresolved or did not cluster together during phylogenetic analysis

(Figures 5.4 and 5.5). There are a few exceptions: haplotypes AcrCR-141, 145, 146, 148

and 149 in putative A. c. paludicola from St. Martin and Cameron LA clustered together,

but this group also included A. c. blanchardi haplotypes from throughout Louisiana as

well as Mississippi, Virginia, and Maryland; and AgrCR-224, 225 and 226 in A. g.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 125 -

dorsalis from Putnam and Collier, FL formed an exclusive, well-supported clade with

long branch length. Additional individuals that would be classified as A. g. dorsalis

based on Conant and Collins (1998) were nearly as divergent from this clade as were all

other A. g. gryllus individuals (Table 5.3). AMOVA also found that very little variation

could be accounted for by the separation of A. c. paludicola from A. c. blanchardi, and

traditional A. g. dorsalis from A. g. gryllus (Table 5.4). However, the variation among

these latter subspecies was much larger and significant when A. g. dorsalis was

considered to be the clade of three haplotypes identified during phylogenetic analysis.

Analysis ofclades

Overall, there appears to be limited genetic structure within the A. c. blanchardi

clade. Although significant, AMOV A showed that slightly more variation was found

within sample sites than among them (Table 5.4). Accordingly, neighbour-joining did

not identify any clades of more than two haplotypes with bootstrap support > 80 ( data not

shown). Bayesian analysis, however, revealed five well-supported clades that were

composed of at least four haplotypes (Figure 5.4). Nearly all individuals belonging to

these clades were found south of the Missouri River (Figure 5.2). Clades 1 and 2 were

found nearly exclusively south of the Arkansas River in Oklahoma, Arkansas, Texas and

Louisiana. Clade 3 occurred primarily in Louisiana, south of the Red River, while Clade

4 was between the Missouri and Red Rivers. Clade 5 was the most widespread, ranging

from Missouri to southern Texas and Louisiana. With the exception of eight small

clusters containing two or three haplotypes, the remaining haplotypes were not resolved,

and these were found across the entire distribution of the A. c. blanchardi clade.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 126 -

~------1 93 ------2 ------3 -----=- __J ------4

1----5

0.6

Figure 5.4. Bayesian phylogenetic tree of 171 A. c. blanchardi control region haplotypes. Lineages with probability values > 0.8 (x 100) and with more than three haplotypes are indicated at the base of the corresponding branch. Haplotype labels have been omitted for clarity. Asterisks indicate haplotypes belonging to individuals classified as A. c. paludicola based on geographical location, and arrows indicate haplotypes belonging to individuals found in St. Mary's MD and Charles City VA. Numbers of major clades correspond to Figure 5.2.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 127 -

AMOVA showed that approximately 75% of the variation within A. c. crepitans

was due to differences among sample sites, suggesting the presence of genetic structure

within this group (Table 5.4). All A. c. crepitans haplotypes clustered into two well­

supported clades, which were apparent in the phylogenetic analysis of all Acris

haplotypes (Figure 5.3). Individual neighbour-joining and Bayesian trees for A. c.

crepitans are therefore not displayed, as they did not identify any additional structure. A

western clade was found exclusively in Tennessee and Kentucky, between the Tennessee

and Ohio Rivers, with several samples north of the Ohio River in Washington IN (Figure

5.2). An eastern clade occurred primarily throughout the southern Appalachian

Highlands and the northern Atlantic Plain, but also included single individuals in Adams

OH, Cumberland TN, and Gadsden FL. Sequence divergence between these two clades

was large, while that within each clade was substantially smaller and of a similar

magnitude to one another (Table 5.3). A similar amount of divergence exists between

these two clades as between the subspecies A. g. gryllus and A. g. dorsalis, and it was

only slightly less than the smallest divergence found between A. c. blanchardi and A. c.

crepitans (Table 5.3). When sample sites were further subdivided into eastern and

western clades, AMOV A demonstrated that > 90% of genetic variation is accounted for by

these two clades (data not shown).

Significant genetic structure was also present within A. gryllus, as AMOVA

revealed that approximately 75% of the variation within this clade was due to differences

among sample sites (Table 5.4). Neighbour-joining and Bayesian analysis of A. gryllus

found several well-supported lineages that were consistent between analyses, although

their exact placement relative to other poorly-supported branches varied among the two

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 128 -

methods. Thus, only the Bayesian results are presented (Figure 5.5). A well-supported

clade contained all haplotypes located west of the Mobile, Alabama and Coosa Rivers

(Figure 5.2). Structure among the haplotypes located east of the Mobile/Alabama/Coosa

system was more complex and overall poorly resolved. A well-supported Florida clade

contained three haplotypes from sites in peninsular Florida (Collier and Putnam). These

were grouped with the western haplotypes with high probability by the Bayesian analysis

(1.00), but with weak bootstrap support in the neighbour-joining analysis ( 47). The

majority ofhaplotypes in North Carolina formed two well-supported clades, although

they did not cluster together with strong support. The remaining eastern haplotypes

represented a large, polytomous group with very short branch lengths and little support

for most clades containing three or more haplotypes. Sequence divergence within each of

these five clades was generally low, while divergences between them were substantially

greater (Table 5.3). Separation of sample sites into these five clades demonstrated that

more than 80% of the variation was due to this arrangement (data not shown).

Barriers to gene flow

Monmonier' s maximum difference algorithm identified four genetic boundaries

that were constructed with an initial genetic distance of> 0.1 (Figure 5.6). Additional

barriers were mainly small, separating only two to four sample sites, suggesting that the

first four are likely the only biologically relevant barriers at this scale. The first barrier

constructed, representing the largest genetic distances, began in a north-south direction

along the Mississippi River. This separated sites containing A. c. blanchardi haplotypes

in Arkansas, Missouri, and western Louisiana from sites harbouring A. gryllus haplotypes

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 129 -

a2~-//------i= E==--

100 100 ] North Carolina 86 100 =:J---Florida

100

92

Western

100 ~--//------c======.!:.A~cr.!..._1 -Acr172 0.6

Figure 5.5. Bayesian analysis of 136 A. gryllus control region haplotypes. Lineages with probability values > 0.8 (x 100) and with more than three haplotypes are indicated at the base of the corresponding branch. Approximate geographic location of significant clades is indicated. Haplotype labels have been omitted for clarity. Asterisks indicate haplotypes belonging to samples classified as A. g. dorsalis based on geographic location, with the exception of three strongly supported haplotypes that are designated as the Florida clade.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 130 -

u . ..

Figure 5.6. Genetic boundaries (heavy lines) identified for Acris control region haplotypes with Monmonier's maximum difference algorithm in the program BARRIER. Boundaries are numbered in order of decreasing initial genetic distance. Points are sample sites.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 131 -

in Mississippi and eastern Louisiana. One exception to this was Carroll MS, which was

found on the western side of the barrier, and contains exclusively A. c. blanchardi

haplotypes. The barrier continued east along the bottom of the Ohio River and then south

along the Tennessee River, separating A. c. crepitans and A. gryllus haplotypes in

Tennessee. As it continued east, it approximately followed the Fall Line, separating A. c.

crepitans haplotypes in northern Alabama and Georgia and western South Carolina and

North Carolina from A. gryllus haplotypes in the south and east of these states. The

remaining barriers were much shorter. The second barrier separated Durham NC, which

contained primarily A. c. crepitans individuals, from five sites in eastern North Carolina

that were predominantly A. gryllus. The third barrier separated St. Mary's MD from

adjacent sites in Virginia, Maryland and Delaware. The fourth barrier corresponded

approximately with the Ohio River, separating A. c. blanchardi haplotypes to the north

and A. c. crepitans haplotypes to the south.

Discussion

Validity ofcurrently recognized species and subspecies

The frequent adjustment of Acris taxonomy and repeated reports of individuals

with intermediate or ambiguous characters indicate that a comprehensive evaluation of

phylogenetic relationships within this complex is long overdue. As stated by Rose et al.

(2006, p. 433), "the lack of interest in elucidating clearly the interspecific and

intraspecific relationships of these two frogs is perplexing because the two species

occupy a major geographical segment...of the continental United States". Several studies

have noted the poor capacity of most morphological traits to clearly differentiate these

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 132 -

groups (Mittleman 1947; Neill 1950; Mccallum and Trauth 2006; Rose et al. 2006;

Micancin 2008). For taxa with extensive phenotypic variation, such as Acris, molecular

genetic markers are much more suitable for resolving phylogenetic relationships as they

are less subjective and not affected by the environment.

The results presented here demonstrate that the distribution of mitochondrial

variation in cricket frogs does not correspond exactly to the distribution of species or

subspecies as defined by morphology or in Conant and Collins (1998) (Figure 5 .1 ).

Rather than two widely-ranging species, the control region haplotypes formed three well­

supported clades that correspond to the current forms of A. c. blanchardi, A. c. crepitans,

and A. gryllus. The relationship among these three groups varied slightly when using

different indices: phylogenetic analyses and sequence divergences indicated that A. c.

blanchardi and A. c. crepitans are most closely related, and both are equally distant from

A. gryllus. However, a measure of genetic distance (

and A. gryllus are the most distant taxa, with A. c. crepitans equidistant from both.

incorporates both nucleotide sequence and haplotype frequency, so the very different

sample sizes used for each taxon may have affected estimates of genetic distance. The

independence of sequence divergence from sample size suggests that this is the more

reliable measurement for this study.

Although A. c. blanchardi and A. c. crepitans are closely related, they represent

independent, monophyletic groups (the few disjunct samples will be addressed later).

Few amphibian studies have employed the mitochondrial control region for phylogenetic

analysis, but those that have typically observed < 6% sequence divergence within species

(Shaffer and McKnight 1996; Church et al. 2003; Shaffer et al. 2004b; Smith and Green

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 133 -

2004; Rowe et al. 2006; Pauly et al. 2007). The range of divergences between A. c.

blanchardi and A. c. crepitans substantially exceeds this, suggesting that these two taxa

do not represent a single species. This supports the recent recommendation by Gamble et

al. (2008) that the western form be considered a distinct species, A. blanchardi, while the

eastern form remains A. crepitans.

Two additional subspecies are currently recognised, A. c. paludicola and A. g.

dorsalis. Although Rose et al. (2006) found that there was no molecular support for

recognising A. c. paludicola, they recommended that it be maintained as a distinct

subspecies based on a series of morphological and behavioural characteristics. Gamble et

al. (2008) also found that A. c. paludicola was nested within A. blanchardi, and suggested

that it does not warrant subspecific status. Results presented here support this view.

Haplotypes belonging to this subspecies were located within the A. blanchardi clade,

with little resolution or clustering for most, and the variation accounted for by separation

into these groups was not significant. This subspecies was- diagnosed from specimens

located at a single site, and based principally on its "decidedly pink" colour, particularly

of the vocal pouches in calling males (Burger et al. 1949). However, the colour of the

vocal pouch appears to be due to blood flow (Rose et al. 2006) and has been observed in

frogs as far away as Michigan (pers. obs.). Moreover, red vertebral stripes are known to

occur in all Acris taxa and across the entire distribution (Nevo 1973b ). It may increase in

frequency on red substrates or in more humid conditions, such as that found in coastal

Louisiana and Texas where this subspecies occurs, as abundant vegetation can enable red

individuals to remain hidden from predators (Pyburn 1961; Nevo 1973b). The very short

adult lifespan and rapid turnover of populations in A eris (Burkett 1984) means that

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 134 -

locally optimal phenotypes could rapidly attain high frequencies, rather than reflecting

long-term isolation. Thus, the primary diagnostic characteristic of this subspecies

(colour) is ambiguous and not distinctive, and, when combined with the genetic data,

provides little support for the recognition of A. c. paludicola.

The validity of A. g. dorsalis has not been previously evaluated. When all

individuals that could be classified as A. g. dorsalis according to Conant and Collins

(1998) were considered, results were similar to A. c. paludicola: haplotypes were

distributed throughout the A. gryllus tree, with little resolution or pattern, and no

significant variation was attributed to this arrangement of groups. However, three

haplotypes (AgrCR-224, 245,246) from peninsular Florida formed a strongly supported

clade separate from non-peninsular A. g. dorsalis haplotypes, and were nearly as different

from these (0.05-0.08 divergence) as they were from all other A. g. gryllus haplotypes

(0.05-0.10 divergence). Thus, the peninsular Florida haplotypes appear to represent a

distinct lineage, but remaining non-peninsular A. g. dorsalis individuals are not

distinguishable from A. g. gryllus based on mtDNA. The presence of a divergent Florida

clade is not surprising, as the Floridian peninsula has a distinct climate, physiography,

and ecology, particularly at its southern extreme (Avise 2000). A disproportionately

large number of endemic species are found in this region (Remington 1968; Delcourt and

Delcourt 1991; Trapnell et al. 2007), and molecular phylogeographic studies have

repeatedly uncovered distinct lineages from here (Avise 2000; Soltis et al. 2006;

Burbrink et al. 2008; Cullingham et al. 2008; Fontanella et al. 2008; Guiher and Burbrink

2008; Morris et al. 2008). Florida likely served as an important refugium for more

temperate species during the glacial episodes of the Pleistocene, when sea levels were

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 135 -

lower and the exposed land mass was more than twice as large (Webb 1990).

Furthermore, although Florida was not likely recently separated from the mainland by a

sea channel, the presence of boreal spruce/pine vegetation at the base of the peninsula

may have served as an effective barrier to peninsular forms during the Pleistocene (Webb

1990). Northern Florida, southern Georgia, and Alabama appear to be regions of

secondary contact between lineages located in refugia in Florida and east Texas­

Louisiana (Remington 1968; Swenson and Howard 2005). A. g. dorsalis conforms to this

pattern: the only two individuals sampled from Collier in the lower peninsula bore very

distinct haplotypes, while Putnam, located within the zone of secondary contact in the

northern peninsula, contained haplotypes from both groups. In spite of relatively

comprehensive sampling, no A. g. dorsalis haplotypes were found elsewhere; it is thus

probable that this lineage is restricted to the Florida peninsula, a pattern similar to that of

the ringneck snake Diadophis punctatus (Fontanella et al. 2008). It therefore appears

likely that A. g. dorsalis represents a divergent lineage that was isolated in Florida

throughout the Pleistocene and perhaps earlier, while the A. g. gryllus lineage survived in

a separate refugium, likely the Gulf and/or Atlantic Coastal Plains (Church et al. 2003;

Zamudio and Savage 2003; Austin et al. 2004). Although A. g. dorsalis may not be

distinct from A. g. gryllus in call characteristics (Nevo and Capranica 1985) or allozymes

(Dessauer and Neva 1969), the mitochondrial data, combined with its distinctive

morphology and behaviour (Netting and Goin 1945; Neill 1950) and the unique

biogeography of the Florida peninsula, indicate that A. g. dorsalis should remain a

separate subspecies until additional information becomes available.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 136 -

Species distributions and biogeographic patterns

1. Acris blanchardi

Substantial adjustment to the distribution of Acris species are suggested by the

mitochondrial data. As proposed here, A. blanchardi is the only species west of the

Mississippi River, and its northeastern range is bordered by the Ohio River to the south.

This species is found ubiquitously throughout this region, but does not appear to extend

much into Kentucky and was not found at all in Tennessee, as described by Conant and

Collins (1998). The lower Mississippi and Ohio Rivers appear to be major barriers to

gene flow in this species. A. blanchardi is a relatively small (maximum SVL = 3.8 cm),

terrestrial/semi-aquatic anuran that preferentially occupies the shorelines of nearly any

type of fresh water, with the exception oflarge lakes and wide rivers (Gray et al. 2005).

It is therefore likely that large rivers such as these prevent extensive dispersal in this

species. The Mississippi River has been considered a major force in the speciation and/or

restriction of gene flow in numerous taxa, including amphibians (Soltis et al. 2006;

Zeisset and Beebee 2008). As was found by Burbrink et al. (2008) for the snake Coluber

constrictor, however, the upper Mississippi River does not appear to be a significant

barrier to gene flow (see also Chapter 4). Though it is less frequently reported as a

barrier, the Ohio River does correspond to phylogenetic breaks in several amphibians and

reptiles (Starkey et al. 2003; Lemmon et al. 2007b; Burbrink et al. 2008).

Although major rivers are considered to be significant barriers to gene flow in

many species, they are often not impermeable (Burbrink et al. 2000; Zamudio and Savage

2003). This appears to be true for A. blanchardi, as individuals were found on opposing

sides of both the Mississippi and Ohio Rivers. The observation of nearly 30 A.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 137 -

blanchardi individuals at seven sites east of the Mississippi River in Louisiana and

Mississippi (see also Gamble et al. 2008) suggests that it is well-established here. How

has such a small vertebrate been able to cross this formidable barrier? Repeated dispersal

via rafting is one possibility (Schiesari et al. 2003), and is supported by reports of A.

blanchardi calling from floating vegetation over deep water some distance from the shore

(Gray et al. 2005). However, it is striking that the eastern extent of A. blanchardi

corresponds to the boundary of the Mississippi Delta (Figure 5.7), the western portion of

the alluvial floodplain of the lower Mississippi River. The lower Mississippi Valley has

experienced repeated flooding, variation in sediment load, and shifting of channels

throughout the Pleistocene due to outwash from melting glaciers (Saucier 1994), and may

be responsible for A. blanchardi' s presence on both sides of the river. A similar scenario

was proposed for a lineage of the snake Diadophis punctatus (Fontanella et al. 2008).

Furthermore, on the basis of earlier morphological studies (Viosca 1944; Boyd 1964) and

its absence from the Delta during this study (Figure 5.7), A. gryllus may be excluded

from this region (although more detailed sampling may indicate otherwise). It is possible

that the transition from alkaline alluvial soils in the Delta to the more acidic sedimentary

rocks of the Coastal Plain to the east (Viosca 1944; Miller and White 1998) maintains the

separation of these two species, as preferences for acidic and alkaline conditions have

been shown to partition them within individual sites (Viosca 1944; Blair 1958; but see

Mecham 1964). The central lineage of the snake Agkistrodon controtrix spanned the

Mississippi River and was limited by the ecological transition from forest to prairie

(Guiher and Burbrink 2008), suggesting that features other than the Mississippi River

may be important in defining biogeographic boundaries for some species.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 138 -

• • • • • • • • ... •

• • • • • • • ......

• •

N 0 100 200 300 400 Kilometres A

Figure 5.7. The Mississippi River Alluvial Plain (shading). The Mississippi Delta is the section of the alluvial plain that extends into western Mississippi. Localities containing A. blanchardi specimens are identified as circles and A. gryllus specimens as triangles. Open circles are samples identified by Gamble et al. (2008) as A. blanchardi using the cytochrome b gene.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 139 -

The distribution of A. blanchardi haplotypes predominantly throughout the

Interior Plains (Figure 5.8) is consistent with the identity of this anuran as a grasslands

species that prefers open vegetation (Beasley et al. 2005; Lehtinen and Skinner 2006).

However, it was also found in the more heavily forested areas of the Ozark Plateau and

Ouachita Mountains, as well as the Coastal Plain. Overall, A. blanchardi appears to be a

generalist in its requirements and is able to survive in a variety of habitats. Its

distribution does not seem to be substantially limited by physiography or ecology, and

large rivers are not always barriers to dispersal, as discussed previously. It may be that a

specific combination of geological and ecological features, which is beyond the scope of

my study, is the primary determinant of this species' distribution.

Within A. blanchardi, the relationship between many haplotypes was unresolved,

with low sequence divergences and short branch lengths compared to those found within

A. crepitans and A. gryllus. This pattern of broadly distributed lineages with shallow

divergence west of the Mississippi River has been found in other amphibians and reptiles

(Shaffer and McKnight 1996; Burbrink et al. 2008; Fontanella et al. 2008). There are two

possible causes of this. A. blanchardi may have diverged only recently from its common

ancestor with A. crepitans, and is thus a very young species. It is also possible that A.

blanchardi recently experienced a bottleneck to one or a few small populations, which

would have eliminated much of its genetic variation (Hoelzel et al. 1993; Frankham et al.

2002; Waldick et al. 2002; England et al. 2003). The lack of genetic structure north of

the Missouri River, combined with detailed analyses of the mitochondrial data (Chapter

4), demonstrate that the Midwest has indeed experienced multiple bottlenecks, likely

caused by repeated elimination of populations by Pleistocene glaciations. In contrast, the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 140 -

D Valley and Ridge D Appalachian Plateau Ill BlueRidge CJ Central Lowland D Great Plains D Interior Low Plateau D Ouachita Mountains D Ozark Plateau D Piedmont D Coastal Plain

N 0 200 400 600 800 Kilometres A

Figure 5.8. Major physiographic divisions of the United States. Localities of Acris specimens are identified as follows: A. blanchardi = circles; A. crepitans = squares; A. gryllus = triangles. The generalised glacial limits for the pre-Illinoian (dashed line) and Wisconsinan glaciations (solid line) are also shown (Reed and Bush 2005).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 141 -

presence of several clades south of the Missouri River, the likely use of this area as a

refugium during the Pleistocene (Chapter 4), and the climatic and habitat stability of the

southern Great Plains for the past -1.6 Ma (Graham 1999), make it probable that this

region has had relatively large, stable populations for some time. It therefore seems

likely that the shallow genetic structure in A. blanchardi is primarily the result of recent

population expansion in the Midwest, but that this pattern may be superimposed on a

short evolutionary history in the southern Great Plains.

Four individuals possessing A. blanchardi haplotypes were found substantially

outside the distribution of this species, in St. Mary's MD and Charles City VA. One

possible explanation for these disjunct individuals is that A. blanchardi once had a

continuous distribution from Ohio through the Appalachians, and a relict population

remained along the northeast coast following range contraction. However, three of the

haplotypes clustered strongly with Louisiana haplotypes in the phylogenetic tree (Figure

5.4), which is not consistent with this hypothesis. A similar scenario was proposed to

explain speciation patterns in trilling frogs (Pseudacris), but was also rejected by

phylogenetic analysis (Lemmon et al. 2007a). Incomplete lineage sorting can cause the

retention of shared polymorphisms from a common ancestor in closely related species,

usually due to random chance and/or insufficient time to achieve reciprocal monophyly.

