the population ecology, molecular ecology, and ... 2016.pdf · the population ecology, molecular...

166
The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) A dissertation presented to the faculty of the College of Arts and Sciences of Ohio University In partial fulfillment of the requirements for the degree Doctor of Philosophy Paul E. Converse August 2016 © 2016 Paul E. Converse. All Rights Reserved.

Upload: others

Post on 19-Jan-2020

23 views

Category:

Documents


3 download

TRANSCRIPT

Page 1: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback

Terrapin (Malaclemys terrapin)

A dissertation presented to

the faculty of

the College of Arts and Sciences of Ohio University

In partial fulfillment

of the requirements for the degree

Doctor of Philosophy

Paul E. Converse

August 2016

© 2016 Paul E. Converse. All Rights Reserved.

Page 2: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

2

This dissertation titled

The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback

Terrapin (Malaclemys terrapin)

by

PAUL E. CONVERSE

has been approved for

the Department of Biological Sciences

and the College of Arts and Sciences by

Willem M. Roosenburg

Professor of Biology

Shawn R. Kuchta

Associate Professor of Biology

Robert Frank

Dean, College of Arts and Sciences

Page 3: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

3

ABSTRACT

CONVERSE, PAUL E., Ph.D., August 2016, Biological Sciences

The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback

Terrapin (Malaclemys terrapin)

Director of Dissertation: Willem M. Roosenburg and Shawn R. Kuchta

The Diamondback Terrapin (Malaclmeys terrapin) is a turtle found in brackish

water habitats along the Gulf and Atlantic coastlines of the North American continent.

Historically, terrapins have had a complex relationship with humans, including over-

harvesting, habitat loss and degradation, and translocations. Furthermore, ecological and

molecular studies of terrapins yield conflicting results with respect to terrapin population

biology. Resolving these conflicts is integral to understanding how past human activities

have influenced the contemporary distribution and abundance of terrapin genetic

diversity. In Chapter 1, I surveyed the terrapin literature and identified incongruences

between ecological and molecular studies. I finish Chapter 1 with recommendations for

future molecular and ecological terrapin studies. In Chapter 2, I demarcated

metapopulation structure and quantified gene flow between populations in Chesapeake

Bay. I detected four populations with weak structure, high admixture, high genetic

diversity, and genetic signatures of anthropogenic translocation. In Chapter 3, I quantified

the effective population sizes of the Chesapeake Bay populations, with both ecological

and molecular approaches. Using mark-recapture, Bayesian model testing, and

approximate Bayesian computation, I recovered incongruent results among methods. I

then used mark-recapture data to rule out spurious molecular estimates, finding that

Page 4: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

4

coalescent models more accurately estimate effective population size in Chesapeake Bay.

In Chapter 4, I used a range-wide dataset to locate major terrapin populations, quantified

historical and contemporary gene flow, and tested for bottlenecks. I used Bayesian model

testing and discriminant analysis of principal components (DAPC) as well as historical

terrapin literature to show the terrapin’s phylogeographic structure was best explained by

anthropogenic translocation events in the early 1900’s. In Chapter 5, I revisited a classic

terrapin phylogeographic hypothesis with approximate Bayesian computation and report

evidence of a younger divergence date between Gulf and Atlantic populations. This

younger divergence date implies the terrapin’s mitochondrial genome has a faster rate of

molecular evolution than previously estimated.

Page 5: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

5

DEDICATION

I dedicate this work to my wife, Michelle, who has always supported my aspirations and

probably won’t read this. Tugboat.

Page 6: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

6

ACKNOWLEDGMENTS

My advisers have shown an inexhaustible supply of patience and understanding,

and I owe much of my growth as a scientist to them. Thank you, Willem and Shawn, for

building my confidence, refining my skill sets, and polishing my writing. The guidance of

a caring adviser has no substitute.

I would also like to thank my committee members: Morgan Vis, Harvey Ballard,

and Matt White. Your input and ideas to improve my dissertation and manuscripts has

greatly improved the quality of my work.

I received funding and logistical support from several sources, and I could not

have completed my work without their generosity. The Ohio University Graduate Student

Senate funded my work through an Original Works Grant (Chapter 2), and the OHIO

Office of the Vice President for Research and Creative Activity funded my work via a

Student Enhancement Award (SEA; Chapters 3-4). The Maryland Department of Natural

Resources provided me with a research assistantship, allowing me to conduct more

research during the spring of 2013 The Ohio Center for Ecology and Evolutionary

Studies also provided me with a research assistantship which helped me publish Chapter

2 prior to defending my dissertation. The United States Army Corps of Engineers

provided logistical support and funding to the mark-recapture study on Poplar Island.

I wish to thank my collaborators on my manuscripts. Timothy King (U.S.

Geological Survey) provided me with the Chesapeake Bay microsatellite dataset

(Chapters 2,3) and helped tremendously with the publication process. Susanne Hauswaldt

further provided me a range-wide genetic dataset, which was integral to Chapters 4 and 5.

Page 7: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

7

Paula F.P. Henry and Michael Haramis provided great input and help on Chapter 2 of my

dissertation.