Under this scenario, the lineages will necessarily coalesce to a time before the species

diverged, and thus will usually cluster together to the exclusion of other existing lineages

(A vise 2000; Cullingham et al. 2008). The lack of clustering of these disjunct haplotypes

and their high similarity to haplotypes in Louisiana (0-0.0021 sequence divergence)

makes incomplete lineage sorting unlikely and is more consistent with recent gene flow

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 142 -

(A vise 2000), but A. blanchardi' s low dispersal capacity (Pyburn 1958; Burkett 1984)

renders this also improbable. A previously continuous distribution from Louisiana or

natural introgression with another species may have enabled these haplotypes to become

established in the northeast. However, these scenarios would require extensive extinction

of all intervening individuals and populations, as no A. blanchardi haplotypes were found

in any other A. crepitans or A. gryllus. Furthermore, only A. gryllus has a nearly

continuous distribution between Louisiana and these individuals, and it is less probable

that A. blanchardi would hybridise extensively with this more distantly related species

than A. crepitans. Given that so few disjunct individuals were observed here, the

simplest and most likely explanation is that these haplotypes became introduced to the

northeast coast via human-mediated translocations. Small frogs are commonly used as

fishing bait in North America, and this has been cited as a potential contributor to the

extirpation of A. blanchardi populations in Canada (Oldham 1992). It is unknown

whether the haplotypes are from translocated individuals, or are the descendants of earlier

translocations. The fact that A. blanchardi occurs at the same latitude and occupies a

wide range of physiographic regions, including the Coastal Plain, implies that individuals

are likely able to survive here. If the translocation occurred some time ago, it is possible

that these haplotypes have become introgressed into the resident A. crepitans population.

A. crepitans and A. gryllus are known to be compatible (Mecham 1964), and the closer

phylogenetic relationship of A. crepitans and A. blanchardi suggests that they are likely

also interfertile. If it occurred very recently, however, it is possible that these individuals

are pure A. blanchardi. Assessment of these disjunct individuals at nuclear loci will

determine whether they are hybrids, and thus clarify the relative timing of the event.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 143 -

2. Acris crepitans

Along the Atlantic coast, the distribution of A. crepitans is similar to that given by

Conant and Collins (1998). In contrast, the western portion of its range deviates

substantially from its currently accepted distribution. No individuals were found west of

the Alabama/Coosa Rivers or in southern Alabama. Although the latter area is

represented by a small sample size at one site, only a single individual was observed in

the Florida panhandle despite profiling nearly 25 samples from several sites. Similar

results have been found in other studies (Gamble et al. 2008), indicating that if A.

crepitans does exist throughout the Gulf Coastal Plain, it is indeed rare. In contrast with

its currently accepted range, this species occurred entirely throughout Kentucky and

Tennessee, with the exception of extreme western Tennessee. The limit of A. crepitans at

the Ohio and Alabama/Coosa Rivers suggests that large rivers are significant barriers to

dispersal for this species. However, similar to A. blanchardi, the Ohio River has not been

completely impassable, and several individuals were found in Indiana and Ohio.

Acris crepitans is generally considered to be an upland species (Neill 1950;

Conant and Collins 1998), and the distribution of haplotypes primarily throughout the

Piedmont and Interior Low Plateau confirms this (Figure 5.8). It was also found in the

lowlands of the Atlantic Coastal Plain from eastern South Carolina to New Jersey,

although it may simply be filling the Coastal Plain niche occupied by A. gryllus further

south (see below) rather than being particularly well-suited for this region. Overall, A.

crepitans occurs only rarely below the Fall Line, the transition zone between the upland

Piedmont region and the lowland Coastal Plain (Figures 5.2 and 5.8). The Fall Line is

considered a major barrier to dispersal for many small aquatic and terrestrial vertebrates

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 144 -

(Aresco 1996; Pauly et al. 2007; Rahel 2007) and marks phylogenetic discontinuities in

some species (Lemmon et al. 2007b; Fontanella et al. 2008). It thus appears that this is a

major biogeographic boundary for A eris, separating upland A. crepitans from lowland A.

gryllus. A. crepitans may be limited by very high elevations, however, as it does not

occur extensively within the highest parts of the Appalachians (Conant and Collins 1998;

Gray and Brown 2005): the Blue Ridge, Valley and Ridge, and Appalachian Plateaus

(Figure 5.8), all of which have elevations in excess of 1200 m (Thornbury 1965). Rather,

it is limited to the lower elevation Piedmont (foothills) in the east and parts of the

Cumberland Plateau (eastern Tennessee/Kentucky) in the west.

Two divergent, well-supported clades were found within A. crepitans,

corresponding to the east and west of its distribution. The western group occurred

primarily in the Interior Low Plateau and western Appalachian Plateau (Figure 5.8),

where it was bounded to the north by the Ohio River and to the west, south and east by

the Tennessee River. The eastern group ranged across the Piedmont and Atlantic Coastal

Plain, east of the Alabama/Coosa and Tennessee Rivers. There is very little divergence

within each clade, and little exchange between them, with the exception of single eastern

individuals found at two western sites (Cumberland TN and Adams OH). Divergence

between the two clades is greater than is typically found within amphibian species ( e.g.

A. blanchardi; Shaffer et al. 2004b; Smith and Green 2004; Rowe et al. 2006), but is

similar to the currently recognised subspecies A. g. gryllus and A. g. dorsalis. Given the

strong phylogenetic clustering, large contrast of within- and between-clade variation,

nearly complete geographic separation, concordance with potential geographic barriers to

gene flow, and comparable divergence to subspecies of the closely related A. gryllus, I

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 145 -

recommend that the eastern and western groups be considered distinct subspecies.

The presence of an eastern and western lineage in A. crepitans is concordant with

the pattern observed in a number of eastern North American amphibians. This is often

attributed to fragmentation in isolated refugia during Pleistocene glaciations, most

commonly within the Appalachian Mountains, Coastal Plain, and Interior Low Plateau

(Austin et al. 2002, 2004; Church et al. 2003; Zamudio and Savage 2003; Hoffman and

Blouin 2004; Lee-Yaw et al. 2008). Glacial refugia are typically identified as unglaciated

regions possessing high diversity relative to other areas (Hewitt 1996; Bernatchez and

Wilson 1998; Hoffman and Blouin 2004; Lee-Yaw et al. 2008). Given its limited

contemporary distribution, the western lineage of A. crepitans was likely isolated within

the Interior Low Plateau, which was beyond the limits of the Pleistocene ice sheets

(Figure 5.8; Mickelson and Colgan 2004), and the Appalachian Mountains and

Ohio/Tennessee Rivers probably continue to be significant barriers to dispersal today.

Based on existing data, it is not possible to determine a specific refugial location for the

more wide-ranging eastern lineage. It is presently found only in unglaciated areas, and

small sample sizes preclude identification of the most diverse area. It is probable that this

lineage survived the Pleistocene glaciations in the southern Appalachians (e.g. North

Carolina, South Carolina, Georgia), for which nucleotide diversity appears to be highest,

and which is frequently considered to be a refugium for amphibians. The presence of

multiple divergent lineages in some amphibians has been attributed to isolation in several

refugia during the Pleistocene (Austin et al. 2002; Church et al. 2003; Hoffman and

Blouin 2004; Lee-Yaw et al. 2008), but the low diversity within each A. crepitans lineage

implies that only a single refugium was likely used for each.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 146 -

3. Acris gryllus

The distribution of A. gryllus corresponded well with its previously recognised

range. It occurred throughout the entire Gulf and Atlantic Coastal Plains east of the

Mississippi River (Figure 5.8) and as far north as the North Carolina/Virginia border.

This is consistent with its recognition as a lowland species (Neill 1950; Conant and

Collins 1998). As discussed previously, the Mississippi River and Fall Line appear to be

significant barriers to gene flow in A. gryllus. These features may have a greater impact

on A. gryllus than on A. blanchardi or A. crepitans: several individuals of this latter

species were found on the opposing sides of both riverine and geological barriers,

whereas A. gryllus was found at only two sites just above the Fall Line (Moore and

Durham NC), and appears to be excluded from sites within 30 km of the Mississippi

River. A. gryllus may be more restricted to a narrow habitat zone, and/or its smaller size

(particularly relative to A. blanchardi; Nevo 1973a; Conant and Collins 1998) may render

it less able to compete with closely related species.

In addition to a distinct Florida clade (A. g. dorsalis), eastern and western clades

were found within A. g. gryllus, completely separated by the Alabama and Coosa Rivers.

This contrasts with the east-west discontinuity corresponding to the Apalachiola or

Tombigbee Rivers seen in other Gulf Coastal Plain taxa (Soltis et al. 2006). Although

only a few sites were sampled on both sides of these rivers, a relatively large number of

individuals ( 14-23) were profiled, suggesting that this proposed boundary is robust. The

western group is monophyletic and contains several well-supported clades, while the east

primarily consists of shallow lineages with little overall structure. Two highly divergent

clades were present in North Carolina, although low bootstrap support and large sequence

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 147 -

divergence between them mean that they may be no more closely related to one another

than either is to other A. g. gryllus haplotypes. Other amphibians and reptiles have also

been shown to possess distinct lineages in the Carolinas (Howes et al. 2006; Lee-Yaw et

al. 2008). This phylogeographic pattern of multiple, divergent lineages is likely the result

of allopatric fragmentation in several refugia during Pleistocene glaciations, as has been

suggested for numerous eastern North American amphibians (Austin et al. 2002; Church

et al. 2003; Hoffman and Blouin 2004; Lee-Yaw et al. 2008). There may have been five

separate refugia for A. gryllus during the Pleistocene: the Florida peninsula (A. g.

dorsalis), the Gulf Coastal Plain in Mississippi (western A. g. gryllus), southern

Alabama/Georgia and the Florida panhandle (most eastern A. g. gryllus), and two

separate areas north of the Pee Dee River (North Carolina clades). There has been little

secondary contact with the western group following deglaciation, suggesting that the

Alabama/Coosa Rivers continue to act as a barrier to gene flow today. In contrast, the

North Carolina clades overlap at several sites and other eastern haplotypes are distributed

widely throughout the Coastal Plain, indicating that barriers no longer exist in this area.

The monophyly, genetic divergence, and geographic separation of western A. g.

gryllus suggest that this group should be treated as a distinct unit for management. In

contrast, the shallow genetic structure in the east, combined with little geographic

separation, does not presently permit identification of discrete population segments here.

The placement of A. g. dorsalis relative to eastern and western A. g. gryllus clades is also

ambiguous (see also Gamble et al. 2008). More comprehensive sampling and the use of

additional genetic markers may clarify the relationship among these groups, after which it

may be possible to recognise additional subspecies within A. gryllus.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 148 -

Comparison with patterns in previous studies

A number of studies have investigated the geographic variation in several

characteristics across the distribution of Acris. It is therefore worthwhile to evaluate the

concordance of the phylogeographic patterns observed here using mitochondrial data

with those seen in other traits. The recent study by Gamble et al. (2008) was largely

consistent with the results presented here, with some differences likely reflecting the

more comprehensive sampling employed in my study. As these differences have already

been discussed in relation to biogeographic patterns and species delimitation, I will not

address them further.

In 1969, Dessauer and N evo conducted an extensive survey of a number of blood

and liver proteins across the entire range of Acris. Based on the distribution of alleles,

they confirmed the reproductive isolation of A. gryllus from A. crepitans/blanchardi, a

pattern that is also supported by call characteristics (Blair 1958; Nevo and Capranica

1985; Micancin 2008), morphology (Neill 1950; Nevo 1973a; Micancin 2008), behaviour

(Neill 1950), ecology (Viosca 1944; Boyd·1964 Bayless 1969; Micancin 2008; but see

Neill 1950; Mecham 1964), and the current mitochondrial data. Dessauer and Nevo

(1969) also found that A. gryllus was nearly monomorphic, with no distinction between

A. g. gryllus and A. g. dorsalis or any of the additional lineages identified here. However,

this is not surprising, as sampling of A. gryllus was limited, and the slow evolutionary

rate of proteins can result in little diversity within species (Murphy et al. 1996).

Within A. crepitans, several allozymes defined a Plains, Delta, and Appalachian

group (Figure 5.9; Dessauer and Nevo 1969) that approximately corresponded to the

subspecies designations of A. crepitans sensu Conant and Collins ( 1998) (compare to

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 149 -

Figure 5.1 ). This is consistent with the distinctiveness of A. blanchardi (Plains group)

relative to eastern A. crepitans (Appalachian group) based on mtDNA data. Allelic

overlap between these groups at some loci is likely due to shared ancestry, which is also

demonstrated by the mitochondrial data, rather than contemporary gene flow. Even

greater allelic overlap between the Delta (A. c. paludicola) and Plains group (Figure 5.9;

Dessauer and Nevo 1969) supports the mitochondrial data in demonstrating that A. c.

paludicola is not different from A. blanchardi. However, the presence of a number of

unique alleles in the Delta group indicates that they are somewhat distinct, perhaps

reflecting rapid selection or isolation from other populations. Although the study by

Dessauer and N evo ( 1969) supports the recognition of three distinct species as proposed

here, it cannot be used to corroborate distribution adjustments as it did not sample many

of the relevant locations.

Acris are known to be highly polymorphic for the colour of both the vertebral

stripe (grey, green, red, red-green, or brown) and dorsum (grey, green, brown, or black)

(Pyburn 1961; Nevo 1973b; Burkett 1984; Gray 1995). At very small geographic scales,

the pattern of colour morphs may primarily reflect selection for cryptic colouration on

local substrates (which can also vary seasonally; Pyburn 1961; Nevo 1973b), apostatic

selection (Milstead et al. 197 4 ), or genetic drift in small populations (Pyburn 1961; Gray

1983, 1984; Gorman 1986). The vertebral stripe has also been observed to vary with

breeding activity, substrate, and moisture level in a single individual (metachrosis;

Pyburn 1961; Gray 1972). Over the entire distribution of Acris, Nevo (1973a) observed

that the frequency of colour morphs differed significantly between A. gryllus and A.

crepitans/blanchardi. Rather than fixed differences between species, however, this

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 150 -

--1·~~"""' ~<..:C:j l v;;,._::- !Tr-.nw.-•-1--,c:;__--. . -o··

I

') ;. ')..,_ ~ _,

Figure 5.9. Geographic distribution of transferrins (A) and globin peptides (B) in populations of A eris. Pie graphs are centred over collecting sites, except for certain populations of A. gryllus. Total area of circle is proportional to the sample size. Shading depicts different allozyme variants. Lines delimit the geographic distribution of A. crepitans as recognised at the time of publication (from Dessauer and Nevo 1969).

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 151 -

appeared to be a clinal trait: southern and central populations of all species were di- or

trimorphic, while northern and western populations, to which only A. crepitans and A.

blanchardi belong, were monomorphic grey. The greater frequency of green and red

morphs in A. gryllus may be due to the more variable and vegetated habitats in the south,

while the large grey morph appears to be adaptively superior in the arid environment of

the northern and western periphery as it is more resistant to desiccation and disease

(Nevo 1973a,b ). Random fixation was regarded as an unlikely cause for the prevalence

of grey in the north because of the large size of some populations (Nevo 1973b ).

However, it is probable that the founder effect associated with colonisation of this area at

the end of the Pleistocene is largely responsible for this monomorphism, which is also

evident in nuclear (Chapter 3) and mitochondrial (Chapter 4) DNA. This is supported by

the fact that the likely source of colonising animals, Kansas (Chapter 4), is also

predominantly grey (Nevo 1973b; Gorman 1986). Overall, it appears that colour patterns

in all A eris are influenced primarily by selection for both climatic variation and

predation, as well as the effects of colonisation, and do not reliably distinguish species.

Morphological investigation of the A eris complex has been extensive, with

species and range designations largely based on these characteristics. However,

numerous studies have documented inconsistencies in diagnostic features (Neill 1950;

Nevo 1973a; McCallum and Trauth 2006; Rose et al. 2006; Micancin 2008), with many

traits varying clinally (Nevo 1973a; McClelland et al. 1998; McCallum and Trauth 2006;

but see Ryan and Wilczynski 1991) and possibly reflecting adaptation to local climates or

habitats (Nevo 1973a). Overall, there appears to be much morphological variation within

taxa, which are also poorly differentiated from one another relative to other North

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 152 -

American hylids (Chantell 1968). These issues have resulted in much uncertainty in

specimen identification, reporting of potential hybrids, and proposed taxonomic

revisions. The mtDNA data described here indicate that one source of confusion is the

incorrect delimitation of each taxon's range; consequently, any conclusions based on

morphology using the previous distributions may be erroneous. For example, McCallum

and Trauth (2006) recommended that A. c. blanchardi be subsumed into A. c. crepitans,

as an extensive survey of morphological characters could not reliably distinguish the

subspecies in Arkansas. According to the distributions presented here, however, this

study actually illustrated that A. blanchardi is quite morphologically uniform, with

Arkansas individuals indistinguishable from those in South Dakota and Nebraska. Neill

(1950) also noted discrepancies in the defining characteristics of A. g. dorsalis, but the

source of these specimens in northern Florida indicates that they may have been A. g.

gryllus. Using phylogenetic analyses, Lemmon et al. (2007b) demonstrated that

hybridisation studies of chorus frogs (Pseudacris), the sister genus to Acris, yielded

similarly ambiguous or erroneous conclusions due to the presence of cryptic species.

Hence, the morphology of each group in the Acris complex should be re-evaluated in the

context of the updated distributions to clarify their diagnostic characteristics (e.g. Stuart

et al. 2006; Padial and de la Riva 2009).

Variation of call characteristics within and among species of Acris has been

investigated extensively. Cluster analysis of a number of temporal and spectral variables

grouped calls according to the three biomes occupied by Acris: grasslands (A.

blanchardi), deciduous woodlands (A. crepitans), and subptropical pineries (A. gryllus)

(Nevo and Capranica 1985). This supports the mitochondrial data presented here, and the

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 153 -

recognition of three distinct species. However, calls of A. blanchardi in eastern Texas,

which is dominated by pine forest, are distinct from grassland populations to the west:

they possess a number of characteristics to increase transmission, as the forests cause

severe degradation of sound (Nevo and Capranica 1985; Ryan et al. 1990; Ryan and

Wilczynski 1991). Furthermore, a population located in an isolated pine forest relict (the

"Lost Pines") had calls more similar to the geographically distant eastern populations

than to a grassland population only 65 km away (Ryan and Wilczynski 1988). It

therefore appears that the call of A. blanchardi (and possibly all Acris) can adapt quickly

for increased transmission efficiency in different habitats, which is likely responsible for

some of the documented variation in A. blanchardi calls. Intraspecific variation in call

characters has been shown to occur in other anuran species (Littlejohn and Roberts 1975;

Ryan et al. 1996; Wycherley et al. 2002), although no studies are known in which this has

been attributed to habitat differences. The convergence of advertisement calls among

widely separated anuran communities (Duellman and Trueb 1986) suggests that habitat

plays a significant role in shaping call characteristics among species, and there is little

reason to doubt that it is also important within species. The definition of anuran species

boundaries purely by call characteristics may therefore be unreliable, particularly across

broad distributions and when large habitat changes exist, as exemplified by Acris (but see

Wycherley et al. 2002).

Although samples from within the pine forest of eastern Texas were not profiled

here, two well-supported clades exist from this region (Figure 5.2): Clade 5.2, entirely

from Louisiana, and Clade 3, from Louisiana and eastern Texas (Anderson and Bowie).

When combined with the presence of a distinct call, this may be indicative of a

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 154 -

genetically distinct population. However, a number of haplotypes from this region

clustered with haplotypes from diverse areas or were poorly resolved, suggesting that it is

not isolated. Furthermore, evidence indicates that A. blanchardi females prefer calls with

low dominant frequencies regardless of whether they are from a local or foreign

population, and that this may increase gene flow in an east-west direction (Ryan et al.

1992; Wilczynski et al. 1992). It is therefore probable that the older, refugial nature of

populations in this region (Chapter 4) has simply resulted in greater genetic structure,

with geographically proximal populations clustering together. More detailed sampling of

this region and the use of higher resolution genetic markers (e.g. microsatellites) should

clarify whether habitat and call differences have generated genetic differences here.

The existence of potential hybrids and individuals with intermediate phenotypes

has been repeatedly reported (Netting and Goin 1945; Neill 1950; Mount 1975;

McCallum and Trauth 2006; J Lee, pers. comm.), and artificial crosses between

sympatric A. crepitans and A. gryllus demonstrated that these species are relatively

interfertile (Mecham 1964). However, it seems probable that hybridisation is actually

quite rare between A eris species. In spite of overlapping distributions between all pairs

of species (Figure 5.2), few sympatric sites were observed in this and other (Micancin

2008) studies. Furthermore, neither hybrid protein patterns (Dessauer and Nevo 1969)

nor intermediate vocalisations (Micancin 2008) have been observed. It is likely that the

general similarity and extensive morphological variability of this group has caused both

the misidentification of pure individuals, and the reporting of intermediates (Mecham

1964; Micancin 2008). Indeed, the use of morphology to identify hybrids of hylid

anurans such as Acris has been shown to be unreliable (Lamb and A vise 1987). A.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 155 -

crepitans and A. blanchardi also appear to be ecologically isolated from A. gryllus in

areas of overlap such as Louisiana, Mississippi, and Alabama, with discrete usage of food

items and microhabitat (Viosca 1944; Blair 1958; Boyd 1964; Bayless 1969; but see

Mecham 1964; Neill 1950). A number of studies also suggest that vocalisation is an

important pre-mating reproductive isolating mechanism in maintaining the identity of A.

gryllus and A. crepitans at sympatric sites (Blair 1958; Nevo and Capranica 1985;

Micancin 2008). Thus, although hybrids may form occasionally under certain

circumstances, this is not likely a major threat to the integrity of these species.

Conclusions

The mitochondrial data provide evidence for three species of Acris, but

adjustments to their currently accepted distributions are required. The distributions of

these species broadly agree with several potential biogeographic barriers, and are

concordant with patterns observed in other species. Furthermore, the limited nuclear

genetic data available supports these divisions (Dessauer and Nevo 1969; see Chapter 6

for more details). The existence of three species is therefore consistent with the

principles of genealogical concordance described by A vise and Ball (1990). Substructure

occurs within each species, particularly A. crepitans and A. gryllus, and some lineages

may be distinct subspecies. These putative subspecies also conform to the principles of

genealogical concordance; for example, the eastern and western lineages of A. crepitans

appear to be phylogenetically distinct units that are largely separated by the Appalachian

Mountains (i.e. allopatric ). In contrast, A. c. paludicola is neither phylogenetically

distinct nor allopatric from A. blanchardi and is thus not supported as a distinct

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 156 -

subspecies. More detailed information from nuclear genetic markers, morphology, and

the degree of reproductive compatibility between groups should be obtained to further

clarify the species or subspecies status of these taxa.

Phylogeographic patterns within this complex are primarily, but not exclusively,

influenced by large rivers and the elevation change at the Fall Line, and are corroborated

by several morphological, ecological, and enzyme studies. However, colour and call

characteristics may be strongly influenced by the environment and therefore are

unreliable as tools for species designation. Overall, Acris conforms to the "longitudinal"

pattern of genetic diversity seen in numerous amphibians and reptiles in North America

( e.g. Howes et al. 2006), supporting the existence of common barriers to gene flow for

small non-volant vertebrates. Although hybridisation between species cannot be

excluded where ranges overlap, it is likely rare owing to the observation of few sympatric

sites, differential habitat utilisation, and distinct call characteristics.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 157 -

Chapter 6

General Discussion

The genetic information obtained during the course of this study provided insight

into the systematics, biogeography, and genetic structure of P. lemur and the genus Acris,

which previously had been unclear when assessed using traditional characters such as

morphology, ecology, and call characteristics. In addition to contributing to the

understanding of regional biogeography and amphibian biology in general, this

information will greatly assist in directing future conservation activities for these

threatened anurans.

Conservation genetics of Peltophryne lemur

Conservation efforts for P. lemur began nearly 30 years ago, with only minimal

genetic data (Lacy and Foster 1987). It is not likely that initial recovery efforts for this

species would have been different had the genetic data presented here been available:

northern and southern populations were kept separate because they were considered

divergent, and the results of this study confirm this to some degree. However, it is this

divergence that suggests the presence of potentially vital adaptive variation in northern

individuals. Given the very small population size and highly inbred nature of the

remaining northern individuals, it is evident that their future in captivity or the wild is

precarious. The genetic information obtained during this study therefore provides the

rationale for the controversial recommendation of mixing divergent lineages in an

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 158 -

attempt to preserve the maximum amount of diversity, and thereby increase the survival

probability of P. lemur in northern Puerto Rico.

Synthesis ofAcris data and conservation genetics ofA. blanchardi

The abundant morphological variability within Acris, combined with ambiguity in

the definitive characteristics of each species, means that genetic data were essential to

determining the divergence among species and their distributions. Mitochondrial

sequences identified a number of inconsistencies in the currently accepted classification

and distribution of taxa, including recognition of A. blanchardi as a distinct species.

Although a small number of samples were profiled, nuclear genetic data supported this

revised taxonomy of Acris. The generally poor amplification of microsatellite loci in A.

crepitans (56%) implies that it is more divergent from A. blanchardi than a subspecies.

The closer relationship of A. blanchardi to A. crepitans than to A. gryllus is apparent in

the even weaker amplification of loci in A. gryllus (44%). Primmer and Merila (2002)

also found very poor amplification of microsatellite loci across several species of Rana

(13-31 %), particularly compared to birds, mammals and reptiles (Primmer and Merila

2002; Barbara et al. 2007). The greater success of cross-species amplification seen in this

study implies that Acris species are more recently diverged than the species of Rana

studied by Primmer and Merila (2002).

A striking difference of A. blanchardi relative to A. crepitans and A. gryllus is its

shallow genetic structure, despite occurring across a much larger area. This pattern of

limited divergence west of the Mississippi River relative to the east has been seen in a

number of amphibians and reptiles (Austin et al. 2002; Starkey et al. 2003; Hoffman and

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 159 -

Blouin 2004; Lemmon et al. 2007b; Burbrink et al. 2008; Fontanella et al. 2008), and is

likely due to a combination of factors. The less topographically varied terrain of the

Great Plains region, in contrast to the more dissected east, means that there are probably

fewer barriers to gene flow such as elevation or habitat transitions; as a consequence,

gene flow can occur to a greater degree and there are fewer opportunities for isolation

and divergence of populations. This same argument can be invoked in terms of the

Pleistocene glaciations: it is probable that populations existed in a single, broad refugium

in the southern Great Plains, rather than the multiple refugia that are frequently proposed

for amphibians and reptiles in the Coastal Plains and Appalachian Mountains (Austin et

al. 2002; Church et al. 2003; Zamudio and Savage 2003; Hoffman and Blouin 2004;

Pauly et al. 2007; Burbrink et al. 2008; Fontanella et al. 2008). Finally, the greater

southward extent of the Pleistocene ice sheets west of the Mississippi River (Richmond

and Fullerton 1986; Figure 5.8) means that populations would have contracted further

south, potentially compressed into a much smaller area. As a corollary, a greater portion

of the Great Plains would have been colonised more recently than eastern North America,

which would further reduce the extent of genetic structure within A. blanchardi relative

to A. crepitans and A. gryllus.

Genetic profiling of A eris identified a number of issues that can contribute to the

efficient conservation of A. blanchardi at the northern edge of its range. Above all, the

elevation of this anuran to species status makes population declines more alarming and

increases the urgency of conservation efforts, as the number of populations, geographical

distribution, and genetic diversity have been considerably reduced. It may also assist in

understanding the causes of declines. The abundance and apparent stability of this

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 160 -

species in southern regions indicates that factors such as habitat loss and alteration,

environmental contaminants, introduced species, and climate change, which are problems

common to all areas, cannot fully explain the declines. The significantly lower genetic

diversity at both nuclear and mitochondrial markers in northern populations suggests that

this may be interacting with external agents to cause declines. Populations of A.

blanchardi were resistant to pesticides in the Mississippi Delta (Boyd et al. 1963;

Ferguson 1963) and Texas (Flickinger et al. 1980), and were relatively unaffected by

high levels of metal accumulation in Texas (Clark et al. 1998). In contrast, persistent

organic pollutants were implicated in morphological abnormalities and population

declines in Illinois and Ohio (Reeder et al. 1998, 2005; Russell et al. 2002). Genetic

composition may affect resistance to environmental contaminants in amphibians (Bridges

and Semlitsch 2001 ), which suggests that the greater genetic variation of southern A.

blanchardi populations may have enabled them to adapt to such stressors, whereas

genetically depauperate northern populations have been unable to cope.

Mitochondrial sequences provided evidence of a distinct northern clade that

largely coincides with the region of decline, and nuclear DNA data support this. Primers

designed for at least two microsatellite loci, which were isolated from Texas animals,

amplified successfully and were polymorphic in Texas individuals. However, they failed

to amplify individuals from Ohio and Illinois ( data not shown), evidence that these

regions have been separated for sufficient time for mutations to have occurred in at least

one priming site ( e.g. Primmer and Merila 2002). In addition, a number of unique

microsatellite alleles were found in both Texas and Illinois, despite much less diversity in

the north. The location of the declining populations at the range margin makes it

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 161 -

probable that they experience highly variable conditions and therefore may be a source

for speciation (Lesica and Allendorf 1995; Garcia-Ramos and Kirkpatrick 1997).

Furthermore, the likelihood of adaptive variation in the north is substantial owing to the

different climate of the region and known differences in life history (e.g. overwintering,

breeding season, drought tolerance). These populations may therefore be vital to the

long-term evolutionary potential of the species as a whole, and are worthy of

conservation efforts (Hunter and Hutchinson 1994; Lesica and Allendorf 1995).

The genetic information for Acris presented here can be used to establish specific

guidelines for conservation efforts. For example, translocations are frequently used to

restore extirpated populations or augment existing ones, and the most appropriate source

populations can be identified using genetic data. Movement of animals should occur

between populations from within the same phylogenetic clade, where possible. This will

minimise the potential for outbreeding depression that may occur if animals with very

different life histories are used (Frankham et al. 2002), as well as maintain any adaptive

variation that may be necessary for survival at the northern edge of the range. Thus,

despite the much greater abundance of A. blanchardi in southern areas such as Texas,

animals should not be taken from here and used for supplementation in the Upper

Midwest. Spatial analysis also indicated that individuals separated by more than 450 km

within Group 1, which encompasses the declining northern populations, were not

correlated at mitochondrial haplotypes. The most conservative recommendation would

therefore be to use populations from within 450 km of the intended recipient population.

Another desirable attribute of source animals is that they be genetically diverse to

increase variation in the recipient populations, which will render populations more

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 162 -

resistant to environmental changes as well as minimise the deleterious effects of

inbreeding and random genetic drift (Frankham et al. 2002). Again, the genetic data

presented here can identify the most diverse region within 450 km of a potential release

site from which source animals could be obtained. Phylogeographic analysis also

demonstrated that major rivers and large changes in elevation are barriers to gene flow, to

various degrees, within and among all species of Acris. This is another issue that should

be addressed when evaluating potential source populations for translocations, as

individuals located on opposing sides of such features are likely to be more differentiated

than those on the same side.

A hypothetical scenario will best illustrate this in practice. If the population in

Lenawee Co. in southeast Michigan was to be augmented, sites within 450 km from this

location range west to north-central Illinois and south to the Ohio River. The most

diverse population from this region is Greene Co., IN, with a nucleotide diversity of

0.0025 (Appendix 1). However, this is still not a very diverse site. It may be decided

that using the greatest genetic diversity possible is more important than precisely

maintaining population structure at this scale. The potential source region could

therefore be extended to incorporate all of phylogenetic Clade I or SAMOVA Group 1,

which includes Missouri, Illinois, Iowa, Nebraska, and eastern Kansas. Although only a

few sample sites within this region are similar in diversity to southern sites ( e.g.

Anderson KS; Appendix 1), most possess different haplotypes that would collectively

increase the genetic variation if multiple sites were incorporated. An approach such as

this also uses animals from a similar climate and environment, therefore ensuring that

animals are likely to be ecologically exchangeable with the recipient population (Crandall

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 163 -

et al. 2000). It would also minimise the potential for transfer of locally endemic diseases

between geographically distant regions.

The recognition of A. blanchardi as a species, and the presence of a distinct

northern clade, suggests that this species should be reevaluated under the Endangered

Species Act in the US (it is already listed as endangered in Canada; Kellar et al. 1997).

The status of A. crepitans and A. gryllus may also need adjustment, as some concern

exists over declining populations in these species as well (Gray and Brown 2005; Jensen

2005; Micancin 2008) and their highly structured nature may increase the urgency of

conservation efforts for certain lineages. This study also demonstrates the importance of

using a large number of samples and populations, if possible, to establish guidelines for

conservation efforts. In addition to more precisely and robustly defining species

boundaries and biogeographic barriers to gene flow, a number of locations were

identified where caution should be exerted during management. For example, the

observation that A. blanchardi is the dominant species throughout the Mississippi Delta

was unexpected, as other phylogeographic studies have shown the Mississippi River to be

a major barrier to dispersal for small vertebrates (Soltis et al. 2006). The Ohio River also

appears to be a clearly defined boundary between A. blanchardi and A. crepitans;

however, both species were observed at sites on opposite sides, something that less

complete sampling probably would have missed. This means that populations adjacent to

biogeographic barriers between closely related species should be avoided for

management purposes, as it is possible that the barriers are not absolute and populations

have become established outside their expected distributions.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 164 -

Value ofgenetics to conservation biology

Molecular methods have been used to uncover genetic structure and resolve

phylogenetic uncertainties in many species, but may be a particularly important tool for

morphologically conservative groups such as amphibians (A vise 1994; Zamudio and

Savage 2003; Stuart et al. 2006; Lemmon et al. 2007b; Vredenburg et al. 2007; Angulo

and Reichle 2008; Elmer and Cannatella 2008; Padial and de la Riva 2009). The genetic

data obtained during this study have been essential to clarifying a number of important

details regarding the biology, life history, and evolution of both P. lemur and Acris. In

tum, this information has provided the foundation to develop comprehensive restoration

programs for each species, including specific guidelines for future activities and the

justification of recommendations that may be viewed as controversial.

The data described in this study represent only a fraction of the information that

can be obtained with genetic markers and used to assist future conservation activities.

The highly polymorphic microsatellites developed here are particularly amenable to an

array of studies that can further elucidate the biology of these species. For example,

powerful statistical methods such as assignment tests can be used with microsatellite data

to determine fine-scale population structure, patterns of dispersal, and hybridisation

(Berry et al. 2004; Mane! et al. 2005). The effective size of a population, a more

biologically relevant measure of a population's stability and future potential than its

census population size, can be estimated using genetic markers (Frankham 1995;

Schwartz et al. 1998). Incorporating detailed maps of landscape features can reveal the

variables that are critical to the species for feeding, reproduction, dispersal, etc. (Mane! et

al. 2003; Storfer et al. 2007). These molecular markers also have immediate, practical

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 165 -

applications for species conservation. I recently assisted with such a situation, in which I

used two of the population-specific P. lemur micro satellite loci described here to

determine the population of origin of several captive individuals that had become

misplaced during handling. Multilocus microsatellite profiles could also be used as

individual genetic "tags" in mark-recapture studies or to monitor the reproductive success

of individuals translocated to a population (Schwartz et al. 2007). At the broadest scale,

such information could be used to design protected reserves or corridors in a fragmented

landscape (Bell and Okamura 2005; Beier et al. 2008; Vandergast et al. 2008; Cushman

et al. 2009), efforts that would benefit multiple species.

Although much information can be gleaned from neutral molecular markers such

as mitochondrial DNA and microsatellite loci, it is the adaptive variation and fitness

attributes that are important to both the evolutionary potential and conservation of a

species. Some studies have shown that variation and/or divergence at quantitative traits

is very poorly correlated with that for neutral molecular markers (Reed and Frankham

2001; McKay and Latta 2002; Holdregger et al. 2006; Knopp et al. 2007), some have

found a good relationship between these (Merila and Cmokrak 2001; Cmokrak and

Merila 2002), and others show that microsatellite heterozygosity and fitness measures are

positively related (Rowe et al. 1999; Reed and Frankham 2002; Lesbarreres et al. 2005;

Charpentier et al. 2005; Hensen and Wesche 2006; Vandewoestijne et al. 2008). Future

work in P. lemur and A eris should evaluate whether populations that are distinct at

neutral markers are also adaptively different from other populations, and whether the low

genetic diversity seen in some populations is indeed manifested as reduced fitness. Both

species represent excellent opportunities for such studies. The highly controlled nature of

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 166 -

P. lemur in captivity could enable controlled experiments to measure the influence of

various environmental factors, such as temperature and precipitation, on individual

fitness. In contrast, the large distribution and abundance of Acris is suitable to evaluating

the fitness of these anurans in response to variables such as environmental gradients,

distance from core populations, habitat alteration, or interaction with other species.

Conclusion

Conservation biology is often crisis discipline, with restoration activities initiated

only after a critical point has been reached and few options remain. In addition, most

conservation decisions must address not only the biological aspects of the species in

question, but also legal, political and economic criteria. At the heart of all conservation

decisions, therefore, is a trade-off between these factors (Shogren et al. 1999; Song and

M'Gonigle 2001; Lundquist and Granek 2005). Genetic data can provide insight into

many unknown biological aspects of a species that may be essential for understanding

causes of declines, and can aid in designing the most appropriate recovery plan.

Furthermore, once molecular markers are developed, they can be used to continue

monitoring populations and the success of an implemented plan (Schwartz et al. 2007).

Incorporating genetic data at the earliest possible stage therefore allows the execution of

a carefully planned, comprehensive program that can evaluate the success or failure of

certain activities within an experimental framework (Hunter and Hutchinson 1994;

Moritz 1999; Trenham and Marsh 2002; McCarthy and Possingham 2007; Seddon et al.

2007). This will provide valuable data that can be applied to other endangered species,

particularly those for which few individuals and little time remains.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 167 -

Literature Cited

Alford RA, Richards SJ (1999) Global amphibian declines: a problem in applied ecology. Annual Review of Ecology and Systematics 30:133-165.

Amato ML, Brooks RJ, Fu J (2008) A phylogeographic analysis of populations of the wood turtle (Glyptemys insculpta) throughout its range. Molecular Ecology 17:570- 581.

Angulo A, Reichle S (2008) Acoustic signals, species diagnosis, and species concepts: the case of a new cryptic species of Leptodactylus (Anura, Leptodactylidae) from the Chapare region, Bolivia. Zoological Journal of the Linnean Society 152:59-77.

Arens P, Bugter R, van't Westende W et al. (2006) Microsatellite variation and population structure of a recovering tree frog (Hyla arborea L.) metapopulation. Conservation Genetics 7:825-835.

Aresco MJ (1996) Geographic variation in the morphology and lateral stripe of the green treefrog (Hyla cinerea) in the southeastern United States. American Midland Naturalist 135:293-298.

Austin JD, Lougheed SC, Neidrauer L, Chek AA, Boag PT (2002) Cryptic lineages in a small frog: the post-glacial history of the spring peeper, Pseudacris crucifer (Anura: Hylidae). Molecular Phylogenetics and Evolution 25:316-329.

Austin JD, Davila JA, Lougheed SC, Boag PT (2003a) Genetic evidence for female­ biased dispersal in the bullfrog, Rana catesbeiana (Ranidae ). Molecular Ecology 12:3165-3172.

Austin JD, Lougheed SC, Moler PE, Boag PT (2003b) Phylogenetics, zoogeography, and the role of dispersal and vicariance in the evolution of the Rana catesbeiana (Anura: Ranidae) species group. Biological Journal of the Linnean Society 80:601-624.

Austin JD, Lougheed SC, Boag PT (2004) Discordant temporal and geographic patterns in maternal lineages of eastern North American frogs, Rana catesbeiana (Ranidae) and Pseudacris crucifer (Hylidae). Molecular Phylogenetics and Evolution 32:799-816.

Avise JC, Nelson WS (1989) Molecular genetic relationships of the extinct dusky seaside sparrow. Science 243:646-648.

A vise JC (1989) A role for molecular genetics in the recognition and conservation of endangered species. Trends in Ecology and Evolution 4:279-281.

A vise JC, Ball RM Jr. (1990) Principles of genealogical concordance in species concepts and biological taxonomy. Oxford Surveys in Evolutionary Biology 7:45-67.

Avise JC (1994) Molecular Markers, Natural History and Evolution. Kluwer Academic Publishers, Boston. 511 pp.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 168 -

Avise JC, Walker D (1999) Species realities and numbers in sexual vertebrates: Perspectives from an asexually transmitted genome. Proceedings of the National Academy of Sciences 96:992-995.

A vise JC (2000) Phylogeography: The History and Formation of Species. Harvard University Press, Cambridge, MA. 447 pp.

Baird SF (1854) Descriptions of new genera and species of North American frogs. Proceedings of the Academy of Natural Sciences of Philadelphia 7:59-62.

Ball SJ, Adams M, Possingham HP, Keller MA (2000) The genetic contribution of single male immigrants to small, inbred populations: a laboratory study using Drosophila melanogaster. Heredity 84:677-684. Ballou JD, Lacy RC (1995) Identifying genetically important individuals for management of genetic diversity in pedigreed populations. Pp. 76-111 in: Population Management for Survival and Recovery: Analytical Methods and Strategies in Small Population Conservation (eds. Ballou JD, Gilpin M, Foose TJ). Columbia University Press, New York.

Bandelt HJ, Forster P, Rohl A (1999) Median-joining networks for inferring intraspecific phylogenies. Molecular Biology and Evolution 16:37-48.

Barbara T, Palma-Silva C, Paggi GM et al. (2007) Cross-species transfer of nuclear microsatellite markers: potential and limitations. Molecular Ecology 16:3759-3767. Barber D (2007) Cooperative Amphibian Programs in AZA: Puerto Rican Crested Toad SSP. Connect February: 13.

Barbour T (1917) Notes on the herpetology of the Virgin Islands. Proceedings of the Biological Society of Washington 30:97-104.

Barinaga M (1990) Where have all the froggies gone? Science 247:1033-1034.

Barribeau SM, Villinger J, Waldman B (2008) Major histocompatibility complex based resistance to a common bacterial pathogen of amphibians. PLoS ONE 3: 1-9.

Bayless LE ( 1969) Ecological divergence and distribution of sympatric A eris populations (Anura: Hylidae). Herpetologica25:181-187.

Beamer DA, Lamb T (2008) Dusky salamanders (Desmognathus, Plethodontidae) from the Coastal Plain: Multiple independent lineages and their bearing on the molecular phylogeny of the genus. Molecular Phylogenetics and Evolution 47:143-153.