I could not have completed these projects without the many undergraduates that

have helped over the years. I especially want to thank ElizaBeth Clowes, Kyle Heckler,

and Renee Harding for all of their hard work and friendship. I also want to thank them for

dealing with me as a housemate.

I want to thank the Roosenburg family (Kate, Selinde, Dirk and Willem) for their

warmness and generosity. Each summer, they open their home to strangers, and somehow

always make it work out. I will miss poorly steering the boat to catch crab, and the

delicious meals that followed.

My fellow graduate students are some of the most reliable friends I’ve made, and

I owe them my sanity. Eric, Andrés, Maggie, Mike, Chris, and Jorge (and others!) helped

me by simply talking science (and whatnot) over beers. Cheers!

Last, I want to thank my family and wife, Michelle (to whom this work is

dedicated). You have supported me more than I deserve, and I cannot express the

gratitude I have for your love.

Page 8: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

8

TABLE OF CONTENTS

Page Abstract ............................................................................................................................... 3

Dedication ........................................................................................................................... 5

Acknowledgments ............................................................................................................... 6

List of Tables .................................................................................................................... 12

List of Figures ................................................................................................................... 13

Chapter 1: Introduction ..................................................................................................... 14

Phylogeographic Studies ............................................................................................... 16

Regional and Population Level Studies ........................................................................ 20

Atlantic Coast ............................................................................................................ 20

Gulf Coast ................................................................................................................. 23

Insights into Reproductive Ecology Provided by Genetic Studies ............................... 26

Discussion ..................................................................................................................... 28

Chapter 2: Spatiotemporal Analysis of Gene Flow in Chesapeake Bay Diamondback

Terrapins ........................................................................................................................... 35

Introduction ................................................................................................................ 35

Materials and Methods ................................................................................................ 38

Sampling Localities and Microsatellite Genotyping ............................................... 38

Population Structure and Genetic Diversity ............................................................. 39

Mutation Rate .......................................................................................................... 40

Effective Population Size ........................................................................................ 40

Historical Gene Flow ............................................................................................ 41

Page 9: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

9

Contemporary Gene Flow ....................................................................................... 42

Comparison of Historical and Contemporary Gene Flow ...................................... 43

Population Bottlenecks ............................................................................................. 43

Results ........................................................................................................................... 44

Population Structure and Genetic Diversity ............................................................. 44

Mutation Rate ............................................................................................................ 45

Effective Population Size .......................................................................................... 45

Historical Gene Flow ................................................................................................ 45

Contemporary Gene Flow ......................................................................................... 46

Comparison of Historical and Contemporary Gene Flow ........................................ 46

Population Bottlenecks ........................................................................................... 47

Discussion .................................................................................................................. 47

Chapter 3: Effective Population Size in Diamondback Terrapins: Incongruence Between

Molecular and Ecological Estimates ................................................................................ 63

Introduction ................................................................................................................... 63

Effective Population Size .......................................................................................... 63

Study System ............................................................................................................ 65

Methods ........................................................................................................................ 66

Molecular Data .......................................................................................................... 66

Ecological Approaches ............................................................................................. 69

Results ........................................................................................................................... 72

Molecular Estimates of Ne ........................................................................................ 72

Ecological Estimates of Ne ........................................................................................ 73

Page 10: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

10

Discussion .................................................................................................................... 73

Molecular Approaches .............................................................................................. 74

Ecological Approaches ............................................................................................. 77

Incongruence Between Molecular and Ecological Estimates of Ne .......................... 78

Congruence between Ecological and Molecular Estimates ...................................... 79

Effective Population Size in Chelonians ................................................................... 80

Chapter 4: Turtle Soup, Prohibition, and the Population Genetic Structure of

Diamondback Terrapins .................................................................................................... 95

Introduction ................................................................................................................... 95

Methods ........................................................................................................................ 98

Sampling and Population Structure .......................................................................... 98

Historical and Contemporary Gene Flow ................................................................. 99

Temporal Changes in Gene Flow ........................................................................... 100

Testing Coastal Population Structure ...................................................................... 101

Bottlenecks .............................................................................................................. 101

Results ......................................................................................................................... 102

Population Structure — STRUCTURE .................................................................. 102

Population Structure — DAPC ............................................................................... 102

Historical and Contemporary Gene Flow ............................................................... 103

Models of Population Structure .............................................................................. 104

Bottlenecks .............................................................................................................. 104

Discussion ................................................................................................................... 105

Chapter 5: Splitting Diamonds: Dating Phylogeographic Divergence in the Diamondback

Terrapin Using Approximate Bayesian Computation ..................................................... 121

Page 11: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

11

Introduction ................................................................................................................. 121

Materials ..................................................................................................................... 123

Sample Collection and Microsatellite Preparation ................................................. 123

Coastal Population Structure ................................................................................... 124

Divergence Dating — Microsatellite Modeling ..................................................... 124

Divergence Dating — Range Scenarios and Priors ................................................ 125

Divergence Dating — Scenario Confidence and Scenario Checking ..................... 126

Results ......................................................................................................................... 127

Coastal Population Structure ................................................................................... 127

Divergence Dating and Effective Population Size .................................................. 127