Beasley VR, Faeh SA, Wikoff B et al. (2005) Risk factors and declines in northern cricket frogs (Acris crepitans). Pp. 75-86 in: Amphibian Declines: The Conservation Status of United States Species ( ed. Lannoo MJ). University of California Press, Berkeley, CA.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 169 -

Beauclerc KB, Johnson B, White BN (2007) Characterization, multiplex conditions, and cross-species utility of tetranucleotide microsatellite loci for Blanchard's cricket frog (A eris crepitans blanchardi). Molecular Ecology Notes 7: 13 3 8-1341.

Beauclerc KB, Johnson B, White BN (in press) Genetic rescue of an inbred captive population of the critically endangered Puerto Rican crested toad (Peltophryne lemur) by mixing lineages. Conservation Genetics.

Beier P, Majka DR, Spencer WD (2008) Forks in the road: Choices in procedures for designing wildland linkages. Conservation Biology 22:836-851.

Bell JJ, Okamura B (2005) Low genetic diversity in a marine nature reserve: re­ evaluating diversity criteria in reserve design. Proceedings of the Royal Society B 272:1067-1074.

Bernatchez L, Wilson CC (1998) Comparative phylogeography ofNearctic and Palearctic fishes. Molecular Ecology 7 :431-452.

Berry 0, Tocher MD, Sarre SD (2004) Can assignment tests measure dispersal? Molecular Ecology 13:551-561.

Bertorelle G, Barbujani G (1995) Analysis of DNA diversity by spatial autocorrelation. Genetics 140:811-819.

Bickford D, Lohman DJ, Sodhi NS et al. (2006) Cryptic species as a window on diversity and conservation. Trends in Ecology and Evolution 22:149-155.

Blair WF (1958) Mating call in the speciation of anuran amphibians. American Naturalist 42:27-51.

Blair WF (1961) Calling and spawning seasons in a mixed population of anurans. Ecology 42:99-110.

Blaustein AR, Wake DB, Sousa WP (1994) Amphibian declines: judging stability, persistence, and susceptibility of populations to local and global extinctions. Conservation Biology 8:60-71.

Bloxam QMC, Tonge SJ (1995) Amphibians: Suitable candidates for breeding-release programmes. Biodiversity and Conservation 4:634-644.

Bohme MU, Schneeweib N, Fritz U, Schlegel M, Berendonk TU (2007) Small edge populations at risk: genetic diversity of the green lizard (Lace rta viridis viridis) in Germany and implications for conservation management. Conservation Genetics 8:555-563.

Bond AB (2007) The evolution of color polymorphism: Crypticity searching images, and apostatic selection. Annual Review of Ecology Evolution and Systematics 38:489-514.

Bos DH, Sites JW (2001) Phylogeography and conservation genetics of the Columbia spotted frog (Rana luteiventris; Amphibia, Ranidae). Molecular Ecology 10:1499- 1513.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 170 -

Boyd CE, Vinson SB, Ferguson DE (1963) Possible DDT resistance in two species of frogs. Copeia 1963:426-429.

Boyd CE ( 1964) The distribution of cricket frogs in Mississippi. Herpetologica 20:201- 202.

Bragg AN (1940) Observations on the ecology and natural history of Anura. III. American Midland Naturalist 24:322-335.

Brede EG, Beebee TJC (2004) Contrasting population structures in two sympatric anurans: implications for species conservation. Heredity 92: 110-117.

Bridges CM, Semlitsch RD (2001) Genetic variation in insecticide tolerance in a population of southern leopard frogs (Rana sphenocephala): implications for amphibian conservation. Copeia 2001:7-13.

Brock MK, White BN (1992) Application of DNA fingerprinting to the recovery program of the endangered Puerto Rican parrot. Proceedings of the National Academy of Sciences 89: 11121-11125.

Bryant EH, Backus VL, Clark ME, Reed DH (1999) Experimental tests of captive breeding for endangered species. Conservation Biology 13: 1487-1496.

Burbrink FT, Lawson R, Slowinski JB (2000) Mitochondrial DNA phylogeography of the polytypic North American rat snake (Elaphe obsoleta): A critique of the subspecies concept. Evolution 54:2107-2118. Burbrink FT, Fontanella F, Pyron RA, Guiher TJ, Jimenez C (2008) Phylogeography across a continent: The evolutionary and demographic history of the North American racer (Serpentes: Colubridae: Coluber constrictor). Molecular Phylogenetics and Evolution 47:274-288.

Burger WL, Smith PW, Smith HM (1949) Notable records of reptiles and amphibians in Oklahoma, Arkansas, and Texas. Journal of the Tennessee Academy of Science 24: 130-134. Burke RL ( 1991) Relocations, repatriations, and translocations of amphibians and reptiles: Taking a broader view. Herpetologica 47:350-357.

Burkett RD (1984) An ecological study of the cricket frog, Acris crepitans. Pp. 89-103 in: Vertebrate Ecology and Systematics: A Tribute to Henry S. Fitch ( eds. Seigel RA, Hunt LE, Knight JL, Malaret L, Zuschlag NL). Museum of Natural History, University of Kansas, Lawrence, KS.

Bums EL, Eldridge MDB, Houlden BA (2004) Microsatellite variation and population structure in a declining Australian Hylid. Molecular Ecology 13:1745-1757.

Bums EL, Eldridge MDB, Crayn DM, Houlden BA (2007) Low phylogeographic structure in a wide spread endangered Australian frog Litoria aurea (Anura: Hylidae). Conservation Genetics 8: 17-32.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 171 -

Burrowes PA, Joglar RL, Green DE (2004) Potential causes for amphibian declines in Puerto Rico. Herpetologica 60:141-154.

Bushnell RJ, Bushnell EP, Parker MV (1939) A chromosome study of five members of the family Hylidae. Journal of the Tennessee Academy of Science 14:209-215.

Campbell C (1978) Reproduction and ecology of turtles and other reptiles and amphibians of Lakes Erie and St. Clair in relation to toxic chemicals. CWS No. 7475/022, Canadian Wildlife Services.

Carpenter DW, Jung RE, Sites JJ (2001) Conservation genetics of the endangered Shenandoah salamander (Plethodon shenandoah, Plethodontidae ). Animal Conservation 4: 111-119.

Chantell CJ (1966) Late Cenozoic hylids from the Great Plains. Herpetologica 22:259- 264.

Chantell CJ (1968) The osteology of A eris and Limnaoedus (Amphibia: Hylidae ). American Midland Naturalist 79:169-182.

Chapuis MP, Estoup A (2007) Microsatellite null alleles and estimation of population differentiation. Molecular Biology and Evolution 24:621-631.

Charpentier M, Setchell JM, Prugnolle F et al. (2005) Genetic diversity and reproductive success in mandrills (Mandrillus sphinx). Proceedings of the National Academy of Sciences 102:16723-16728.

Chek AA, Lougheed SC, Bogart JP, Boag PT (2001) Perception and history: molecular phylogeny of a diverse group of neotropical frogs, the 30-chromosome Hyla (Anura: Hylidae). Molecular Phylogenetics and Evolution 18:370-385.

Church SA, Kraus JM, Mitchell JC, Church DR, Taylor DR (2003) Evidence for multiple Pleistocene refugia in the postglacial expansion of the eastern tiger salamander, Ambystoma tigrinum tigrinum. Evolution 57:372-383.

Clark DJ, Cantu R, Cowman DF, Maxson DJ (1998) Uptake of arsenic and metals by tadpoles at an historically contaminated Texas site. Ecotoxicology 7:61-67.

Collins JP, Storfer A (2003) Global amphibian declines: sorting the hypotheses. Diversity and Distributions 9:89-98.

Conant R, Collins JT (1998) A Field Guide to Reptiles and Amphibians of Eastern and Central North America, 3rd ed. Houghton Mifflin, New York, NY.

Cope ED (1868) Sixth contribution to the herpetology of Tropical America. Proceedings of the Academy of Natural Sciences of Philadelphia 20:305-313.

Cracraft J (1983) Species concepts and speciation analysis. Pp. 159-187 in: Current Ornithology (ed. Johnston RF). Plenum Press, New York.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 172 -

Crandall KA, Bininda-Emonds ORP, Mace GM, Wayne RK (2000) Considering evolutionary processes in conservation biology. Trends in Ecology and Evolution 15:290-295.

Crawshaw GJ (2000) Diseases and Pathology of Amphibians and Reptiles. Pp. 199-231 in: Ecotoxicology of Amphibians and Reptiles (eds. Sparling DW, Linder G, Bishop CA). SETAC Press, Pensacola, FL.

Crump ML, Hensley FR, Clark KL (1992) Apparent decline of the golden toad: Underground or extinct? Copeia 1992:413-420.

Cullingam Cl, Kyle CJ, Pond BA, White BN (2008) Genetic structure of raccoons in eastern North America based on mtDNA: implications for subspecies designation and rabies disease dynamics. Canadian Journal of Zoology 86:947-958.

Culver M, Johnson WE, Pecon-Slattery J, O'Brien SJ (2000) Genomic ancestry of the American puma (Puma concolor). Journal of Heredity 91:186-197.

Cushman SA, McKelvey KS, Schwartz MK (2009) Use of empirically derived source­ destination models to map regional conservation corridors. Conservation Biology 23:369-376. Darling JA, Bagley MJ, Roman J, Tepolt CK, Geller JB (2008) Genetic patterns across multiple introductions of the globally invasive crab genus Carcinus. Molecular Ecology 17:4992-5007.

Daugherty CH, Cree A, Hay JM, Thompson MB (1990) Neglected taxonomy and continuing extinctions oftuatara (Sphenodon). Nature 347:177-179.

Debes PV, Zachos FE, Hanel R (2008) Mitochondrial phylogeography of the European sprat (Sprattus sprattus L., Clupeidae) reveals isolated climatically vulnerable populations in the Mediterranean Sea and range expansion in the northeast Atlantic. Molecular Ecology 17:3873-3888.

Delcourt HR, Delcourt PA (1991) Quaternary Ecology: A Paleoecological Perspective. Chapman and Hall, New York.

Delcourt PA, Delcourt HR (1980) Vegetation maps for eastern North America: 40 000 yr BP to the present. Pp. 123-165 in: Geobotany II (ed. Romans RC). Plenum Press, NY. Denton JS, Hitchings SP, Beebee TJC, Gent A (1997) A recovery program for the natterjack toad (Bufo calamita) in Britain. Conservation Biology 11:1329-1338.

Descimon H, Napolitano M (1993) Enzyme polymorphism, wing pattern variability, and geographical isolation in an endangered butterfly species. Biological Conservation 66: 117-123.

Dessauer HC, Nevo E (1969) Geographic variation of blood and liver proteins in cricket frogs. Biochemical Genetics 3: 171-188.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 173 -

Dever JA (2007) Fine-scale genetic structure in the threatened foothill yellow-legged frog (Rana boylii). Journal of Herpetology 41:168-173.

Diana SG, Beasley VR (1998) Amphibian toxicology. Pp. 266-277 in: Status and Conservation of Midwestern Amphibians (ed. Lannoo MJ). University oflowa Press, Iowa City.

Diniz-Filho JAF, Telles MPDC (2002) Spatial autocorrelation analysis and the identification of operational units for conservation in continuous populations. Conservation Biology 16:924-935.

Dodd CK, Jr., Seigel RA (1991) Relocation, repatriation, and translocation of amphibians and reptiles: Are they conservation strategies that work? Herpetologica 47:336-350.

Drew RE, Hallett JG, Aubry KB et al. (2003) Conservation genetics of the fisher (Martes pennanti) based on mitochondrial DNA sequencing. Molecular Ecology 12:51-62.

Dubey S, Shine R (2008) Origin of the parasites of an invading species, the Australian cane toad (Bufo marinus): are the lungworms Australian or American? Molecular Ecology 17:4418-4424.

Duellman WE, Trueb L (1986) Biology of Amphibians. The Johns Hopkins University Press, Baltimore, MD.

Dumeril AMC, Bibron G (1841) Erpetologie generale ou histoire naturelle des reptiles. Libraire Encyclopedique de Roret, Paris.

Dunn ER (1938) Notes on frogs of the genus Acris. Proceedings of the Academy of Natural Sciences of Philadelphia 90: 153-154.

Dupanloup I, Schneider S, Excoffier L (2002) A simulated annealing approach to define the genetic structure of populations. Molecular Ecology 11:2571-2581.

Eagle AC, Hay-Chmielewski EM, Cleveland KT et al. (2005) Michigan's Wildlife Action Plan. Michigan Department of Natural Resources, Lansing, ML

Easteal S ( 1985) The ecological genetics of introduced populations of the giant toad, Bufo marinus. Ill. Geographical patterns of variation. Evolution 39:1065-1075.

Eckert CG, Samis KE, Lougheed SC (2008) Genetic variation across species' geographical ranges: the central-marginal hypothesis and beyond. Molecular Ecology 17:1170-1188.

Eckstein RL, O'Neill RA, Danihelka J, Otte A, Kohler W (2006) Genetic structure among and within peripheral and central populations of three endangered floodplain violets. Molecular Ecology 15:2367-2379.

Eizirik E, Kim J, Menotti-Raymond Met al. (2001) Phylogeography, population history and conservation genetics ofjaguars (Panthera onca, Mammalia, Felidae). Molecular Ecology 10:65-79.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 174 -

Elmer KR, Cannatella DC (2008) Three new species of leaflitter frogs from the upper Amazon forests: cryptic diversity within Pristimantis "ockendeni" (Anura: Strabomantidae) in Ecuador. Zootaxa 11-38.

England PR, Osler GHR, Woodworth LM et al. (2003) Effects of intense versus diffuse population bottlenecks on microsatellite genetic diversity and evolutionary potential. Conservation Genetics 4:595-604.

Estep AD (2000) Geographic variation in wariness, morphology and genetics of the cricket frog, Acris crepitans, in Oklahoma. Unpublished M.Sc. dissertation. University of Central Oklahoma.

Evanno G, Regnaut S, Goudet J (2005) Detecting the number of clusters of individuals using the software STRUCTURE: a simulation study. Molecular Ecology 14:2611-2620.

Evans BJ, Brown RM, McGuire JA et al. (2003) Phylogenetics of fanged frogs: Testing biogeographical hypotheses at the interface of the Asian and Australian faunal zones. Systematic Biology 52:794-819.

Excoffier L, Smouse PE, Quattro JM (1992) Analysis of molecular variance inferred from metric distances among DNA haplotypes: Application to human mitochondrial DNA restriction data. Genetics 131:479-491.

Excoffier L, Laval G, Schneider S (2005) Arlequin ver. 3.0: An integrated software package for population genetics data analysis. Evolutionary Bioinformatics 1:47-50.

Fenneman NM, Johnson DW (1946) Physiographic divisions of the conterminous U.S. USGS, Washington, DC.

Ferguson DE (1963) Mississippi Delta wildlife developing resistance to pesticides. Agricultural Chemicals September:32-34.

Fitzinger L (1843) Systema Reptilium. Society for the Study of Amphibians and Reptiles. 106 pp.

Flickinger EL, King KA, Stout WF, Mohn MM (1980) Wildlife hazards from Furadan 3G applications to rice in Texas. Journal of Wildlife Management 44:190-197.

Fontanella F, Feldman CR, Siddall ME, Burbrink FT (2008) Phylogeography of Diadophis punctatus: Extensive lineage diversity and repeated patterns of historical demography in a trans-continental snake. Molecular Phylogenetics and Evolution 46: 1049-1070.

Ford MJ (2002) Selection in captivity during supportive breeding may reduce fitness in the wild. Conservation Biology 16:815-825.

Forsman A, Ahnesjo J, Caesar S, Karlsson M (2008) A model of ecological and evolutionary consequences of color polymorphism. Ecology 89:34-40.

Frankham R (1995) Effective population size/adult population size ratio in wildlife: a review. Genetical Research 66:95-107.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 175 -

Frankham R, Ballou JD, Briscoe DA (2002) Introduction to Conservation Genetics. Cambridge University Press, Cambridge, United Kingdom.

Fraser DJ, Bernatchez L (2001) Adaptive evolutionary conservation: towards a unified concept for defining conservation units. Molecular Ecology 10:2741-2752.

Friar EA, Ladoux T, Roalson HE, Robichaux HR (2000) Microsatellite analysis of a population crash and bottleneck in the silversword, sandwicense ssp. sandwicense (), and its implications for reintroduction. Molecular Ecology 9:2027-2034.

Frost DR, Grant T, Faivovich Jet al. (2006) The amphibian tree oflife. Bulletin of the American Museum of Natural History 297:1-370.

Fu YX ( 1997) Statistical tests of neutrality of mutations against population growth, hitchhiking and background selection. Genetics 147:915-925.

Funk WC, Blouin MS, Corn PS et al. (2005) Population structure of Columbia spotted frogs (Rana luteiventris) is strongly affected by the landscape. Molecular Ecology 14:483-496.

Galbreath PF, Adams ND, Guffey SZ, Moore CJ, West JL (2001) Persistence of native southern Appalachian brook trout populations in the Pigeon River system, North Carolina. North American Journal of Fisheries Management 21:927-934. Galeotti P, Rubolini D, Dunn PO, Fasola M (2003) Colour polymorphism in birds: causes and functions. Journal of Evolutionary Biology 16:635-646.

Gamble T, Berendzen PB, Shaffer HB, Starkey DE, Simons AM (2008) Species limits and phylogeography of North American cricket frogs (Acris: Hylidae). Molecular Phylogenetics and Evolution 48:112-125.

Garcia-Diaz J (1967) Rediscovery of Buja lemur (Cope) and additional records ofreptiles from Puerto Rico. Stahlia 10: 1-6.

Garcia-Moreno J, Matocq MD, Roy MS, Geffen E, Wayne RK (1996) Relationships and genetic purity of the endangered Mexican wolf based on analysis of microsatellite loci. Conservation Biology 10:376-389.

Garcia-Ramos G, Kirkpatrick M (1997) Genetic models of adaptation and gene flow in peripheral populations. Evolution 51 :21-28.

Germano JM, Bishop P J (2009) Suitability of amphibians and reptiles for translocation. Conservation Biology 23:7-15.

Givnish TJ, Millam KC, Mast AR et al. (2009) Origin, and diversification of the Hawaiian lobeliads (: Campanulaceae). Proceedings of the Royal Society B 276:407-416.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 176 -

Godoy JA, Negro JJ, Hiraldo F, Donazar JA (2004) Phylogeography, genetic structure and diversity in the endangered bearded vulture (Gypaetus barbatus, L.) as revealed by mitochondrial DNA. Molecular Ecology 13:371-390.

Goebel AM (1996) Systematics and conservation ofbufonids in North America and in the Bufo boreas species group. Unpublished Ph.D. dissertation. University of Colorado.

Goebel AM, Donnelly JM, Atz ME ( 1999) PCR primers and amplification methods for l 2S ribosomal DNA, the control region, cytochrome oxidase I, and cytochrome b in bufonids and other frogs, and an overview of PCR primers which have amplified DNA in amphibians successfully. Molecular Phylogenetics and Evolution 11:163-199.

Good JM, Demboski JR, Nagorsen DW, Sullivan J (2003) Phylogeography and introgressive hybridization: chipmunks (genus Tamias) in the northern Rocky Mountains. Evolution 57:1900-1916.

Goodman SJ, Tamate HB, Wilson R et al. (2001) Bottlenecks, drift and differentiation: The population structure and demographic history of sika deer (Cervus nippon) in the Japanese archipelago. Molecular Ecology 10:1357-1370.

Gorman WL ( 1986) Patterns of color polymorphism in the cricket frog, A eris crepitans, in Kansas. Copeia 1986:995-999.

Gorman WL, Gaines MS (1987) Patterns of genetic variation in the cricket frog, Acris crepitans, in Kansas. Copeia 1987:352-360.

Gasner KL, Black IH (1957) The effects of acidity on the development and hatching of New Jersey frogs. Ecology 38:256-262.

Graham A (1999) Late Cretaceous and Cenozoic History of North American Vegetation North of Mexico. Oxford University Press, Oxford. 350 pp.

Grant C (1931) Notes on Bufo marinus (Linnaeus). Copeia 1931:62.

Gray RH (1971) Fall activity and overwintering of the cricket frog (Acris crepitans) in central Illinois. Copeia 1971:748-750.

Gray RH (1972) Metachrosis of the vertebral stripe in the cricket frog, Acris crepitans. The American Midland Naturalist 87:549-551.

Gray RH (1983) Seasonal, annual and geographic variation in color morph frequencies of the cricket frog, Acris crepitans, in Illinois. Copeia 1983:300-311.

Gray RH ( 1984) Effective breeding size and the adaptive significance of color polymorphism in the cricket frog (Acris crepitans) in Illinois, USA. Amphibia-Reptilia 5:101-107.

Gray RH (1995) An unusual color pattern variant in cricket frogs (Acris crepitans) from southern Illinois. Transactions of the Illinois State Academy of Science 88: 13 7-13 8.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 177 -

Gray RH (2000) Historical occurrence of malformations in the cricket frog, Acris crepitans, in Illinois. Transactions of the Illinois State Academy of Science 93:279- 284.

Gray RH, Brown LE (2005) Decline of northern cricket frogs (Acris crepitans). Pp. 47-54 in: Amphibian Declines: The Conservation Status of United States Species (ed. Lannoo MJ). University of California Press, Berkeley, CA.

Gray RH, Brown LE, Blackbum L (2005) A eris crepitans Baird, 1854. Pp. 441-443 in: Amphibian Declines: The Conservation Status of United States Species (ed. Lannoo MJ). University of California Press, Berkeley, CA.

Graybeal A ( 1997) Phylogenetic relationships of bufonid frogs and tests of alternate macroevolutionary hypotheses characterizing their radiation. Zoological Journal of the Linnean Society 119:297-338.

Green DM (2003) The ecology of extinction: population fluctuation and decline in amphibians. Biological Conservation 111:331-343. Greenwell M, Beasley VR, Brown LE (1996) The mysterious decline of the cricket frog. Aquaticus 26:48-54.

Guiher TJ, Burbrink FT (2008) Demographic and phylogeographic histories of two venomous North American snakes of the genus Agkistrodon. Molecular Phylogenetics and Evolution 48:543-553. Guo SW, Thompson EA (1992) Performing the exact test of Hardy-Weinberg proportion for multiple alleles. Biometrics 48:361-372.

Gunther A (1858) Catalogue of the Batrachia Salientia in the Collection of the British Museum. Francis and Collins, London.

Hakansson J, Bratt C, Jensen P (2007) Behavioural differences between two captive populations of red jungle fowl (Gallus gallus) with different genetic background, raised under identical conditions. Applied Animal Behaviour Science 102:24-38.

Hall TA (1999) BIOEDIT: a user-friendly biological sequence alignment editor and analysis program for Windows 95-98-NT. Nucleic Acids Symposium Series 41:95-98.

Halverson MA, Skelly DK, Caccone A (2006) Inbreeding linked to amphibian survival in the wild but not in the laboratory. Journal of Heredity 97:499-507.

Hamilton MB, Pincus EL, Di Fiore A, Fleischer RC (1999) Universal linker and ligation procedures for construction of genomic DNA libraries enriched for microsatellites. Biotechniques 27:500-507. Hammerson GA, Livo LJ (1999) Conservation status of the northern cricket frog (Acris crepitans) in Colorado and adjacent areas at the northwestern extent of the range. Herpetological Review 30:78-80.

Hanski I (1998) Metapopulation dynamics. Nature 396:41-49.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 178 -

Harper F (1947) A new cricket frog (Acris) from the middle western states. Proceedings of the Biological Society of Washington 60:39-40.

Hass CA, Maxson LR, Hedges SB (2001) Relationships and divergence times of West Indian amphibians and reptiles: insights from albumin immunology. Pp. 157-174 in: Biogeography of the West Indies: Patterns and Perspectives (eds. Woods CA, Sergile FE). CRC Press, Boca Raton, FL.

Hay RM (1998) Blanchard's cricket frogs in Wisconsin: a status report. Pp. 79-82 in: Status and Conservation of Midwestern Amphibians ( ed. Lannoo MJ). University of Iowa Press, Iowa City. Hebert PDN, Cywinska A, Ball SL, deWaard JR (2003) Biological identifications through DNA barcodes. Proceedings of the Royal Society of London B 270:313-321. Hedges SB, Hass CA, Maxson LR (1992) Caribbean biogeography: Molecular evidence for dispersal in West Indian terrestrial vertebrates. Proceedings of the National Academy of Sciences 89:1909-1913.

Hedges SB (1996) The origin of West Indian amphibians and reptiles. Pp. 95-128 in: Contributions to West Indian Herpetology: A Tribute to Albert Schwartz (eds. Powell R, Henderson RW). Society for the Study of Amphibians and Reptiles, Ithaca, NY.

Hedges SB, Joglar RL, Thomas R (2004) Bufo lemur. 2006 IUCN Red List of Threatened Species, IUCN 2006. www.iucnredlist.org. Accessed 6-11-2008. Hedrick PW (1995) Gene flow and genetic restoration: The Florida panther as a case study. Conservation Biology 9:996-1007.

Hedrick PW, Kalinowski ST (2000) Inbreeding depression in conservation biology. Annual Review of Ecology and Systematics 31: 139-162.

Hedrick PW (2001) Conservation genetics: where are we now? Trends in Ecology and Evolution 16:629-636.

Hemesath LM (1998) Iowa's frog and toad survey, 1991-1994. Pp. 206-216 in: Status and Conservation of Midwestern Amphibians (ed. Lannoo MJ). University of Iowa Press, Iowa City. Hensen I, Wesche K (2006) Relationships between population size, genetic diversity and fitness components in the rare Dictamnus albus in central Germany. Biodiversity and Conservation 15:2249-2261.

Hewitt GM (1996) Some genetic consequences of ice ages, and their role in divergence and speciation. Biological Journal of the Linnean Society 58:247-276.

Hewitt GM (2000) The genetic legacy of the Quaternary ice ages. Nature 405:907-913.

Hewitt GM, Ibrahim KM (2001) Inferring glacial refugia and historical migrations with molecular phylogenies. Pp. 271-294 in: Integrating Ecology and Evolution in a Spatial Context (eds. Silvertown J, Antonovics J). Cambridge University Press, Cambridge.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 179 -

Hewitt GM (2004) Genetic consequences of climatic oscillations in the Quaternary. Philosophical Transactions of the Royal Society of London B 359: 183-195.

Hoelzel AR, Halley J, Campagna C et al. ( 1993) Elephant seal genetic differentiation and the use of simulation models to investigate historical population bottlenecks. Journal of Heredity 84:443-449.

Hoffman EA, Blouin MS (2004) Evolutionary history of the northern leopard frog: Reconstruction of phylogeny, phylogeography, and historical changes in population demography from mitochondrial DNA. Evolution 58: 145-159.

Hoffman EA, Schueler FW, Blouin MS (2004) Effective population sizes and temporal stability of genetic structure in Rana pipiens, the northern leopard frog. Evolution 58:2536-2545.

Holderegger R, Kamm U, Gugerli F (2006) Adaptive vs. neutral genetic diversity: implications for landscape genetics. Landscape Ecology 21:797-807.

Holman JA (1995) Pleistocene Amphibians and Reptiles in North America, Oxford University Press, Oxford.

Houlahan JE, Findlay CS, Schmidt BR, Meyer AH, Kuzmin SL (2000) Quantitative evidence for global amphibian population declines. Nature 404:752-755. Howes BJ, Lindsay B, Lougheed SC (2006) Range-wide phylogeography of a temperate lizard, the five-lined skink (Eumeces fasciatus). Molecular Phylogenetics and Evolution 40: 183-194.

Hunter MLJ, Hutchinson A (1994) The virtues and shortcomings of parochialism: Conserving species that are locally rare, but globally common. Conservation Biology 8:1163-1165.

Huson DH, Bryant D (2005) Application of phylogenetic networks in evolutionary studies. Molecular Biology and Evolution 23:254-267.

Hutchison DW (2003) Testing the central/peripheral model: Analyses of microsatellite variability in the collared lizard (Crotaphytus collaris collaris). American Midland Naturalist 149:148-162.

Ibrahim KM, Nichols RA, Hewitt GM (1996) Spatial patterns of genetic variation generated by different forms of dispersal during range expansion. Heredity 77:282- 291.

Irwin DE (2002) Phylogeographic breaks without geographic barriers to gene flow. Evolution 56:2383-2394. Irwin JT, Costanzo JP, Lee RE (1999) Terrestrial hibernation in the northern cricket frog, A eris crepitans. Canadian Journal of Zoology 77: 1240-1246.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 180 -

Irwin JT (2005) Overwintering in northern cricket frogs (Acris crepitans). Pp. 55-58 in: Amphibian Declines: The Conservation Status of United States Species (ed. Lannoo MJ). University of California Press, Berkeley, CA.

James CH, Moritz C (2000) Intraspecific phylogeography in the sedge frog Litoria fallax (Hylidae) indicates pre-Pleistocene vicariance of an open forest species from eastern Australia. Molecular Ecology 9:349-358.

Jancovich JK, Davidson EW, Parameswaran N et al. (2005) Evidence for emergence of an amphibian iridoviral disease because of human-enhanced spread. Molecular Ecology 14:213-224.

Jehle R, Arntzen JW (2002) Microsatellite markers in amphibian conservation genetics. Herpetological Journal 12:1-9.

Jensen JB (2005) Acris gryllus (LeConte, 1825). p.443 in: Amphibian Declines: The Conservation Status of United States Species (ed. Lannoo MJ). University of California Press, Berkeley, CA.

Jirku M, Bolek MG, Whipps CM et al. (2006) A new species of Myxidium (Myxosporea: Myxidiidae ), from the western chorus frog, Pseudacris triseriata triseriata, and Blanchard's cricket frog, Acris crepitans blanchardi (Hylidae), from eastern Nebraska: morphology, phylogeny, and critical comments on amphibian Myxidium taxonomy. Journal of Parasitology 92:611-619.

Johansson M, Primmer CR, Sahlsten J, Merila J (2005) The influence of landscape structure on occurrence, abundance and genetic diversity of the common frog, Rana temporaria. Global Change Biology 11:1664-1679.

Johnson B (1999) Recovery of the Puerto Rican Crested Toad. Endangered Species Bulletin 24:8-9.

Johnson BK, Christiansen JL (1976) The food and food habits of Blanchard's cricket frog, Acris crepitans blanchardi (Amphibia, Anura, Hylidae ), in Iowa. Journal of Herpetology 10:63-74.

Johnson RR (1990) Release and translocation strategies for the Puerto Rican crested toad, Peltophryne lemur. Endangered Species Update 8:54-57.

Kalendar R (2005) F ASTPCR: A PCR primer design and repeat sequence searching software with additional tools for the manipulation and analysis of DNA and proteins. Helsinki, Finland.

Kalinowski ST (2004) Counting alleles with rarefaction: Private alleles and hierarchical design. Conservation Genetics 5:539-543.

Kalinowski ST (2005) HP-RARE 1.0: a computer program for performing rarefaction on measures of allelic richness. Molecular Ecology Notes 5: 187-189.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 181 -

Keane TM, Creevey CJ, Pentony MM, Naughton TJ, Mclnerney JO (2006) Assessment of methods for amino acid matrix selection and their use on empirical data shows that ad hoc assumptions for choice of matrix are not justified. BMC Evolutionary Biology 6:29.

Kellar TC, Waldron G, Bishop C et al. (1997) National Recovery Plan for Blanchard's Cricket Frog. Recovery of Nationally Endangered Wildlife Committee, Ottawa.

Kiesecker JM, Blaustein AR, Belden LK (2001) Complex causes of amphibian population declines. Nature 410:681-684.

Knopp T, Cano JM, Crochet P-A, Merila J (2007) Contrasting levels of variation in neutral and quantitative genetic loci on island populations of moor frogs (Rana arvalis). Conservation Genetics 8:45-56.

Knopp T, Merila J (2008) The postglacial recolonization of Northern Europe by Rana arvalis as revealed by microsatellite and mitochondrial DNA analyses. Heredity 102: 174-181.

Knutson MG, Sauer JR, Olsen DA et al. (2000) Landscape associations of frog and toad species in Iowa and Wisconsin, U.S.A. Journal of the Iowa Academy of Science 107: 134-145. Koscinski D, Handford P, Tubaro PL, Sharp S, Lougheed SC (2008) Pleistocene climatic cycling and diversification of the Andean treefrog, Hypsiboas andinus. Molecular Ecology 17:2012-2025. Kozak KH, Larson A, Bonett RM, Harmon LJ (2005) Phylogenetic analysis of ecomorphological divergence, community structure, and diversification rates in dusky salamanders (Plethodontidae: Desmognathus). Evolution 59:2000-2016.

Kraaijeveld-Smit FJL, Beebee TJC, Griffiths RA, Moore RD, Schley L (2005) Low gene flow but high genetic diversity in the threatened Mallorcan midwife toad Alytes muletensis. Molecular Ecology 14:3307-3315.

Kumar S,Tamura K, Nei M (2004) MEGA3: Integrated software for Molecular Evolutionary Genetics Analysis and sequence alignment. Briefings in Bioinformatics 5:150-163.

Kyle CJ, Wilson CC (2007) Mitochondrial DNA identification of game and harvested freshwater fish species. Forensic Science International 166:68-76.

Lacy R, Foster M (1987) Summary of electrophoretic analyses for Puerto Rican crested toad, Chicago Zoological Society.

Lada H, Mac Nally R, Taylor AC (2008) Phenotype and gene flow in a marsupial (Antechinus jlavipes) in contrasting habitats. Biological Journal of the Linnean Society 94:303-314.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 182 -

Lamb T, Avise JC (1987) Morphological variability in genetically defined categories of anuran hybrids. Evolution 41:157-165.

Land D, Shindle D, Cunningham M, Lotz M, Ferree B (2004) Florida panther genetic restoration and management annual report 2003-04. Florida Fish and Wildlife Conservation Commission, Tallahassee, FL.

Lannoo MJ, Lang K, Waltz T, Phillips GS (1994) An altered amphibian assemblage: Dickinson County, Iowa, 70 years after Frank Blanchard's survey. American Midland Naturalist 131:311-319.

Lannoo MJ (1998a) Conclusion. Pp. 429-432 in: Status and Conservation of Midwestern Amphibians ( ed. Lannoo MJ). University of Iowa Press, Iowa City.

Lannoo MJ (1998b) Amphibian conservation and wetland management in the upper Midwest: a catch-22 for the cricket frog? Pp. 330-339 in: Status and Conservation of Midwestern Amphibians (ed. Lannoo MJ). University oflowa Press, Iowa City.

Larson S, Jameson R, Etnier M, Fleming M, Bentzen P (2002) Loss of genetic diversity in sea otters (Enhydra lutris) associated with the fur trade of the 18th and 19th centuries. Molecular Ecology 11: 1899-1903.

LeConte J ( 1825) Remarks on the American species of the Genera Hyla and Rana. Annals of the Lyceum of Natural History of New York 1:278-282.

Lee-Yaw JA, Irwin JT, Green DM (2008) Postglacial range expansion from northern refugia by the wood frog, Rana sylvatica. Molecular Ecology 17:867-884.

Lehtinen RM (2002) A historical study of the distribution of Blanchard's cricket frog (Acris crepitans blanchardi) in southeastern Michigan. Herpetological Review 33:194-197.

Lehtinen RM, Skinner AA (2006) The enigmatic decline of Blanchard's cricket frog (A eris crepitans blanchardi): A test of the habitat acidification hypothesis. Copeia 2006:159-167.

Leja WT (1998) Aquatic Habitats in the Midwest: Waiting for Amphibian Conservation Initiatives. Pp. 345-353 in: Status and Conservation of Midwestern Amphibians (ed. Lannoo MJ). University of Iowa Press, Iowa City.

Lemmon EM, Lemmon AR, Cannatella DC (2007a) Geological and climatic forces driving speciation in the continentally distributed trilling chorus frogs (Pseudacris). Evolution 61:2086-2103.

Lemmon EM, Lemmon AR, Collins JT, Lee-Yaw JA, Cannatella DC (2007b) Phylogeny-based delimitation of species boundaries and contact zones in the trilling chorus frogs (Pseudacris). Molecular Phy lo genetics and Evolution 44: 1068-1082.

Lentini A (2000) Puerto Rican Crested Toad (Peltophryne lemur) SSP Husbandry Manual, Keeper and Curator Edition. Toronto Zoo, Scarborough, ON.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 183 -

Lesbarreres D, Primmer CR, Laurila A, Merila J (2005) Environmental and population dependency of genetic variability-fitness correlations in Rana temporaria. Molecular Ecology 14:311-323. Lesica P, Allendorf FW ( 1995) When are peripheral populations valuable for conservation? Conservation Biology 9:753-760.

Lessa EP, Cook JA, Patton JL (2003) Genetic footprints of demographic expansion in North America, but not Amazonia, during the Late Quaternary. Proceedings of the National Academy of Sciences 100:10331-10334.

Lips KR, Brem F, Brenes R et al. (2006) Emerging infectious disease and the loss of biodiversity in a Neotropical amphibian community. Proceedings of the National Academy of Sciences 103:3165-3170. Littlejohn MJ, Roberts JD (1975) Acoustic analysis of an intergrade zones between two call races of the Limnodynastes tasmaniensis complex (Anura: Leptodactylidae) in south-eastern Australia. Australian Journal of Zoology 23: 113-122.

Lo mo lino MV, Channell R ( 1995) Splendid isolation: patterns of geographic range collapse in endangered mammals. Journal ofMammalogy 76:335-347. Lougheed SC, Gascon C, Jones DA, Bogart JP, Boag PT (1999) Ridges and rivers: a test of competing hypotheses of Amazonian diversification using a dart-poison frog (Epipedobatesfemoralis). Proceedings of the Royal Society of London B 266:1829- 1835.

Lundquist CJ, Granek EF (2005) Strategies for successful marine conservation: integrating socioeconomic, political, and scientific factors. Conservation Biology 19: 1771-1778.

Lynch M, Pfrender M, Spitze Ket al. (1999) The quantitative and molecular genetic architecture of a subdivided species. Evolution 53: 100-110. Mace GM, Smith TB, Bruford MW, Wayne RK (1996) An Overview of the Issues. Pp. 1- 21 in: Molecular Genetic Approaches in Conservation (eds. Smith TB, Wayne RK). Oxford University Press, New York.

Madsen T, Shine R, Olsson M, Wittzell H (1999) Restoration of an inbred adder population. Nature 402:34-35.

Maglia AM, Pugener LA, Mueller JM (2007) Skeletal morphology and postmetamorphic ontogeny of A eris crepitans (Anura: Hylidae ): a case of miniaturization in frogs. Journal ofMorphofogy 268:194-223.

Manel S, Schwartz MK, Luikart G, Taberlet P (2003) Landscape genetics: combining landscape ecology and population genetics. Trends in Ecology and Evolution 18: 189- 197.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 184 -

Manel S, Gaggiotti OE, Waples RS (2005) Assignment methods: matching biological questions with appropriate techniques. Trends in Ecology and Evolution 20: 136-142.

Manni F, Guerard E, Heyer E (2004) Geographic patterns of (genetic, morphologic, linguistic) variation: How barriers can be detected by using Monmonier's algorithm. Human Biology 76:173-190.

Marsh DM, Trenham PC (2001) Metapopulation dynamics and amphibian conservation. Conservation Biology 15:40-49. Martinez-Solano I, Rey I, Garcia-Paris M (2005) The impact of historical and recent factors on genetic variability in a mountain frog: the case of Rana iberica (Anura: Ranidae). Animal Conservation 8:431-441.

Maruska EJ (1986) Amphibians: review of zoo breeding programmes. International Zoo Yearbook 24/25:56-65.

Masta SE, Laurent NM, Routman EJ (2003) Population genetic structure of the toad Bufo woodhousii: an empirical assessment of the effects of haplotype extinction on nested cladistic analysis. Molecular Ecology 12:1541-1554.

Matos-Torres JJ (2006) Habitat characterization for the Puerto Rican crested toad (Peltophryne lemur) at Guanica State Forest, Puerto Rico. Unpublished M.Sc. dissertation. University of Puerto Rico - Mayaguez. Mayr E (1963) Animal Species and Evolution. Harvard University Press, Cambridge.

McCallum ML (2007) Amphibian decline or extinction? Current declines dwarf background extinction rate. Journal of Herpetology 41:486-491.

McCallum ML, Trauth SE (2003) A forty-three year museum study of northern cricket frog (Acris crepitans) abnormalities in Arkansas: Upward trends and distributions. Journal of Wildlife Diseases 39:522-528. McCallum ML, Trauth SE (2004) Blanchard's cricket frog in Nebraska and South Dakota. Prairie Naturalist 36: 129-135.

McCallum ML, Trauth SE (2006) An evaluation of the subspecies Acris crepitans blanchardi (Anura, Hylidae). Zootaxa 1104:1-21.

McCallum ML, Trauth SE (2007) Physiological trade-offs between immunity and reproduction in the northern cricket frog (Acris crepitans). Herpetologica 63:269-274.

McCarthy MA, Possingham HP (2007) Active adaptive management for conservation. Conservation Biology 21:956-953.

McClelland BE, Wilczynski W, Ryan MJ (1998) Intraspecific variation in laryngeal and ear morphology in male cricket frogs (Acris crepitans). Biological Journal of the Linnean Society 63:51-67.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 185 -

McKay JK, Latta RG (2002) Adaptive population divergence: markers, QTL and traits. Trends in Ecology and Evolution 17:285-291.

McLeod DS (2005) Nebraska's declining amphibians. Pp. 292-294 in: Amphibian Declines: The Conservation Status of United States Species (ed. Lannoo MJ). University of California Press, Berkeley, CA.

Mecham JS ( 1964) Ecological and genetic relationships of the two cricket frogs, genus Acris, in Alabama. Herpetologica 20:80-91.

Mehlman DW (1997). Change in avian abundance across the geographic range in response to environmental change. Ecological Applications 7:614-624.

Meyer E, Newell D, Hines H et al. (2004) Rheobatrachus situs. 2006 IUCN Red List of Threatened Species, IUCN 2006. www.iucnredlist.org. Accessed 14-04-2007.

Micancin JP (2008) Acoustic variation and species discrimination in southeastern sibling species, the cricket frogs Acris crepitans and Acris gryllus. Unpublished Ph.D. dissertation. University of North Carolina-Chapel Hill.

Mickelson DM, Colgan PM (2004) The southern Laurentide Ice Sheet. Pp. 1-16 in: The Quaternary Period in the United States (eds. Gillespie AR, Porter SC, Atwater BF). Elsevier, Amsterdam.

Mierzwa KS ( 1998) Status of northeastern Illinois amphibians. Pp. 115-124 in: Status and Conservation of Midwestern Amphibians ( ed. Lannoo MJ). University of Iowa Press, Iowa City.

Miller DA, White RA (1998) A conterminous United States multi-layer soil characteristics data set for regional climate and hydrology modeling. Earth Interactions 2:1-26.

Miller TJ (1985) Husbandry and breeding of the Puerto Rican toad (Peltophryne lemur) with comments on its natural history. Zoo Biology 4:281-286.

Milstead WW, Rand AS, Steward MM (1974) Polymorphism in cricket frogs: An hypothesis. Evolution 28:489-491.

Minton SA (1998) Observations on Indiana amphibian populations: a forty-five-year overview. Pp. 217-220 in: Status and Conservation of Midwestern Amphibians (ed. Lannoo MJ). University of Iowa Press, Iowa City.

Mittleman MB (1947) Miscellaneous notes on Indiana amphibians and reptiles. American Midland Naturalist 38:466-484.