Microsatellite Mutation ........................................................................................... 128

Discussion ................................................................................................................... 129

Literature Cited ............................................................................................................... 137

Page 12: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

12

LIST OF TABLES

Page

Table 2.1: Sampling locations around Chesapeake Bay……………………………….54

Table 2.2: Summary of ABC priors for mutation estimation .........................................56

Table 2.3: Population genetic metrics for loci…………………………………………58

Table 2.4: Dest results for STRUCTURE clusters .............................................................59

Table 3.1: Sample of statistics to estimate Ne .................................................................81

Table 3.2: Priors for DIYABC analyses ..........................................................................82

Table 3.3:Results from MIGRATE and Bézier scores .....................................................83

Table 3.4: Ne estimates from molecular data ..................................................................85

Table 3.5: Mark-recapture results 2006—2105 ..............................................................87

Table 3.6: POPAN models and results for Nc estimates .................................................88

Table 3.7: Ne estimates from mark-recapture data and non-Fisherian sex-ratios ...........89

Table 3.8: Summary of molecular software features .....................................................90

Table 4.1: Historical gene flow (M) estimates ..............................................................110

Table 4.2: Contemporary gene flow (m) estimates) .....................................................110

Table 4.3: Temporal changes in gene flow (∆m) ..........................................................111

Table 4.4: BOTTLENECK results ..................................................................................112

Table 5.1: Scenario posterior probabilities and divergence date estimates ..................133

Table 5.2: Types 1 and 2 errors for scenario selection .................................................134

Page 13: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

13

LIST OF FIGURES

Page Figure 1.1: Map of locations for previous terrapin genetic studies. ...............................34

Figure 2.1: STRUCTURE results and sampling localities in Chesapeake Bay. ...............60

Figure 2.2: ABC posterior estimates for Gmu loci mutation ..........................................61

Figure 2.3: Historical and contemporary gene flow and ∆m estimates ..........................62

Figure 3.1: Chesapeake Bay and Poplar Island sampling localities ...............................91

Figure 3.2: Models from MIGRATE for population structure. ........................................92

Figure 3.3: Schematic for model selection in MIGRATE. ...............................................93

Figure 3.4: Comparisons of Ne by molecular software ...................................................94

Figure 4.1: Range wide STRUCTURE results and sampling localities. ........................114

Figure 4.2: Depiction of the models tested in MIGRATE and Bézier scores ................115

Figure 4.3: BIC results for downstream DAPC cluster analyses. .................................116

Figure 4.4: DAPC results for k = 6 ...............................................................................117

Figure 4.5: Membership probabilities for individuals under k = 6 ...............................118

Figure 4.6: DAPC results for k = 12 .............................................................................119

Figure 4.7: Membership probabilities for individuals under k = 12 .............................120

Figure 5.1: STRUCTURE results and sampling localities .............................................135

Figure 5.2: Priors and results from DIYABC analyses ..................................................136

Page 14: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

14

CHAPTER 1: INTRODUCTION

The Diamondback Terrapin (Malaclemys terrapin) is one of the most charismatic

and recognizable turtle species in North America, populating brackish water habitats

along the Gulf and Atlantic coasts from Corpus Christi, TX to Cape Cod, MA, as well as

a small disjunct population in Bermuda. Historically, the terrapin has been divided into

seven subspecies: the Texas Diamondback Terrapin (Malaclemys terrapin littoralis), the

Mississippi Diamondback Terrapin (Malaclemys t. pileata), the Ornate Diamondback

Terrapin (M. t. macrospilota), the Mangrove Diamondback Terrapin (M. t.

rhizophorarum), the Eastern Florida Diamondback Terrapin (M. t. tequesta), the Carolina

Diamondback Terrapin (M. t. centrata), and the Northern Diamondback Terrapin (M. t.

terrapin). All subspecies were delimited based on regional variation in size, color, and

shape. The recognition of seven morphologically variable subspecies suggests substantial

amounts of phylogeographic and population genetic structure should be present within

the terrapin.

The natural history and ecology of the terrapin added weight to this thesis.

Ecological studies show that terrapins generally exhibit low vagility, reside in small

home ranges, and demonstrate high nest site philopatry (Auger 1989; Lovich and

Gibbons 1990; Roosenburg 1996; Spivey 1998; Roosenburg et al. 1999). Indeed,

Diamondback Terrapins have been recaptured in the same creek or river system for

decades (Gibbons et al. 2001). On the other hand, not all females demonstrate perfect

nest site fidelity, and some individuals are known to disperse long distances (8 km;

Sheridan et al. 2010). Moreover, the dynamics of juvenile dispersal are largely unknown.

Page 15: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

15

Thus, the ecological data are mixed in their predictions on levels of expected population

genetic structure.

Unfortunately, during the 19th and 20th Centuries, before patterns of terrapin

diversity were well studied, populations along the East Coast underwent severe

demographic declines as a consequence of demand for terrapin meat (Coker 1906). The

result was regional and local extirpations of terrapin populations. Moreover, as natural

populations declined, government and private terrapin farms were created to meet the

demand for their meat (Barney 1924; Hildebrand and Hatsel 1926; Hildebrand 1929; see

Chapter 13). Farm-raised terrapins were collected from poorly documented locations,

often locally but also from other parts of the terrapin’s range (Coker 1920). For example,

as terrapin populations in Chesapeake Bay diminished, terrapins from North Carolina

were translocated to replenish the dwindling stock (Coker 1920). Terrapins from South

Carolina were moved to North Carolina and into Chesapeake Bay (Coker 1920).

Additionally, Gulf terrapins were also translocated into south Atlantic populations (Coker

1920; Hildebrand 1933). Indeed, terrapins were moved around their range with such

frequency that Coker (1920) quipped “Had these terrapin carried hand-bags, they might

have displayed an array of hotel stickers to shame the traveler returned from Europe.”

When demand for terrapin meat subsided, farms purportedly simply released their

terrapins. The ability of released terrapins to interbreed with local populations and the

numbers and success of the translocated terrapins remains unknown.

Thus, the terrapin has a complicated history with humans, which creates significant

challenges to understanding the patterns of genetic variation. Most genetic work has

Page 16: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

16

focused primarily on conservation needs and management strategies. I survey studies of

genetic variation in the Diamondback Terrapin. Specifically, I address the terrapin’s

phylogeographic and population genetic structure, and the abundance and distribution of

terrapin genetic diversity, examining the relationship between terrapin genetic studies and

ecological findings. I end with recommendations for future genetic work.

Phylogeographic Studies

The Diamondback Terrapin has extensive phenotypic variation among its seven

putative subspecies. Accordingly, there was the expectation of underlying high levels of

genetic structure across its range. Lamb and Avise (1992) conducted the first analysis of

genetic structure throughout the terrapin’s range using mitochondrial DNA (mtDNA) and

18 restriction enzymes. They recovered 74 restriction fragments from 53 terrapins, with

six haplotypes. Compared to other vertebrates, this level of divergence was unusually low

(uncorrected p = 0.001%), and geographic structure was weak, with only one

phylogeographic break near Cape Canaveral, FL (referred to as the “Gulf/Atlantic divide”

in the literature; Figure 1.1).

In contrast with taxonomy, Lamb and Avise (1992) did not recover any of the

subspecies as diagnosable entities with molecular data. However, the genetic divide near

Cape Canaveral did correspond with a noticeable change in terrapin morphology.

Terrapins to the north lack knobs on their medial keels while terrapins in the south and in

the Gulf possess knobby keels.

One possible cause of the discordance between subspecific designations based on

morphology and terrapin phylogeography may be that terrapins exhibit low levels of

Page 17: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

17

mtDNA variation. Previous research suggested that Chelonians exhibit low rates of

nuclear and mitochondrial evolution relative to other vertebrates (Avise et al. 1992;

Broham 2002; Shaffer et al. 2013). The slow-down in terrapin mtDNA evolutionary rate

was inferred from concordant phylogeographic breaks among taxa (Lamb and Avise

1992; Avise et al. 1992). For instance, the Eastern Oyster (Crassostrea virginica) and

Seaside Sparrow (Ammodramus maritimus) exhibit a phylogeographic break in the Gulf,

while the Horseshoe Crab (Limulus polyphemus) exhibits a phylogeographic break near

Cape Canaveral, FL, suggesting these taxa share a common vicariant history with the

terrapin (Avise et al. 1992). However, relative to these other taxa, the terrapin

demonstrates a paucity of mtDNA variation and thus would require a much slower rate of

mtDNA evolution relative to other vertebrates (approximately 14-fold) if they shared this

vicariant event (Avise et al. 1992). However, Avise et al. (1992) also noted that terrapins

might simply have been isolated near Cape Canaveral more recently.

Perhaps as a consequence of the low genetic variation found by Lamb and Avise

(1992), no further phylogeographic work was conducted on the terrapin for over a

decade, until revisited by Hauswaldt (2004) and Hauswaldt and Glenn (2005). Hauswaldt

(2004) surveyed the mtDNA control region, a rapidly evolving marker in most

vertebrates, across the terrapin’s range. She recovered 28 mtDNA haplotypes, with

unique haplotypes in Maryland, Texas, and Florida. The phylogeographic break found by

Lamb and Avise (1992) was recovered, and haplotypes for the Atlantic and Gulf Coast

formed separate clades. However, limited phylogenetic structure was recovered in both

Page 18: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

18

clades, and even the Atlantic/Gulf divide lacked strong support (bootstrap < 70%; Hillis

and Bull 1993).

In a parallel study, Hauswaldt and Glenn (2005) quantified genetic diversity

across the terrapin’s range using six nuclear microsatellites (TerpSH; Hauswaldt and

Glenn 2003). Population structure, population assignment, and bottleneck detection were

the primary foci of the work. Most sampling took place along the Atlantic coast (NY, NJ,

MD, NC, and three locations in SC), with additional sampling in the Gulf (FL and TX).