Monroe WH (1976) The karst landforms of Puerto Rico, 899 ed. USGS Professional Paper. US Government Printing Office, Washington, DC.

Monsen KJ, Blouin MS (2003) Genetic structure in a montane ranid frog: restricted gene flow and nuclear-mitochondrial discordance. Molecular Ecology 12:3275-3286.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 186 -

Montgomery ME, Ballou JD, Nurthen RK et al. (1997) Minimizing kinship in captive breeding programs. Zoo Biology 16:377-389.

Moore DS (1997) Molecular Analyses of Temperate Hylid Frogs using the Mitochondrial Cytochrome b Gene. Unpublished Ph.D. dissertation. University of Alabama at Birmingham.

Moreno JA. Unpublished notes on Peltophryne lemur Cope. 1985. Internal document for the Puerto Rico Department of Natural Resources.

Moriarty JJ (1998) Status of amphibians in Minnesota. Pp. 166-168 in: Status and Conservation of Midwestern Amphibians ( ed. Lannoo MJ). University of Iowa Press, Iowa City.

Moritz C, Dowling TE, Brown WM (1987) Evolution of animal mitochondrial DNA: relevance for population biology and systematics. Annual Review of Ecology and Systematics 18:269-292.

Moritz C (1999) Conservation units and translocations: strategies for conserving evolutionary processes. Hereditas 130:217-228.

Morris AB, Ickert-Bond SM, Brunson DB, Soltis DE, Soltis PS (2008) Phylogeographical structure and temporal complexity in American sweetgum (Liquidambar styraciflua; Altingiaceae). Molecular Ecology 17:3889-3900.

Mossman MJ, Hartman LM, Hay RM, Sauer JR, Dhuey BJ (1998) Monitoring long-term trends in Wisconsin frog and toad populations. Pp. 169-198 in: Status and Conservation of Midwestern Amphibians ( ed. Lannoo MJ). University of Iowa Press, Iowa City.

Mount RH (1975) The Reptiles and Amphibians of Alabama, Auburn University Agricultural Experiment Station, Auburn, AL.

Murphy RW, Sites JW, Buth DG, Haufler CH (1996) Proteins: Isozyme Electrophoresis. Pp. 51-120 in: Molecular Systematics (eds. Hillis DM, Moritz C, Mable BK). Sinauer Associates Inc., Sunderland, MA. ·

Myers N, Mittermeier RA, Mittermeier CG, da Fonseca GAB, Kent J (2000) Biodiversity hotspots for conservation priorities. Nature 403:853-858.

Nei M (1972) Genetic distance between populations. American Naturalist 106:283-292.

Neill WT (1950) Taxonomy, nomenclature, and distribution of southeastern cricket frogs, genus Acris. American Midland Naturalist 43:152-156.

Netting MG, Goin CJ (1945) The cricket-frog of peninsular Florida. Quarterly Journal of the Florida Academy of Sciences 8:304-310.

Nevo E (1973a) Adaptive variation in size of cricket frogs. Ecology 54:1271-1281.

Nevo E (1973b) Adaptive color polymorphism in cricket frogs. Evolution 27:353-367.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 187 -

Nevo E, Capranica RR (1985) Evolutionary origin of ethological reproductive isolation in cricket frogs, Acris. Pp. 147-214 in: Evolutionary Biology, vol. 19 (eds. Hecht MK, Wallace B, Prance GT). Plenum Press, New York.

NYSDEC (2007) Checklist of Amphibians, Reptiles, Birds and Mammals of New York State. New York State Department of Environmental Conservation, Division of Fish, Wildlife and Marine Resources, Wildlife Diversity Group.

Newman D, Tallmon DA (2001) Experimental evidence for beneficial fitness effects of gene flow in recently isolated populations. Conservation Biology 15: 1054-1063.

Newman RA, Squire T (2001) Microsatellite variation and fine-scale population structure in the wood frog (Rana sylvatica). Molecular Ecology 10:1087-1100.

Nielson M, Lohman K, Sullivan J (2001) Phylogeography of the tailed frog (Ascaphus truei): Implications for the biogeography of the Pacific Northwest. Evolution 55:147- 160.

Nittinger F, Gamauf A, Pinsker W, Wink M, Haring E (2007) Phylogeography and population structure of the saker falcon and the influence of hybridization: mitochondrial and microsatellite data. Molecular Ecology 16:1497-1517.

Noonan BP, Chippindale PT (2006) Dispersal and vicariance: The complex evolutionary history of boid snakes. Molecular Phylogenetics and Evolution 40:347-358. Oden NL (1984) Assessing the significance of a spatial correlogram. Geographical Analysis 16: 1-16.

Oldham MJ (1992) Declines in Blanchard's cricket frog in Ontario. Pp. 30-31 in: Declines in Canadian amphibian populations: designing a national monitoring strategy ( eds. Bishop CA, Pettit KE). Canadian Wildlife Service, Occasional Paper No. 76.

Ouellet M (2000) Amphibian Deformities: Current State of Knowledge. Pp. 617-661 in: Ecotoxicology of Amphibians and Reptiles (eds. Sparling DW, Linder G, Bishop CA). SETAC Press, Pensacola, FL.

Padial JM, de la Riva I (2009) Integrative taxonomy reveals cryptic Amazonian species of Pristimantis (Anura: Strabomantidae). Zoological Journal of the Linnean Society 155:97-122.

Paetkau D, Waits LP, Clarkson PL, Craighead L, Strobeck C (1997) An empirical evaluation of genetic distance statistics using microsatellite data from bear (Ursidae) populations. Genetics 14 7: 194 3-19 5 7.

Palo JU, O'Hara RB, Laugen AT et al. (2003) Latitudinal divergence of common frog (Rana temporaria) life history traits by natural selection: evidence from a comparison of molecular and quantitative genetic data. Molecular Ecology 12: 1963-1978.

Palo JU, Schmeller DS, Laurila A et al. (2004) High degree of population subdivision in a widespread amphibian. Molecular Ecology 13:2631-2644.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 188 -

Parent CE, Caccone A, Petren K (2008) Colonization and diversification of Galapagos terrestrial fauna: a phylogenetic and biogeographical synthesis. Philosophical Transactions of the Royal Society B 363:3347-3361.

Patenaude NJ, Quinn JS, Beland P, Kingsley M, White BN (1994) Genetic variation of the St. Lawrence beluga whale population assessed by DNA fingerprinting. Molecular Ecology 3:375-381.

Pauly GB, Hillis DM, Cannatella DC (2004) The history of a nearctic colonization: Molecular phylogenetics and biogeography of the nearctic toads (Bufo ). Evolution 58:2514-2535.

Pauly GB, Piskurek 0, Shaffer HB (2007) Phylogeographic concordance in the southeastern United States: the flatwoods salamander, Ambystoma cingulatum, as a test case. Molecular Ecology 16:415-429.

Peakall R, Smouse PE (2006) GENALEX 6: Genetic analysis in Excel. Population genetic software for teaching and research. Molecular Ecology Notes 6:288-295.

Pechmann JHK, Scott DE, Semlitsch RD et al. (1991) Declining amphibian populations: The problem of separating human impacts from natural fluctuations. Science 253:892- 895.

Pechmann JHK, Wilbur HM (1994) Putting declining amphibian populations in perspective: natural fluctuations and human impacts. Herpetologica 50:65-84.

Petersen SD, Mason T, Akber Set al. (2007) Species identification of tarantulas using exuviae for international wildlife law enforcement. Conservation Genetics 8:497-502.

Phillips K (1990) Where Have All the Frogs and Toads Gone? Bioscience 40:422.

Pico R (1974) The Geography of Puerto Rico. Aldine Publishing Co., Chicago. 439 pp.

Posada D, Crandall KA (1998) MODEL TEST: testing the model of DNA substitution. Bioinformatics 14:817-818.

Posada D, Crandall KA (2001) Intraspecific gene genealogies: trees grafting into networks. Trends in Ecology and Evolution 16: 37-45.

Pounds JA, Crump ML (1994) Amphibian declines and climate disturbance: the case of the golden toad and the harlequin frog. Conservation Biology 8:72-85.

Pounds A, Savage J (2004) Bufo periglenes. 2006 IUCN Red List of Threatened Species, IUCN 2006. www.iucnredlist.org. Accessed 14-04-2007.

Pounds JA, Bustamante MR, Coloma LA et al. (2006) Widespread amphibian extinctions from epidemic disease driven by global warming. Nature 439:161-167.

Pounds JA, Fogden MPL, Savage JM, Gorman GC (1997) Tests of null models for amphibian declines on a tropical mountain. Conservation Biology 11:1307-1322.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 189 -

Pounds JA, Fogden MPL, Campbell JH (1999) Biological response to climate change on a tropical mountain. Nature 398:611-615.

Pramuk JB, Hass CA, Hedges SB (2001) Molecular phylogeny and biogeography of West Indian toads (Anura: Bufonidae). Molecular Phylogenetics and Evolution 20:294-301.

Pramuk JB (2002) Combined evidence and cladistic relationships of West Indian toads (Anura:Bufonidae). Herpetological Monographs 16:121-151.

Pray LA, Goodnight CJ (1995) Genetic variation in inbreeding depression in the red flour beetle Tribolium castaneum. Evolution 49: 176-188.

PRCT SSP (2006) The Puerto Rican Crested Toad Species Survival Plan Website. www.crestedtoadssp.org. Accessed 04-06-2008.

Pregill G (1981a) Cranial morphology and the evolution of West Indian toads (Salientia: Bufonidae): Resurrection of the genus Peltophryne Fitzinger. Copeia 1981:273-285.

Pregill G ( 1981 b) Late Pleistocene herpetofaunas from Puerto Rico. University of Kansas Miscellaneous Publications No. 71. University of Kansas, Lawrence, KS.

Primmer CR, Merila J (2002) A low rate of cross-species microsatellite amplification success in Ranid frogs. Conservation Genetics 3:445-449. Pritchard JK, Stephens M, Donnelly P (2000) Inference of population structure using multilocus genotype data. Genetics 155:945-959.

Pyburn WF (1958) Size and movements of a local population of cricket frogs (A eris crepitans). Texas Journal of Science 10:325-342.

Pyburn WF (1961) The inheritance and distribution of vertebral stripe color in the cricket frog. Pp. 235-261 in: Vertebrate Speciation (ed. Blair WF). Univ. of Texas, Austin.

Rahel FJ (2007) Biogeographic barriers, connectivity and homogenization of freshwater faunas: it's a small world after all. Freshwater Biology 52:696-710.

Ralin DB, Rogers JS (1972) Aspects of tolerance to desiccation inAcris crepitans and Pseudacris streckeri. Copeia 1972:519-525.

Ramos-Onsins SE, Rozas J (2002) Statistical properties of new neutrality tests against population growth. Molecular Biology and Evolution 19:2092-2100.

Raymond M, Rousset F (1995) GENEPOP (Version 1.2): Population genetics software for exact tests and ecumenicism. Journal of Heredity 86:248-249. Reed DH, Frankham R (2001) How closely correlated are molecular and quantitative measures of genetic variation? A meta-analysis. Evolution 55: 1095-1103.

Reed DH, Frankham R (2002) Correlation between fitness and genetic diversity. Conservation Biology 17:230-237.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 190 -

Reed JC Jr., Bush CA (2005) Generalized Geologic Map of the Conterminous United States ver. 1.2. USGS, Denver, CO. http://pubs.usgs.gov/atlas/geologic/48States/.

Reeder AL, Foley GL, Nichols DK et al. (1998) Forms and prevalence of intersexuality and effects of environmental contaminants on sexuality in cricket frogs (Acris crepitans). Environmental Health Perspectives 106:261-266.

Reeder AL, Ruiz MO, Pessier A et al. (2005) Intersexuality and the cricket frog decline: historic and geographic trends. Environmental Health Perspectives 113:261-265.

Remington CL (1968) Suture-zones of hybrid interaction between recently joined biotas. Evolutionary Biology 2:321-428.

Richards CM (2000) Inbreeding depression and genetic rescue in a plant metapopulation. American Naturalist 155:383-394.

Richmond GM, Fullerton DS (1986) Summation of Quaternary glaciations in the United States of America. Pp. 183-196 in: Quaternary Glaciations in the Northern Hemisphere (eds. Sibrava V, Bowen DQ, Richmond GM). Pergamon Press, Oxford.

Rickard A, Sonntag E (2006) Blanchard's cricket frog (Acris crepitans blanchardi) morphological abnormalities in southeast Michigan. Endangered Species Update Oct­ Dec: 131-135. Rius M, Pascual M, Turon X (2008) Phylogeography of the widespread marine invader Microcosmus squamiger reveals high genetic diversity of introduced populations and non-independent colonizations. Diversity and Distributions 14:818-828.

Rivero JA (1978) The Amphibians and Reptiles of Puerto Rico. Editorial Universitaria, Universidad de Puerto Rico, Mayaguez. 152 pp.

Rivero JA, Mayorga H, Estremera E, Izquierdo I (1980) Sohre el Bufo lemur (Cope) (Amphibia, Bufonidae). Caribbean Journal of Science 15:33-40.

Rogers AR, Harpending H (1992) Population growth makes waves in the distribution of pairwise genetic differences. Molecular Biology and Evolution 9:552-569.

Rohr JR, Raffel TR, Romansic JM, McCallum H, Hudson PJ (2008) Evaluating the links between climate, disease spread, and amphibian declines. Proceedings of the National Academy of Sciences 105:17436-17441.

Ronquist F, Huelsenbeck JP (2003) MRBAYES 3: Bayesian phylogenetic inference under mixed models. Bioinformatics 19:1572-1574.

Rose FL, Simpson TR, Forstner MRJ, McHenry DJ, Williams J (2006) Taxonomic status of Acris gryllus paludicola: in search of the pink frog. Journal of Herpetology 40:428- 434.

Rowe G, Beebee TJC, Burke T (1998) Phylogeography of the natterjack toad Bufo calamita in Britain: genetic differentiation of native and translocated populations. Molecular Ecology 7:751-760.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 191 -

Rowe G, Beebee TJC, Burke T (1999) Microsatellite heterozygosity, fitness and demography in natterjack toads Bufo calamita. Animal Conservation 2:85-92.

Rowe G, Harris DJ, Beebee TJC (2006) Lusitania revisited: A phylogeographic analysis of the natterjack toad Bufo calamita across its entire biogeographical range. Molecular Phylogenetics and Evolution 39:335-346.

Ruokonen M, Kvist L, Tegelstroem H, Lumme J (2000) Goose hybrids, captive breeding and restocking of the Fennoscandian populations of the lesser white-fronted goose (Anser erythropus). Conservation Genetics 1:277-283.

Russell RW, Lipps GJ, Hecnar SJ, Haffner GD (2002) Persistent organic pollutants in Blanchard's cricket frogs (Acris crepitans blanchardi) from Ohio. Ohio Journal of Science 102: 119-122.

Ryan MJ, Wilczynski W (1988) Coevolution of sender and receiver: Effect on local mate preference in cricket frogs. Science 240: 1786-1788.

Ryan MJ, Cocroft RB, Wilczynski W (1990) The role of environmental selection in intraspecific divergence of mate recognition signals in the cricket frog, A eris crepitans. Evolution 44: 1869-1872.

Ryan MJ, Wilczynski W (1991) Evolution of intraspecific variation in the advertisement call of a cricket frog (Acris crepitans, Hylidae). Biological Journal of the Linnean Society 44:249-271.

Ryan MJ, Perrill SA, Wilczynski W (1992) Auditory tuning and call frequency predict population-based mating preferences in the cricket frog, Acris crepitans. American Naturalist 139:1370-1383.

Ryan MJ, Rand AS, Weight LA (1996) Allozyme and advertisement call variation in the tungara frog, Physalaemus pustulosus. Evolution 50:2435-2453. Salthe SN, Nevo E (1969) Geographic variation oflactate dehydrogenase in the cricket frog, Acris crepitans. Biochemical Genetics 3:335-341.

Saucier RT (1994) Geomorphology and quaternary geologic history of the lower Mississippi Valley. Mississippi River Commission, US Army Corps of Engineers, Vicksburg, MS.

Schelly R, Salzburger W, Koblmuller S, Duftner N, Sturmbauer C (2006) Phylogenetic relationships ofthe lamprologine cichlid genus Lepidiolamprologus (Teleostei: Perciformes) based on mitochondrial and nuclear sequences, suggesting introgressive hybridization. Molecular Phylogenetics and Evolution 38:426-438.

Schiesari L, Zuanon J, Azevedo-Ramos C et al. (2003) Macrophyte rafts as dispersal vectors for fishes and amphibians in the Lower Solimoes River, Central Amazon. Journal of Tropical Ecology 19:333-336.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 192 -

Schmidt KP (1928) Scientific Survey of Porto Rico and the Virgin Islands. New York Academy of Sciences, New York.

Schneider S, Excoffier L (1999) Estimation of past demographic parameters from the distribution of pairwise differences when the mutation rates vary among sites: application to human mitochondrial DNA. Genetics 152: 1079-1089.

Schwartz A, Thomas R (1975) A Check-List of West Indian Amphibian and Reptiles. Carnegie Museum of natural History, Pittsburgh. 216 pp.

Schwartz A, Henderson RW (1991) Amphibians and reptiles of the West Indies: Descriptions, distributions, and natural history. University of Florida Press, Gainesville, FL. 720 pp.

Schwartz MK, Tallmon DA, Luikart G (1998) Review of DNA-based census and effective population size estimators. Animal Conservation 1:293-299.

Schwartz MK, Luikart G, Waples RS (2007) Genetic monitoring as a promising tool for conservation and management. Trends in Ecology and Evolution 22:25-33.

Seddon PJ, Armstrong DP, Maloney RF (2007) Developing the science ofreintroduction biology. Conservation Biology 21:303-312.

Seibert EA, Lillywhite HB, Wassersug RJ (1974) Cranial coossification in frogs: relationship to rate of evaporative water loss. Physiological Zoology 47:261-265. Seigel RA, Dodd CK (2002) Translocations of amphibians: Proven management method or experimental technique? Conservation Biology 16:552-554.

Shaffer HB, McKnight ML (1996) The polytypic species revisited: genetic differentiation and molecular phylogenetics of the tiger salamander Ambystoma tigrinum (Amphibia: Caudata) complex. Evolution 50:417-433.

Shaffer HB, Fellers GM, Magee A, Voss SR (2000) The genetics of amphibian declines: population substructure and molecular differentiation in the Yosemite Toad, Bufo canorus (Anura, Bufonidae) based on single-strand conformation polymorphism analysis (SSCP) and mitochondrial DNA sequence data. Molecular Ecology 9:245- 257.

Shaffer HB, Fellers GM, Voss SR, Oliver JC, Pauly GB (2004a) Species boundaries, phylogeography and conservation genetics of the red-legged frog (Rana aurora/draytonii) complex. Molecular Ecology 13:2667-2677.

Shaffer HB, Pauly GB, Oliver JC, Trenham PC (2004b) The molecular phylogenetics of endangerment: cryptic variation and historical phylogeography of the California tiger salamander, Ambystoma californiense. Molecular Ecology 13:3033-3049.

Shogren JF, Tschirhart J, Anderson T, et al. (2001) Why economics matters for endangered species protection. Conservation Biology 13:1257-1261.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 193 -

Shuster SM, Miller MP, Lang BK et al. (2005) The effects of controlled propagation on an endangered species: genetic differentiation and divergence in body size among native and captive populations of the Socorro Isopod (Crustacea: Flabellifera). Conservation Genetics 6:355-368.

Siddle HV, Kreiss A, Eldridge MOB et al. (2007) Transmission of a fatal clonal tumor by biting occurs due to depleted MHC diversity in a threatened carnivorous marsupial. Proceedings of the National Academy of Sciences 104:16221-16226.

Skelly DK, Yurewicz KL, Werner EE, Relyea RA (2003) Estimating decline and distributional change in amphibians. Conservation Biology 17:744-751.

Smith GR, Todd A, Rettig JE, Nelson F (2003) Microhabitat selection by northern cricket frogs (Acris crepitans) along a west-central Missouri creek: Field and experimental observations. Journal of Herpetology 37:383-385.

Smith HM, Goebel AM (1994) The proper spelling and dates for certain names associated with the endemic West Indian genus of toads (Amphibia: Bufonidae). Herpetological Review 25:49-50. Smith MA, Green DM (2004) Phylogeography of Bufo fowleri at its northern range limit. Molecular Ecology 13:3723-3733.

Smith MA, Green DM (2005) Dispersal and the metapopulation paradigm in amphibian ecology and conservation: are all amphibian populations metapopulations? Ecography 28:110-128.

Smith MA, Green DM (2006) Sex, isolation and fidelity: unbiased long-distance dispersal in a terrestrial amphibian. Ecography 29:649-658.

Sodhi NS, Bickford D, Diesmos AC et al. (2008) Measuring the meltdown: drivers of global amphibian extinction and decline. PLoS ONE 3:e1636.

Soltis DE, Morris AB, McLachlan JS, Mano PS, Soltis PS (2006) Comparative phylogeography ofunglaciated eastern North America. Molecular Ecology 15:4261- 4293.

Soltis PS, Gitzendanner MA ( 1999) Molecular systematics and the conservation of rare species. Conservation Biology 13:471-483.

Song SJ, M'Gonigle RM (2002) Science, power, and system dynamics: the political economy of conservation biology. Conservation Biology 15:980-989.

Sparling DW, Lowe TP, Day D, Dolan K (1995) Responses of amphibian populations to water and soil factors in experimentally-treated aquatic macrocosms. Archives of Environmental Contamination and Toxicology 29:455-461.