Across the terrapin’s range, they found high levels of heterozygosity (Ho; 0.70-0.87) and

large numbers of alleles at most loci (A; 7.5-13), except in Florida (Ho = 0.56; A = 6.5)

Using an analysis of molecular variance (AMOVA; Excoffier et al. 1992), Hauswaldt and

Glenn (2005) found that little variance could be explained by grouping all sampling

localities (16.8%), while partitioning populations into Gulf and Atlantic groups explained

the most, but still little, variance (25.1%). In addition, Hauswaldt and Glenn (2005)

demonstrated that individuals in some populations could be correctly assigned to their

population of origin using Bayesian assignment tests. Terrapins in the Gulf had higher

instances of correct assignment (71-100%) than populations in the Atlantic (10-75%).

Interestingly, terrapins along the Atlantic were more similar to Texas populations than

Florida populations, consistent with the observation that terrapin translocations occurred

during the early 20th Century. Surprisingly, the program BOTTLENECK (Piry et al. 1999)

found no evidence of demographic bottlenecks in any population, including populations

known to have been historically over-harvested.

Page 19: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

19

More recently, Hart et al. (2014) used 12 nuclear microsatellites (Gmu; King and

Julian 2004) to revisit phylogeographic structure throughout the terrapin’s range.

Populations in the North Atlantic, Gulf, and Florida showed low levels of heterozygosity

(0.32-0.47), while mid-Atlantic populations showed moderate levels (0.66-0.71). Genetic

clusters (populations) were inferred using the Bayesian clustering program STRUCTURE

(Pritchard et al. 2000). STRUCTURE operates by dividing individuals into clusters such

that conformity to Hardy-Weinberg equilibrium is maximized while linkage

disequilibrium is minimized. Hart et al. (2014) initially detected the presence of two

genetic clusters. The first consisted of terrapins from Atlantic populations while the

second consisted of terrapins from Florida. STRUCTURE plots indicated that populations

sampled from the Gulf (LA and TX) were admixed with samples from the Atlantic and

Florida. Within the Atlantic cluster, two subclusters were evident: Massachusetts vs. the

remaining Atlantic populations. Further analysis showed that Louisiana and Texas

formed a cluster, while Florida formed the final cluster, for a total of four clusters. Hart et

al. (2014) also found support for these clusters using neighbor-joining phenograms,

AMOVAs, and FST calculations. In contrast with previous studies, evidence of population

contraction was detected using BOTTLENECK, with each locality exhibiting some

signature of a bottleneck. Hart et al. (2014) concluded that precipitous drops in

population sizes had occurred.

Page 20: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

20

Regional and Population Level Studies

Atlantic Coast

While phylogeographic studies did not recover large genetic discontinuities or

incipient species, the investigation of patterns of genetic variation within and among

populations has been an ongoing research interest. Such studies aim to reveal smaller

scale patterns of differentiation, quantify directions and levels of gene flow, probe

demographic history, and direct conservation and management.

Terrapins were historically harvested at unsustainable levels (Garber 1990), and

the genetic impacts of this exploitation are a common thread running through terrapin

research. To date, a range-wide, coordinated conservation plan for terrapins is not in

place; rather, conservation efforts are left to individual states, with inconsistent

enforcement (Glenos 2013; Hart et al. 2014). To investigate the problem of illegal

harvests, Lester (2007) conducted a covert study to identify the source populations of

terrapins sold at the Fulton Fish Market in New York City. She obtained blood samples

from 63 terrapins in 2004, and, using assignment tests and 12 Gmu microsatellites,

determined that terrapins had been collected along much of the East Coast, from South

Carolina to New York. The majority originated from Chesapeake Bay (34/63), especially

from Sandy Island Cove, MD. Maryland closed their legal terrapin fishery in 2007 due to

its negative impacts on local populations (Roosenburg et al. 2008). How illegal markets

currently affect terrapin populations remains unclear.

Many studies have examined genetic variation within and among sets of

populations, with an eye towards documenting population structure and habitat use (e.g.,

Page 21: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

21

Callens et al. 2011). Elucidating how habitat features impact patterns of gene flow has

important implications for metapopulation dynamics and conservation planning. In a

creative study, Sheridan (2010) used six Gmu microsatellite markers and computational

modeling to compare and contrast barriers to dispersal in Barnegat Bay, NJ. First, she

measured genetic differentiation among sampling localities using FST and G’ST (Fst for

multiple alleles) statistics. She then estimated gene flow among populations under a

diversity of models. Simple models quantified genetic differentiation as a function of

straight-line and shoreline distances. Complex models included landscape features, and

assigned costs to traverse them. Open, deep and shallow waters, emergent wetlands,

development, roads, and riprap (rocks used to armor shorelines) were investigated for

their propensity to facilitate or impede gene flow. Sheridan (2010) found that open water

and developed landscapes were the biggest impediments to gene flow, while shallow

water and emergent wetlands facilitated genetic exchange. These results are consistent

with ecological studies of terrapin movement (Roosenburg et al. 1999).

Populations located at the edge of a species range commonly have smaller

population sizes and lower levels of genetic diversity than more centrally located

populations (Kirkpatrick and Barton 1997). McCafferty et al. (2013) investigated this

phenomenon in terrapins by examining populations near the northern edge of the range in

Massachusetts. Populations in this region are at the terminus of the terrapin’s range, but

also may have undergone bottlenecks as a consequence harvesting. To investigate this,

McCafferty et al. (2013) combined nuclear sequence data (class 1 major

histocompatibility complex, MHC1) with six Gmu microsatellite loci. Despite their study

Page 22: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

22

populations being at the northern edge of the terrapin’s range, McCafferty et al. (2013)

found moderate levels of heterozygosity (0.62-0.71), contrary to the lower levels

typically associated with range margins and population bottlenecks. Using STRUCTURE,

microsatellites demarcated two populations, while Fst statistics documented low to

moderate levels of population differentiation (0.03-0.11). The first STRUCTURE cluster

spanned from Wellfleet Harbor and Sandy Neck, MA, while samples in Sippican Harbor,

MA constituted a second cluster. In contrast with microsatellites, no variation was

recovered at the nuclear locus, MCH1. McCafferty et al. (2013) proposed this result was

the consequence of the locus operating under purifying selection, although recent range

expansion or low effective population size could lead to the same lack of variation.

Translocation may also affect population structure in this region. In 1972, approximately

100 terrapins from New Jersey were released in Buzzards Bay (Lazell 1976).

Not all population level studies have recovered variation among populations. In

the Hudson River of New York, Piermont Marsh harbors a terrapin population that is

small and broadly distributed, with few high quality nesting sites. This population may be

isolated from neighboring populations, and Wiktor et al. (2001) hypothesized that

terrapins in Piermont Marsh may be genetically differentiated from other populations in

the region. To test this hypothesis, they compared terrapins from Piermont Marsh with

terrapins from Rhode Island, Connecticut, and New Jersey using two nuclear inter-simple

sequence repeat (ISSR) loci. In contrast with their predictions, they found little genetic

differentiation between Piermont Marsh and other populations in the North Atlantic.

Wiktor et al. (2001) concluded the Piermont Marsh population must be connected to

Page 23: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

23

others via high levels of gene flow, most likely from populations in the Hudson River. An

alternative explanation is that the Piermont Marsh population has been founded relatively

recently. That is, it could in fact be currently isolated demographically, yet this isolation

may not yet have left a genetic signature.

Most recently, Converse et al. (2015) quantified historical and contemporary gene

flow among terrapin populations in Chesapeake Bay. Using 12 Gmu microsatellite loci

and STRUCTURE, they found four terrapin populations that were weakly structured, yet

harbored moderate to high levels of heterozygosity (0.69-0.79). Using MIGRATE (Beerli

2008) and BAYESASS (Wilson and Rannala 2003) to estimate historical and

contemporary gene flow, respectively, Converse et al. (2015) showed large increases in

gene flow into Chesapeake Bay, as well as large increases between the Patuxent River

and Kent Island, potentially due to terrapin translocation events. By contrast, levels of

gene flow decreased over time among most populations within the Bay. They attributed

this latter pattern to habitat loss and the impacts of a large crabbing industry, which

negatively impacts terrapin dispersal, especially of juveniles and males (Roosenburg et

al. 1997; Roosenburg and Green 2000).

Gulf Coast

Until recently, populations in the Gulf were understudied by molecular

ecologists. The work of Hart (2005, 2014) and Hauswaldt and Glenn (2005) suggested

that Gulf Coast populations contained lower levels of genetic diversity than populations

in the mid-Atlantic, but both studies included limited sampling in the Gulf region.

Coleman (2011) was the first to include a wide sampling of Gulf populations. Using 12

Page 24: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

24

Gmu microsatellites and STRUCTURE, he identified three clusters of populations:

Florida, South Carolina, and the Gulf (TX, LA, AL). STRUCTURE suggested little

admixture among clusters. While earlier studies found mixed support for population

bottlenecks, Coleman (2011) found strong support for bottlenecks in populations in the

Gulf, Florida, and Atlantic using M-ratios (all M < 0.40). Moderate levels of

heterozygosity (0.50-0.65) and significant deviations from Hardy-Weinberg equilibrium

in the Alabama and South Carolina populations provided support for genetic bottlenecks.

His findings differed from Hauswaldt and Glenn (2005), who found much higher levels

of heterozygosity (0.56 - 0.80) and no evidence for bottlenecks in Gulf populations.

Work on Gulf Coast terrapins continued with Glenos (2013) whose sampling was

most dense in Texas, but included Louisiana and Alabama. She did not find significant

differentiation between the two sampling localities within Texas, suggesting terrapins

along the Texas coast constitute a single population. On the other hand, she found

terrapins from Alabama and Louisiana to be genetically separated from the Texas

population. Consistent with Coleman (2011), her 12 Gmu microsatellites exhibited low

levels of heterozygosity (0.39-0.52), few private alleles, and low numbers of alleles (3.0-

5.5) in all populations.