Starkey DE, Shaffer HB, Burke RL et al. (2003) Molecular systematics, phylogeography, and the effects of Pleistocene glaciation in the painted turtle (Chrysemys picta) complex. Evolution 57:119-128.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 194 -

Steiner S, Lehtinen RM (2008) Occurrence of the amphibian pathogen Batrachochytrium dendrobatidis in Blanchard's cricket frog (Acris crepitans blanchardi) in the U.S. Midwest. Herpetological Review 39: 193-196.

Stejneger L ( 1902) The herpetology of Porto Rico. Pp. 549-724 in: Report of US National Museum.

Stiff BJ, Hansel AK (2004) Quaternary glaciations in Illinois. Pp. 71-81 in: Quaternary Glaciations - Extent and Chronology, Part II: North America (eds. Ehlers J, Gibbard PL). Elsevier, Amsterdam.

Storfer A, Murphy MA, Evans JS et al. (2007) Putting the "landscape" in landscape genetics. Heredity 98: 128-142.

Stuart BL, Inger RF, Voris HK (2006) High level of cryptic species diversity revealed by sympatric lineages of Southeast Asian forest frogs. Biology Letters 2:470-474.

Stuart SN, Chanson JS, Cox NA et al. (2004) Status and trends of amphibian declines and extinctions worldwide. Science 306: 1783-1786.

Swenson NG, Howard DJ (2005) Clustering of contact zones, hybrid zones, and phylogeographic breaks in North America. American Naturalist 166:581-591.

Tajima F (1989) Statistical method for testing the neutral mutation hypothesis by DNA polymorphism. Genetics 123:585-595. Tallmon DA, Luikart G, Waples RS (2004) The alluring simplicity and complex reality of genetic rescue. Trends in Ecology and Evolution 19:489-496.

Tamura K, Nei M (1993) Estimation of the number of nucleotide substitutions in the control region of mitochondrial DNA in humans and chimpanzees. Molecular Biology and Evolution 10:512-526.

Tarr CL, Fleischer RC (1999) Population boundaries and genetic diversity in the endangered Mariana crow (Corvus kubaryi). Molecular Ecology 8:941-949.

Tautz D, Arctander P, Minelli A, Thomas RH, Vogler AP (2003) A plea for DNA taxonomy. Trends in Ecology and Evolution 18:70-74.

Tennessen JA, Zamudio KR (2003) Early male reproductive advantage, multiple paternity and sperm storage in an amphibian aggregate breeder. Molecular Ecology 12:1567-1576.

Teplitsky C, Mills JA, Alho JS, Yarrall JW, Merila J (2008) Bergmann's rule and climate change revisited: Disentangling environmental and genetic responses in a wild bird population. Proceedings of the National Academy of Sciences 105: 13492-13496.

Theodorakis CW, Rinchard J, Carr JA et al. (2006) Thyroid endocrine disruption in stonerollers and cricket frogs from perchlorate-contaminated streams in east-central Texas. Ecotoxicology 15:31-50.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 195 -

Thompson JD, Gibson TJ, Plewniak F, Jeanmougin F, Higgins DG (1997) The CL UST AL_x windows interface: flexible strategies for multiple sequence alignment aided by quality analysis tools. Nucleic Acids Research 25:4876-4882.

Thornbury WD (1965) Regional Geomorphology of the United States. John Wiley and Sons, Inc, New York.

Trapnell DW, Schmidt JP, Quintana-Ascencio PF, Hamrick JL (2007) Genetic insights into the biogeography of the southeastern North American endemic, Ceratiola ericoides (Empetraceae). Journal of Heredity 98:587-593.

Trauth SE, Robison HW, Plummer MV (2004) The Amphibians and Reptiles of Arkansas. University of Arkansas Press, Fayetville, AR. 440 pp.

Trenham PC, Marsh DM (2002) Amphibian translocation programs: Reply to Seigel and Dodd. Conservation Biology 16:555-556.

USFWS (1992) Recovery plan for the Puerto Rican crested toad (Peltophryne lemur). U.S. Fish and Wildlife Service, Southeast Region, Atlanta, Georgia.

Van Gorp CD (2001) Changes in the distribution of A eris crepitans blanchardi, with studies of nesting microhabitat and the possible impact ofultraviolet-B radiation. Unpublished M.A. dissertation. Drake University.

Van Oosterhout C, Weetman D, Hutchinson WF (2006) Estimation and adjustment of microsatellite null alleles in nonequilibrium populations. Molecular Ecology Notes 6:255-256.

Vandergast AG, Bohonak AJ, Hathaway SA, Boys J, Fisher RN (2008) Are hotspots of evolutionary potential adequately protected in southern California? Biological Conservation 141: 1648-1664.

Vandewoestijne S, Schtickzelle N, Baguette M (2008) Positive correlation between genetic diversity and fitness in a large, well-connected metapopulation. BMC Biology 6:46.

Veldman RB (1997) Trend toward extinction of the cricket frog (Acris crepitans): prediction of declines using field data and the Vortex simulation program. Unpublished M.Sc. dissertation. Illinois State University.

Vianna JA, Bonde RK, Caballero S et al. (2006) Phylogeography, phylogeny and hybridization in trichechid sirenians: implications for manatee conservation. Molecular Ecology 15:433-447.

Vieites DR, Min MS, Wake DB (2007) Rapid diversification and dispersal during periods of global warming by plethodontid salamanders. Proceedings of the National Academy of Sciences 104: 19903-19907.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 196 -

Vila C, Sundqvist A-K, Flagstad O et al. (2003) Rescue of a severely bottlenecked wolf (Canis lupus) population by a single immigrant. Proceedings of the Royal Society of London B 270:91-97.

Viosca PA (1923) An ecological study of the cold-blooded vertebrates of southeastern Louisiana. Copeia 1923:35-44.

Viosca PA ( 1944) Distribution of certain cold-blooded animals in Louisiana in relationship to the geology and physiography of the state. Proceedings of the Louisiana Academy of Sciences 8:4 7-62.

Volis S, Mendlinger S, Olsvig-Whittaker L, Safriel UN, Orlovsky N (1998) Phenotypic variation and stress resistance in core and peripheral populations of Hordeum spontaneum. Biodiversity and Conservation 7:799-813.

Vredenburg VT, Bingham R, Knapp R et al. (2007) Concordant molecular and phenotypic data delineate new taxonomy and conservation priorities for the endangered mountain yellow-legged frog. Journal of Zoology 271:361-374.

Wake DB, Vredenburg V (2008) Are we in the midst of the sixth mass extinction? A view from the world of amphibians. Proceedings of the National Academy of Sciences 105:11466-11473.

Wagner AP, Creel S, Kalinowski ST (2006) Estimating relatedness and relationships using microsatellite loci with null alleles. Heredity 97:336-345. Waldick RC, Kraus S, Brown M, White BN (2002) Evaluating the effects of historic bottleneck events: an assessment of microsatellite variability in the endangered, North Atlantic right whale. Molecular Ecology 11:2241-2249.

Waldman B, Rice JE, Honeycutt RL (1992) Kin recognition and incest avoidance in toads. American Zoologist 32: 18-30.

Wang JY, Chou LS, White BN (1999) Mitochondrial DNA analysis of sympatric morphotypes ofbottlenose dolphins (genus: Tursiops) in Chinese waters. Molecular Ecology 8:1603-1612.

Wang JY, Frasier TR, Yang SC, White BN (2008) Detecting recent speciation events: the case of the finless porpoise (genus Neophocaena). Heredity 101:145-155.

Ward R, Rutledge CJ, Zimmerman EG (1987) Genetic variation and population subdivision in the cricket frog Acris crepitans. Biochemical Systematics and Ecology 15:377-384.

Webb SD (1990) Historical Biogeography. Pp. 70-100 in: Ecosystems of Florida (eds. Myers RL, Ewel JJ). University of Central Florida Press, Orlando.

Weber JL, Wong C (1993) Mutation of human short tandem repeats. Human Molecular Genetics 2:1123-1128.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 197 -

Weir BS, Cockerham CC (1984) Estimating F-statistics for the analysis of population structure. Evolution 38:1358-1370.

Westemeier RL, Brawn JD, Simpson SA et al. (1998) Tracking the long-term decline and recovery of an isolated population. Science 282: 1695-1698.

Whitehouse AM, Harley EH (2001) Post-bottleneck genetic diversity of elephant populations in South Africa, revealed using microsatellite analysis. Molecular Ecology 10:2139-2149.

WI DNR (2005) Wisconsin's Strategy for Wildlife Species of Greatest Conservation Need. Wisconsin Department of Natural Resources, Madison, WI.

Widder PD, Bidwell JR (2008) Tadpole size, cholinesterase activity, and swim speed in four frog species after exposure to sub-lethal concentrations of chlorpyrifos. Aquatic Toxicology 88:9-18.

Wiens JJ, Graham CH, Moen DS, Smith SA, Reeder TW (2006) Evolutionary and ecological causes of the latitudinal diversity gradient in hylid frogs: treefrog trees unearth the roots of high tropical diversity. The American Naturalist 168:579-596.

Wilczynski W, Keddy-Hector AC, Ryan MJ (1992) Call patterns and basilar papilla tuning in cricket frogs. 1. Differences among populations and between sexes. Brain, Behavior and Evolution 39:229-237.

Williams DA, Muchugu E, Overholt WA, Cuda JP (2007) Colonization patterns of the invasive Brazilian peppertree, Schinus terebinthifolius, in Florida. Heredity 98:284- 293.

Wilson GA, Fulton TL, Kendell Ket al. (2008) Genetic diversity and structure in Canadian northern leopard frog (Rana pipiens) populations: implications for reintroduction programs. Canadian Journal of Zoology 86:863-874.

Wilson PJ, Grewal S, Lawford ID, et al. (2000) DNA profiles of the eastern Canadian wolf and the red wolf provide evidence for a common evolutionary history independent of the gray wolf. Canadian Journal of Zoology 78:2156-2166.

Wright S (1951) The genetical structure of populations. Annals of Eugenics 15:323-354.

Wu H, Wan QH, Fang SG, Zhang SY (2005) Application of mitochondrial DNA sequence analysis in the forensic identification of Chinese sika deer subspecies. Forensic Science International 148:101-105.

Wycherley J, Doran S, Beebee TJC (2002) Frog calls echo microsatellite phylogeography in the European pool frog (Rana lessonae). Journal of Zoology 258:479-484.

Wyman RL (1990) What's happening to the amphibians? Conservation Biology 4:350- 352.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 198 -

Wyner YM, Amato G, Desalle R (1999) Captive breeding, reintroduction, and the conservation genetics of black and white ruffed lemurs, Varecia variegata variegata. Molecular Ecology 8:S107-Sl 15.

Zamudio KR, Savage WK (2003) Historical isolation, range expansion, and secondary contact of two highly divergent mitochondrial lineages in spotted salamanders (Ambystoma maculatum ). Evolution 57: 1631-1652.

Zamudio KR, Wieczorek AM (2007) Fine-scale spatial genetic structure and dispersal among spotted salamander (Ambystoma maculatum) breeding populations. Molecular Ecology 16:257-274.

Zeisset I, Beebee T (2001) Determination of biogeographical range: an application of molecular phylogeography to the European pool frog Rana lessonae. Proceedings of the National Academy of Sciences 268:933-938.

Zeisset I, Beebee TJC (2008) Amphibian phylogeography: a model for understanding historical aspects of species distributions. Heredity 101: 101-119.

Zink RM, Barrowclough GF, Atwood JL, Blackwell-Rago RC (2000) Genetics, taxonomy, and conservation of the threatened California gnatcatcher. Conservation Biology 14:1394-1405.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. haplotype(s)

Appendix

mitochondrial

Delaware Arkansas Arkansas Arkansas Arkansas Delaware Arkansas Arkansas Alabama Alabama Alabama Arkansas

Arkansas

Georgia

Georgia

Georgia Georgia

Florida

Florida

Florida

Florida Florida

Illinois Illinois

Illinois

State

1.

observed,

Details

DNA

(Abl

of

nucleotide

sample

New

Tuscaloosa

Tallapoosa

Champaign

Santa

Covington

=

Effingham

Columbia

Lafayette

Sebastian

Franklin

Gadsden

Garland

County

Monroe

Putnam Benton

Liberty

Greene

Collier Fulton

A.

Brown

Taylor

Baker

Stone

Kent

blanchardi; Castle

Rosa

collection

diversity

sites

Acr

(nn;

14

20

N

22

10

5 3

2

2 3

2

6 5

3 1

5 2

9

4

1 4 5

2

1 8

5

=

(counties)

for

A.

sites

crepitans;

with>

221,223,224,225,226

207,242-244,247,248

for

207,208,209,210

175,207,

54,59,63,81,94

205,221,222 89,90,91,92

A

202,

207,249,250

225,

25,

Ha£lotypes

Agr

eris

1

174,200

245,246

174,

176,

1,

98,99

96,97

73,99

individual),

100,

1,

211-220

172

173

251-253

66-69

=

93

36

specimens.

1

45

A.

227-241 175

177

101

gryllus),

species

The

and

number

coordinates

present

0.0033 0.0028

0.0063

0.0064

0.0125 0.0106 0.0042 0.0159

0.0042

0.0059 0.0229 0.0273

0.0014 0.0041 0.0063 0.0031

0.0013

0.007

0.027

0.004

n/a

n/a

n/a

7l"n

of

0

based

0

individuals

in

on

Acr/Agr

decimal

Species

Agr

Agr

Acr

Abl Abl

Abl Abl Agr Abl Abl Abl

Agr Abl Acr Acr Agr

Agr Agr Agr Acr

Acr Abl Abl

Abl

phylogenetic

(N),

degrees

Latitude

32.8501 33.2356 36.3542 33.2494 35.4776

33.2723 35.2939 35.8741 34.5453

30.2875

26.1551 39.0927 39.6956

30.6141 33.5623 31.8095 30.6033 40.1368 32.5499

33.0324 39.9503

39.0697

31.251

36.355

29.598

mitochondrial

analysis

are

provided.

Longitude

-85.83501

-87.51169

-75.56027

-82.23627

-75.61171

-81.67131 -84.61833

-87.02181 -83.15764

-81.75851 -84.22702

-83.91935

-86.4162

-93.2298 -93.8845 -91.7293 -94.2468 -93.5631 -94.3518 -92.1699 -93.0759

-81.5388

-90.7402 -88.2168

-88.5923

of

-

199

-

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Indiana Indiana Indiana

Indiana Indiana Indiana Indiana Indiana

Illinois Illinois

Kansas

Kansas Kansas Kansas Kansas Kansas Kansas Kansas

Kansas

Kansas Kansas Kansas Kansas

Iowa Iowa Iowa

Iowa Iowa Iowa Iowa

Leavenworth

Washington

Woodbury

Anderson

Cherokee

Jennings Madison

Atchison

Jefferson

Bourbon

Monona

Douglas

Marion

Greene

Marion Cowley

Harvey Marion

Brown

Henry White

Kiowa

Posey

Parke

Lake Story

Gove

Linn

Pike Polk

21

9

5

5

3 5 5 4 5 5 9

4 3

3 2 3 7 2

2 1 3 2

3

1

1

1 1 1

1

1

1,

1,

1,5,6,45

178,

36,

12,

50,

25,36 1, 20,21 1,

1, 1, 1,

1,

1,

22

23

23 24

25 54

6,

1 3-7

1 1

1

1 1 1

1

1

1 42

41

14 70-72

83 13, 20

51

179

7

14

0.0022 0.0025 0.0049 0.0013

0.0013 0.0014 0.0017 0.0038

0.0028 0.0035

0.0021

0.0167 0.0063

0.025

n/a

n/a n/a

n/a n/a n/a n/a n/a

0

0

0 0 0 0 0

0

Abl Abl Abl Abl

Abl Abl Abl Abl Abl Acr Abl

Abl Abl Abl Abl Abl Abl Abl Abl Abl Abl

Abl

Abl Abl Abl Abl Abl Abl Abl Abl

38.1041 41.5296 39.0008 38.8125

39.0515 39.7638

39.7928 38.5896

42.0504 40.9824 38.0259

42.0242 41.3256 42.4385 41.6279

39.5431

39.8131 38.2148

37.8561

37.1421 37.2362 39.2344 38.9907 38.0428

37.5793 39.2267 38.3385

38.412

38.927

42.03

-86.09681

-85.6268 -90.0188 -88.1801 -87.0374 -87.2192 -87.3953 -86.1418

-87.2371 -87.8488

-91.5428

-93.0692 -95.9537 -93.6309 -91.6306 -95.2527 -93.5287 -96.2331 -95.5627 -95.2833

-94.7869

-94.7874

-97.4085 -96.9436

-99.2578 -100.475 -94.9798 -97.0861

-95.256 -95.391

-

200

-

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Louisiana Louisiana Louisiana Kentucky Kentucky Kentucky

Louisiana Louisiana Louisiana Louisiana Louisiana Louisiana Louisiana Louisiana Louisiana Louisiana Louisiana oiin Jefferson Louisiana

Louisiana Louisiana

Kansas

Kansas Kansas Kansas Kansas Kansas

Kansas Kansas Kansas Kansas

East

Pointe Pottawatomie

East

Montgomery

Natchitoches

Washington

Wabaunsee

Evangeline Henderson

Ascension

Livingston

Avoyelles

Calcasieu

Shawnee

Bienville

Cameron

Bracken

Wallace

Pawnee

Iberville

Baton

La

Bossier

Morris

Miami

Meade

Osage

Allen

Grant

Feliciana

Salle

Coupee

Rouge

3

3 5

5 1

1 4 5 2 2 7 1 1

1 9 1 3 1 5

2 7 5 1 4

1

8 1 5 1 1

114,

114,

127,

202,251,255-257,274

107-10~ 119,

3,15 4,16 160, 156, 143, 135, 134,

129,149,152,153

18,

118,

132, 157, 158, 157, 132,

263,

25,46,47

25,28,53

19,

1,

1,27,28

104, 129,

147,

112,

145-149

120-122,

22,24

15,

127,128,162

181

119 112

133 113 106 26

275-277 54 23,

55 27

1

1

130

105

151

142

180

52,

131

54

159,

164

161

0.0167 0.0013 0.0167

0.0117

0.0608

0.0063 0.0165 0.0356

0.0201 0.0196 0.0158 0.0188

0.0176

0.0133

0.0159

0.013

0.007

0.02

n/a

n/a

n/a n/a n/a

n/a

n/a

n/a n/a

n/a

n/a

n/a

Abl/Acr

Ahl

Ahl Ahl Ahl Ahl Ahl Ahl Ahl

Ahl Ahl Abl Acr

Ahl Ahl Ahl Ahl Agr Ahl Ahl Ahl Ahl

Ahl Ahl Agr Ahl

Ahl

Ahl Ahl

Ahl

38.5685 37.1481 38.7032 38.6398 38.1754 39.3366 39.0392 38.9747 38.8998 39.7771

37.8026 37.9489

31.0541 32.3934 30.6537 29.8732 30.4801 32.5946 30.2038 30.2374

31.5832 30.8382 30.6873

30.2752

31.7298 29.9352 30.4739 31.7298 30.6472

38.731

-86.13037

-93.29176 -92.05937 -93.04347 -91.11965 -93.65335 -90.94097 -91.07849 -93.21489 -92.56922

-92.37796

-91.28719

-92.18556 -90.80364 -93.09522 -91.53067

-95.7011

-95.7274 -94.8513 -99.2075 -96.3059 -96.6403 -95.7103 -96.2024 -101.758 -97.0772 -84.0809

-87.5681

-92.8218

-90.1514

-

201

-

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Mississippi

Mississippi Mississippi Mississippi Mississippi Mississippi Mississippi

Mississippi Mississippi Mississippi

Louisiana Louisiana

Louisiana Louisiana Louisiana Louisiana Louisiana Louisiana

Louisiana Louisiana Louisiana Louisiana

Louisiana Maryland

Maryland Maryland

Michigan Michigan

Missouri Missouri

St.

West

St.

John

Washington Tangipahoa

St. Tishomingo Washtenaw

Vermillion

St.

St.

St.

St.