The value of diversity studies is always highlighted by environmental disasters,

such as the Deepwater Horizon oil spill in 2010, which deeply impacted coastal ecology

in Louisiana. Drabeck et al. (2014) studied the effects of the oil spill on terrapin

populations. Using four TerpSH and 12 Gmu microsatellites, they found no meaningful

population structure among Louisiana terrapins, nor did they detect population

Page 25: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

25

contractions using the program BOTTLENECK, in contrast with Coleman (2011) and Hart

et al. (2014). However, consistent with Lamb and Avise (1992) and Hart et al. (2014),

Drabeck et al. (2014) found Florida, the Gulf, and Atlantic terrapin populations to be

diagnosable entities. AMOVAs showed that most genetic variance (7.47%) was

explained by combining Gulf and Atlantic terrapin populations while excluding Florida

populations. Interestingly, similar to Hauswaldt and Glenn (2005), they also found

evidence for terrapin translocations during the 20th Century, as Gulf Coast terrapins were

more similar to coastal Atlantic populations than adjacent populations in the Florida

Keys. This finding may be due to the undesirable traits of Florida terrapins. Florida

terrapins were considered insipid in taste and too small to sell or hybridize; rather,

terrapins from Texas were preferred for shipment to the Atlantic seaboard (Hay 1917;

Hildebrand 1933).

Further work on Louisiana terrapins was conducted by Petre et al. (2015).

Twenty-six sampling localities (analyzed as eight sites) along the Louisiana coastline

were examined for population structure, genetic diversity, population connectivity, and

bottlenecks, using 13 Gmu microsatellites. While STRUCTURE did not detect any

meaningful population structure, isolation-by-distance was detected along Louisiana’s

coastline and MIGRATE-n found a stepping stone model best explained patterns of gene

flow. Petre et al. (2015) found fairly high levels of heterozygosity (0.74-0.77) and allele

number (5.8-10.4) among populations. Despite the high levels of genetic diversity,

BOTTLENECK found that two of the eight sites exhibited the genetic signature of

population contraction.

Page 26: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

26

In summary, while many ecological studies document high levels of population

structure, phylogeographic and regional genetic studies have found more limited genetic

variation among populations. This contrast could be due to genetic and ecological studies

quantifying structure on different time scales, with ecological studies documenting real

time population structure that is yet to manifest itself in the genome. It remains possible

that the terrapin recently underwent a rapid range expansion across much of its range,

resulting in limited phylogeographic structure. Terrapin populations along the coasts may

also have experienced sporadic episodes of increased gene flow, including human-

mediated gene flow.

Insights into Reproductive Ecology Provided by Genetic Studies

Large documented declines in terrapin populations during the early 20th Century

were a major impetus to study the patterns of terrapin genetic diversity. By contrast,

studies of terrapin reproductive biology received less attention. Two themes within

reproductive biology have been studied by molecular ecologists: natal philopatry and

multiple paternity. Natal philopatry is the propensity for offspring to return to their birth

location to nest. Whether or not emerging terrapin hatchlings exhibit natal philopatry was

unknown until recently, though it had been documented in the Mississippi Map Turtle

(Graptemys pseudogeographica kohni), a closely related species (Freedberg et al. 2005).

High levels of natal philopatry are important for studies of genetic differentiation because

they represent a mechanism by which genetic structure among populations can evolve. If

mothers and their daughters return to the same nesting locality each season, genetic

evidence should complement ecological findings of nest site fidelity by mothers. In

Page 27: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

27

Barnegat Bay NJ, Sheridan et al. (2010) documented natal philopatry in terrapins using

six Gmu microsatellites and spatial autocorrelation analysis. They calculated genetic

correlations among terrapins, and then surveyed for relationships with geographic

distance. The only demographic group with a positive genetic autocorrelation was nesting

females (at 0-50 m), indicating they exhibit spatial structure that could result from nest-

site philopatry. Sheridan et al. (2010) also found that many juvenile males were transient

at their study site. High levels of juvenile movement may be partly responsible for the

lack of population genetic structure found in terrapins.

A clutch of Diamondback Terrapin hatchlings share a mother, but not necessarily

a father. Multiple paternity in terrapins was not well understood until advances in genetic

techniques allowed researchers to estimate paternity accurately but highly suspect as

terrapins can store sperm for up to four years (Hildebrand and Hatsel 1926). The first

study of multiple paternity in terrapins was conducted by Hauswaldt (2004), who

surveyed six TerpSH microsatellite in 439 hatchlings from 26 mothers in Oyster Bay,

NY. She found a majority of nests (82%) were sired by one male. This result was

somewhat surprising, and suggested multiple paternity in the terrapin was uncommon

relative to other Chelonians (e.g., Galbraith et al. 1993; Valenzuela 2000; Ireland et al.

2003).

Like any trait, multiple paternity can vary by region or population, or even over

time in the same location. Following Hauswaldt (2004), Sheridan (2010) surveyed six

Gmu microsatellites in 1,608 hatchlings and 1,558 adults from four sites in Barnegat Bay,

NJ and Poplar Island in Chesapeake Bay. She recorded higher levels of multiple paternity

Page 28: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

28

than did Hauswaldt (2004), with a mean of 29%. However, the occurrence of multiple

paternity varied widely among nesting locations, from 12.5 to 45.7%. Moreover, this

variation was correlated with population sex ratio: populations with female-biased sex

ratios (4.7:1) demonstrated the highest levels of multiple paternity, while populations

with extreme female-biased sex ratios (9:1) had lower levels.