Wilkinson

Lenawee

Rapides

Hancock Fredrick Webster

Caldwell Harrison

Grenada

Charles

Tammany

Carroll

Greene Forrest

Union

Perry

Charles

Helena

Cole Martin

Tate

Feliciana Mary's

James

the

Baptist

14

20

3

3

6 2 1 2 1

9 3 1 6 6 4

7 1 5

9 1 5 2

6 4

3 5 1 3

1

1

255,258,261,265-267,278,279,281

163,202,259,260,264,268-273

111,

105,123,124,137,150,155

110,111,112,116,117

260,262,284,296-298

141,

1, 107,

202,282,283

262, 256,300,301 278,

41,

262,268

136,

143,

125,126 102,

172,

202,282

202-204

280

115 138,

140

183

202 165

299 289-295 285-288

166-168

13

1

1

1

137

144, 154, 144,

103

182

139

155

0.0237

0.0167

0.0143

0.0083 0.0125 0.0347 0.0105 0.0109 0.0129 0.0191

0.0014

0.0158

0.0094 0.0125

0.0094

0.0042

0.009

0.018

0.01

n/a

n/a

n/a

n/a

n/a

n/a

n/a

n/a

0

0

0

Abl/Agr

Agr

Ahl Ahl

Agr Ahl Ahl Ahl Agr Ahl Ahl

Acr Ahl Ahl

Acr Agr Ahl Agr Ahl Agr Ahl Agr Ahl Agr Agr Agr Agr

Agr

Ahl

Ahl

29.9424 31.2621

30.8136 30.0247

30.0652 30.6074 30.1863 30.3984

32.8074 32.7292 30.8297 38.5221 30.8675

41.9176 39.4637 42.2603 38.2939

33.4598

31.2171

30.3581 30.4279 34.7289 31.2146 34.6447

31.1756 39.6593

29.962

31.263

33.775

38.54

-92.47659 -90.39639 -90.72945 -90.75803

-91.78212 -90.53133 -90.46295 -92.38723 -89.93062 -92.22191

-91.38578 -89.97935 -76.97227 -77.41393 -76.59763

-88.63965 -89.90714 -89.27621 -89.46246 -89.80345 -88.99095 -89.07954 -88.24274 -89.96467

-91.26623

-93.3432

-84.0659

-83.7595

-92.2407

-93.979

-

202

-

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. North North North North North North North North

New

Nebraska

Nebraska Nebraska Nebraska Nebraska

Nebraska

Missouri Missouri Missouri Missouri Missouri Missouri Missouri Missouri Missouri

Ohio Ohio Ohio Ohio Ohio Ohio

Carolina

Carolina Carolina Carolina Carolina Carolina

Carolina

Carolina

Jersey

New

Mecklenburg

Washington

Cape

Nodaway

Crawford

Reynolds

Jefferson

Auglaize

Johnson

Jackson

Durham

Pawnee

Pulaski

Greene

Clinton

Gaston

Wayne Adams

Greene

Moore

Fulton Cedar

Butler

Gates Knox

Rock

Linn

Cass

Ray

Hanover

May

13

17

4

4 5 4 2 5 3 5

5

2 4 2

4 1 4 7 1 1

1 5 1

5 1 5

5 3

5

5

172,187,310,311,312,313,316

188,311,314,316,

311,314,315

1,20,36,80

33,36,44

25,33,34

25,37,38

1,

186,187 172,205 184,

302-305 306-309

1,

1,

1,

1,

35,

184

18 18 18 18

1

1

1

1

1

188 1

1 1

23 39

40

185

36

·

317,318

0.0054 0.0042

0.0033

0.0153

0.0021

0.0113 0.0125

0.0042

0.0618

0.0985 0.0084 0.0038

0.0426 0.0241

0.0834 0.0358

0.0012

n/a n/a n/a n/a n/a n/a

0

0

0

0 0

0

0

Acr/Agr

Acr/Agr

Acr/Agr

Abl/Acr

Abl Abl Abl Abl Abl Abl Abl Abl

Abl Abl Abl Abl Abl Abl Abl Acr Agr Acr Agr Acr

Agr

Abl Abl

Abl

Abl Abl

37.2163 40.3597

38.0252

40.9365 42.6053 38.7455 39.8331

40.1386 37.8465 42.5071 39.3211 37.3658 42.6292

39.0881 35.9983 35.2845 36.4499 35.2318 35.2378

40.5472 35.8553 35.3647 38.8464 41.5966 34.2024 39.4279 39.4191

39.7126

39.024

40.162

-81.16645

-78.90036

-80.83086 -79.45015

-76.61634 -77.99534

-77.89504

-91.3263 -93.3161

-94.8779 -94.4632 -93.7963 -93.0869

-92.1989 -94.0232

-96.2086 -96.0934 -97.2399 -97.1451 -97.8545 -90.9778

-99.4701 -74.7991

-84.5132 -83.8286 -83.5022 -84.2546

-84.1245 -83.9494

-76.743

-

203

-

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. South

South South South South

South

South

South

Oklahoma Oklahoma Oklahoma Oklahoma Oklahoma Oklahoma

Oklahoma Oklahoma Oklahoma Oklahoma Oklahoma Oklahoma

Tennessee

Tennessee

Ohio

Ohio Ohio

Ohio

Ohio Ohio Ohio Ohio

Carolina Carolina Carolina Carolina

Carolina Carolina

Dakota

Dakota

Montgomery

Cumberland

Darlington

Muskogee

Hamilton

Abbeville Cherokee Sequoyah Canadian

Paulding

Barnwell

Wagoner

Berkeley

McNairy

Yankton

Leflore

Warren

Pickens

Miami

Marion

Kiowa

Custer Preble

Lucas

Logan

Atoka

Wood

Roger

Union

Ellis

16

18

10 4

10 5

5

5 5 7 5

4 5 2 2 6

3

5 4 2 2

6 5 6

2 1 2

4 2 5

20,25,28,36,60,61

207,221,

22,22,57,64,65

174,176,192

172,193,201

22,62,63

57,58,86

20,36,59

334,335

203,206 336,337

194,197

188-191

22,87

59,95

20,22 23,32 54,56

1,

1,

22

81

19

1 1 1 1

1 1 1

19

2

319-335

0.0008

0.0167 0.0176 0.0251 0.0021 0.0125

0.0121 0.0125 0.0058 0.0167 0.0208 0.0101

0.0085 0.0146 0.0056

0.0004

0.0088 0.0104

0.0487

0.027

n/a

0 0 0 0

0 0 0

0

0

Ahl Ahl Ahl Ahl Ahl Ahl Ahl Ahl Ahl Ahl Ahl

Ahl Ahl Ahl Ahl Ahl Agr Agr Ahl Acr Ahl Ahl Ahl

Agr

Agr Acr

Acr Ahl Ahl Acr

41.6493 40.0593

41.1184 39.7503 39.1795

41.4051 39.4402

39.7439 35.5223

34.3821

35.8872 35.6015 35.8933 36.2681 34.9307 34.9764 35.4844 35.6866 34.2376 35.6702 33.3038 35.9705

42.8246 42.9604 33.1271 34.3295 34.1554 34.8288 35.1799

35.9462

-82.45252 -81.35233 -79.98064

-79.33633

-88.55074

-84.99825

-84.2186 -84.4938 -83.6085 -84.2426

-84.2218 -84.5863 -84.6372 -97.8963 -83.6104 -96.0861

-99.7953 -98.9964 -94.6778 -98.9202 -94.7889 -97.4429 -95.3671 -99.6673

-95.5448

-96.6576 -97.3853 -82.7015

-95.001

-79.978

-

204

-

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. Tennessee

Virginia

Texas Texas Texas Texas

Texas

Texas

Texas Texas

Montgomery

Charles

McMullen

Anderson

Kendall

Menard

Bowie

Travis

Hood

Wise

City

30

11

4 4 2

1

8

1

1

1

8-11,30,48,84,85,88

169-171,

10,

49,

6 17, 16,

11,30,

195,

183,

43

79,82

30 29

31

196

73

74-78

198,

199

0.0059

0.0042 0.0042

0.0538 0.0144

0.013

n/a

n/a n/a

n/a

Abl/Acr

Acr Abl Abl Abl Abl Abl Abl Abl Abl

36.5106 31.7975 33.4381 28.3561 32.4323

30.8973 30.3218 33.2059 37.3449

29.898

-87.37043

-77.07161

-95.6299

-97.7962 -98.7308 -98.5378 -94.2502 -99.8102 -97.7698 -97.6743

-

205

-

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 206 -

Appendix 2. Details of mitochondrial control region haplotypes found for A eris species. The number of occurrences, states in which they were found, and GenBank accession numbers are provided.

Haplotype Number States GenBankNo. Acris blanchardi OH, IA, IL, IN, KS, AcrCR-1 226 EU828245 KY, MI, MO, NE, SD AcrCR-2 1 OH EU828246 AcrCR-3 1 IL EU828247 AcrCR-4 1 IL EU828248 AcrCR-5 6 IL,IN EU828249 AcrCR-6 6 IL,IN EU828250 AcrCR-7 3 IL,IN EU828251 AcrCR-8 2 TX EU828252 AcrCR-9 1 TX EU828253 AcrCR-10 7 TX EU828254 AcrCR-11 8 TX EU828255 AcrCR-12 1 IA EU828256 AcrCR-13 4 IA,MO EU828257 AcrCR-14 5 IA, IN EU828258 AcrCR-15 1 KY EU828259 AcrCR-16 2 TX EU828260 AcrCR-17 1 TX EU828261 AcrCR-18 5 KS,NE EU828262 AcrCR-19 20 KS,SD EU828263 AcrCR-20 9 KS,MO, OK EU828264 AcrCR-21 1 KS EU828265 AcrCR-22 15 KS,OK EU828266 AcrCR-23 8 KS,MO,OK EU828267 AcrCR-24 4 KS EU828268 AcrCR-25 8 KS, MO, OK, AR EU828269 AcrCR-26 1 KS EU828270 AcrCR-27 3 KS EU828271 AcrCR-28 3 KS,OK EU828272 AcrCR-29 1 TX EU828273 AcrCR-30 4 TX EU828274 AcrCR-31 1 TX EU828275 AcrCR-32 1 OK EU828276 AcrCR-33 3 MO EU828277 AcrCR-34 1 MO EU828278 AcrCR-35 1 MO EU828279 AcrCR-36 11 IL, KS, MO, OK, AR EU828280 AcrCR-37 1 MO EU828281 AcrCR-38 2 MO EU828282 AcrCR-39 1 MO EU828283

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 207 -

AcrCR-40 3 OH EU828284 AcrCR-41 1 IN EU828285 AcrCR-42 1 IN EU828286 AcrCR-43 1 TX EU828287 AcrCR-44 2 MO EU828288 AcrCR-45 4 IN, IL EU828289 AcrCR-46 1 KS EU828290 AcrCR-47 1 KS EU828291 AcrCR-48 11 TX EU828292 AcrCR-49 2 TX EU828293 AcrCR-50 3 IN EU828294 AcrCR-51 2 IN EU828295 AcrCR-52 1 KS EU828296 AcrCR-53 1 KS EU828297 AcrCR-54 5 KS,OK,AR EU828298 AcrCR-55 1 KS EU828299 AcrCR-56 1 OK EU828300 AcrCR-57 2 OK EU828301 AcrCR-58 1 OK EU828302 AcrCR-59 4 OK,AR EU828303 AcrCR-60 1 OK EU828304 AcrCR-61 1 OK EU828305 AcrCR-62 1 OK EU828306 AcrCR-63 2 OK,AR EU828307 AcrCR-64 1 OK EU828308 AcrCR-65 1 OK EU828309 AcrCR-66 1 IL EU828310 AcrCR-67 1 IL EU828311 AcrCR-68 1 IL EU828312 AcrCR-69 1 IL EU828313 AcrCR-70 2 IL EU828314 AcrCR-71 2 IL EU828315 AcrCR-72 1 IL EU828316 AcrCR-73 2 TX,AR EU828317 AcrCR-74 1 TX EU828318 AcrCR-75 1 TX EU828319 AcrCR-76 2 TX EU828320 AcrCR-77 1 TX EU828321 AcrCR-78 1 TX EU828322 AcrCR-79 1 TX EU828323 AcrCR-80 1 MO EU828324 AcrCR-81 2 OK,AR EU828325 AcrCR-82 1 TX EU828326 AcrCR-83 1 KS EU828327 AcrCR-84 1 TX EU828328 AcrCR-85 1 TX EU828329

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 208 -

AcrCR-86 1 OK EU828330 AcrCR-87 1 OK EU828331 AcrCR-88 1 TX EU828332 AcrCR-89 1 AR EU828333 AcrCR-90 2 AR EU828334 AcrCR-91 2 AR EU828335 AcrCR-92 1 AR EU828336 AcrCR-93 1 AR EU828337 AcrCR-94 1 AR EU828338 AcrCR-95 1 OK EU828339 AcrCR-96 1 AR EU828340 AcrCR-97 1 AR EU828341 AcrCR-98 1 AR EU828342 AcrCR-99 2 AR EU828343 AcrCR-100 1 AR EU828344 AcrCR-101 1 AR EU828345 AcrCR-102 1 LA FJ829506 AcrCR-103 1 LA FJ829507 AcrCR-104 1 LA FJ829508 AcrCR-105 2 LA FJ829509 AcrCR-106 1 LA FJ829510 AcrCR-107 2 LA FJ829511 AcrCR-108 1 LA FJ829512 AcrCR-109 2 LA FJ829513 AcrCR-110 2 LA FJ829514 AcrCR-111 2 LA FJ829515 AcrCR-112 3 LA FJ829516 AcrCR-113 1 LA FJ829517 AcrCR-114 2 LA FJ829518 AcrCR-115 1 LA FJ829519 AcrCR-116 1 LA FJ829520 AcrCR-117 1 LA FJ829521 AcrCR-118 1 LA FJ829522 AcrCR-119 2 LA FJ829523 AcrCR-120 1 LA FJ829524 AcrCR-121 1 LA FJ829525 AcrCR-122 1 LA FJ829526 AcrCR-123 1 LA FJ829527 AcrCR-124 1 LA FJ829528 AcrCR-125 2 LA FJ829529 AcrCR-126 1 LA FJ829530 AcrCR-127 2 LA FJ829531 AcrCR-128 1 LA FJ829532 AcrCR-129 2 LA FJ829533 AcrCR-130 1 LA FJ829534 AcrCR-131 1 LA FJ829535

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 209 -

AcrCR-132 1 LA FJ829536 AcrCR-133 1 LA FJ829537 AcrCR-134 1 LA FJ829538 AcrCR-135 1 LA FJ829539 AcrCR-136 2 LA FJ829540 AcrCR-137 2 LA FJ829541 AcrCR-138 1 LA FJ829542 AcrCR-139 1 LA FJ829543 AcrCR-140 1 LA FJ829544 AcrCR-141 1 LA FJ829545 AcrCR-142 1 LA FJ829546 AcrCR-143 2 LA FJ829547 AcrCR-144 1 LA FJ829548 AcrCR-145 1 LA FJ829549 AcrCR-146 1 LA FJ829550 AcrCR-147 2 LA FJ829551 AcrCR-148 1 LA FJ829552 AcrCR-149 3 LA FJ829553 AcrCR-150 1 LA FJ829554 AcrCR-151 2 LA FJ829555 AcrCR-152 1 LA FJ829556 AcrCR-153 1 LA FJ829557 AcrCR-154 1 LA FJ829558 AcrCR-155 2 LA FJ829559 AcrCR-156 1 LA FJ829560 AcrCR-157 1 LA FJ829561 AcrCR-158 1 LA FJ829562 AcrCR-159 2 LA FJ829563 AcrCR-160 2 LA FJ829564 AcrCR-161 1 LA FJ829565 AcrCR-162 1 LA FJ829566 AcrCR-163 1 LA FJ829567 AcrCR-164 1 LA FJ829568 Acris crepitans AcrCR-165 1 MD FJ829569 AcrCR-166 1 MS FJ829570 AcrCR-167 1 MS FJ829571 AcrCR-168 1 MS FJ829572 AcrCR-169 1 VA FJ829573 AcrCR-170 1 VA FJ829574 AcrCR-171 1 VA FJ829575 AcrCR-172 14 SC, DE, NC, MD FJ829576 AcrCR-173 1 DE FJ829577 AcrCR-174 5 SC,GA,AL FJ829578 AcrCR-175 3 FL,GA FJ829579 AcrCR-176 2 GA,SC FJ829580

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 210 -

AcrCR-177 1 GA FJ829581 AcrCR-178 3 IN FJ829582 AcrCR-179 2 IN FJ829583 AcrCR-180 2 KY FJ829584 AcrCR-181 1 KY FJ829585 AcrCR-182 2 MD FJ829586 AcrCR-183 2 MD,VA FJ829587 AcrCR-184 5 NC FJ829588 AcrCR-185 1 NC FJ829589 AcrCR-186 1 NJ FJ829590 AcrCR-187 2 NJ,NC FJ829591 AcrCR-188 4 OH,SC,NC FJ829592 AcrCR-189 1 SC FJ829593 AcrCR-190 1 SC FJ829594 AcrCR-191 1 SC FJ829595 AcrCR-192 2 SC FJ829596 AcrCR-193 2 SC FJ829597 AcrCR-194 2 TN FJ829598 AcrCR-195 1 TN FJ829599 AcrCR-196 1 TN FJ829600 AcrCR-197 2 TN FJ829601 AcrCR-198 3 VA FJ829602 AcrCR-199 1 VA FJ829603 AcrCR-200 2 AL FJ829604 AcrCR-201 2 SC FJ829605 Acris gryllus AcrCR-202 14 MS,AL,LA FJ829606 AcrCR-203 2 MS,TN FJ829607 AcrCR-204 1 MS FJ829608 AcrCR-205 2 NC,FL FJ829609 AcrCR-206 1 TN FJ829610 AcrCR-207 11 AL, FL, GA, SC FJ829611 AcrCR-208 1 AL FJ829612 AcrCR-209 1 AL FJ829613 AcrCR-210 1 AL FJ829614 AcrCR-211 1 AL FJ829615 AcrCR-212 1 AL FJ829616 AcrCR-213 2 AL FJ829617 AcrCR-214 1 AL FJ829618 AcrCR-215 1 AL FJ829619 AcrCR-216 1 AL FJ829620 AcrCR-217 1 AL FJ829621 AcrCR-218 1 AL FJ829622 AcrCR-219 1 AL FJ829623 AcrCR-220 1 AL FJ829624 AcrCR-221 4 FL,SC FJ829625

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 211 -

AcrCR-222 1 FL FJ829626 AcrCR-223 1 FL FJ829627 AcrCR-224 1 FL FJ829628 AcrCR-225 6 FL,GA FJ829629 AcrCR-226 1 FL FJ829630 AcrCR-227 1 FL FJ829631 AcrCR-228 2 FL FJ829632 AcrCR-229 1 FL FJ829633 AcrCR-230 1 FL FJ829634 AcrCR-231 1 FL FJ829635 AcrCR-232 1 FL FJ829636 AcrCR-233 1 FL FJ829637 AcrCR-234 1 FL FJ829638 AcrCR-235 1 FL FJ829639 AcrCR-236 1 FL FJ829640 AcrCR-237 1 FL FJ829641 AcrCR-238 1 FL FJ829642 AcrCR-239 1 FL FJ829643 AcrCR-240 1 FL FJ829644 AcrCR-241 1 FL FJ829645 AcrCR-242 1 FL FJ829646 AcrCR-243 1 FL FJ829647 AcrCR-244 2 FL FJ829648 AcrCR-245 1 FL FJ829649 AcrCR-246 1 FL FJ829650 AcrCR-247 1 FL FJ829651 AcrCR-248 1 FL FJ829652 AcrCR-249 1 GA FJ829653 AcrCR-250 1 GA FJ829654 AcrCR-251 3 GA FJ829655 AcrCR-252 1 GA FJ829656 AcrCR-253 1 GA FJ829657 AcrCR-254 1 LA FJ829658 AcrCR-255 3 LA FJ829659 AcrCR-256 2 LA,MS FJ829660 AcrCR-257 1 LA FJ829661 AcrCR-258 1 LA FJ829662 AcrCR-259 2 LA FJ829663 AcrCR-260 2 LA,MS FJ829664 AcrCR-261 1 LA FJ829665 AcrCR-262 3 LA,MS FJ829666 AcrCR-263 1 LA FJ829667 AcrCR-264 1 LA FJ829668 AcrCR-265 1 LA FJ829669 AcrCR-266 1 LA FJ829670 AcrCR-267 1 LA FJ829671

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 212 -

AcrCR-268 2 LA FJ829672 AcrCR-269 1 LA FJ829673 AcrCR-270 1 LA FJ829674 AcrCR-271 1 LA FJ829675 AcrCR-272 1 LA FJ829676 AcrCR-273 1 LA FJ829677 AcrCR-274 1 LA FJ829678 AcrCR-275 1 LA FJ829679 AcrCR-276 1 LA FJ829680 AcrCR-277 1 LA FJ829681 AcrCR-278 3 LA,MS FJ829682 AcrCR-279 1 LA FJ829683 AcrCR-280 1 LA FJ829684 AcrCR-281 1 LA FJ829685 AcrCR-282 2 MS FJ829686 AcrCR-283 1 MS FJ829687 AcrCR-284 1 MS FJ829688 AcrCR-285 1 MS FJ829689 AcrCR-286 1 MS FJ829690 AcrCR-287 1 MS FJ829691 AcrCR-288 1 MS FJ829692 AcrCR-289 1 MS FJ829693 AcrCR-290 1 MS FJ829694 AcrCR-291 1 MS FJ829695 AcrCR-292 1 MS FJ829696 AcrCR-293 1 MS FJ829697 AcrCR-294 1 MS FJ829698 AcrCR-295 1 MS FJ829699 AcrCR-296 1 MS FJ829700 AcrCR-297 1 MS FJ829701 AcrCR-298 1 MS FJ829702 AcrCR-299 1 MS FJ829703 AcrCR-300 1 MS FJ829704 AcrCR-301 1 MS FJ829705 AcrCR-302 1 NC FJ829706 AcrCR-303 1 NC FJ829707 AcrCR-304 1 NC FJ829708 AcrCR-305 1 NC FJ829709 AcrCR-306 2 NC FJ829710 AcrCR-307 1 NC FJ829711 AcrCR-308 1 NC FJ829712 AcrCR-309 1 NC FJ829713 AcrCR-310 1 NC FJ829714 AcrCR-311 8 NC FJ829715 AcrCR-312 1 NC FJ829716 AcrCR-313 2 NC FJ829717

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission. - 213 -

AcrCR-314 3 NC FJ829718 AcrCR-315 1 NC FJ829719 AcrCR-316 3 NC FJ829720 AcrCR-317 1 NC FJ829721 AcrCR-318 1 NC FJ829722 AcrCR-319 1 SC FJ829723 AcrCR-320 1 SC FJ829724 AcrCR-321 1 SC FJ829725 AcrCR-322 1 SC FJ829726 AcrCR-323 1 SC FJ829727 AcrCR-324 1 SC FJ829728 AcrCR-325 1 SC FJ829729 AcrCR-326 1 SC FJ829730 AcrCR-327 1 SC FJ829731 AcrCR-328 1 SC FJ829732 AcrCR-329 1 SC FJ829733 AcrCR-330 1 SC FJ829734 AcrCR-331 1 SC FJ829735 AcrCR-332 1 SC FJ829736 AcrCR-333 1 SC FJ829737 AcrCR-334 1 SC FJ829738 AcrCR-335 1 SC FJ829739 AcrCR-336 1 SC FJ829740 AcrCR-337 1 SC FJ829741

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.