Discussion

Studies of genetic variation have taught us much about terrapin biology. At the

broadest scale, range-wide phylogeographic analyses by Lamb and Avise (1992),

Hauswaldt (2004), Hauswaldt and Glenn (2005), and Hart et al. (2014) have shown that

genetic structure across the range of the terrapin is limited, and do not support the

existence of seven subspecies, at least if one views subspecies as incipient evolutionary

lineages. Thus, the subspecific designations in terrapins are in need of re-evaluation.

Genetic studies of the terrapin are made more difficult by limited sequence

variation and the specter of frequent human translocations. In general, mtDNA exhibits

higher levels of structure than nuclear markers (microsatellites excluded), making

mtDNA useful for phylogeography. However, in Chelonians, mtDNA is thought to

mutate more slowly than in other vertebrates (Avise et al. 1992; Bromham 2002; Shaffer

et al. 2013), resulting in less structure and weaker inference. Ultimately, many nuclear

loci from next generation sequencing resources are needed to elucidate the evolutionary

history of the terrapin.

At the range-wide and regional scales, a lack of genetic structure may be a

consequence of the terrapin rapidly expanding its range from a refugium. Because the

Page 29: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

29

terrapin’s range closely follows the Gulf and Atlantic coasts, it is essentially structured as

a linear stepping stone, which should produce isolation-by-distance. STRUCTURE

assumes no isolation-by-distance (Pritchard 2010), and yet isolation-by-distance is

present at regional and range-wide scales in terrapins (Hauswaldt and Glenn 2005; Hart

et al. 2014). Violation of the assumptions of FST statistics (equal population sizes, island

migration model, etc.) may also obfuscate population structure estimates. For this reason,

it is important that future work take full advantage of the latest advances in population

genetics, especially coalescent theory, which represents a powerful stochastic model for

evaluating demographic history (Wakeley 2008). For example, Bayesian skyline plots

could be used to examine variation in effective population sizes through time.

Alternatively, historical events could also be investigated using Approximate Bayesian

Computations, which allow complex models to be constructed and tested against genetic

data (Csilléry et al. 2010). For instance, models including admixture, population

divergence dates, population sizes, and translocation could be assembled and tested

against one another.

Among population genetic studies, a shared finding is that terrapin genetic

diversity is the lowest in southern populations, especially in Florida. This finding is in

contrast with many taxa in the Eastern U.S., especially terrestrial taxa, which display the

highest levels of genetic diversity in the south and the lowest levels in the north (Avise

2000; Soltis et al. 2006). In terrapins, the legacy of terrapin harvests may contribute to

lowered genetic diversity. However, historical factors likely contributed as well. For

example, if the terrapin had a Pleistocene refugium along the Atlantic coast and only

Page 30: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

30

recently (within the last several thousand years) expanded into the Gulf, Gulf populations

would be expected to exhibit low levels of genetic diversity.

By contrast, genetic diversity is highest in the mid-Atlantic region, possibly due to

translocations into the region. Translocation is supported by Bayesian population

assignment tests, which often assign Atlantic terrapins to Gulf populations (Hauswaldt

and Glenn 2005). In addition, Drabeck et al. (2014) found that Atlantic populations were

genetically more similar to Gulf populations than neighboring populations. Alternatively,

the mid-Atlantic could be the origin of the terrapin’s historical range. If the terrapin

originated in this portion of its range, genetic diversity is expected to be higher there

while peripheral populations (the North Atlantic, Florida, and the Gulf) are expected to

exhibit lower genetic diversity.

A common finding among genetic studies is that Florida populations stand out as

distinctive. Hart el al. (2014), Drabeck et al. (2014), and Hauswaldt and Glenn (2005) all

found Florida terrapins to be genetically differentiated from neighboring populations in

the Atlantic and Gulf. The cause of this differentiation is not entirely clear, but terrapin

populations in Florida exhibit much lower levels of heterozygosity (0.363; Hart et al.

2013) and fewer private alleles than populations in other regions, consistent with a severe

genetic bottleneck, inbreeding, or small population sizes. One possible cause is the

Suwanne Seaway, which separated peninsular Florida from the North American continent

by sea level rise during the Miocene and Pliocene (Felder and Staton 1994). This Seaway

would have isolated terrapin populations in Florida, while facilitating gene flow between

Page 31: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

31

Gulf and Atlantic populations. The Suwanne Seaway could also account for the

Gulf/Atlantic genetic divide first documented by Lamb and Avise (1992).

It is interesting to note that the specific microsatellite markers that terrapin studies

have used influence the recovered patterns of diversity. Two sets of microsatellite

primers are commonly used: TerpSH (Hauswaldt and Glenn 2003) and Gmu (King and

Julian 2004). These two sets of markers differ in mutation rate. Hauswaldt (2004)

estimated her loci to have an average mutation rate of 2.72 x 10-3 mutations-1 site-1

generation-1, while loci developed by King and Julian (2004) have a mean mutation rate

of 4.34 x 10-4 mutations site-1 generation-1 (Converse et al. 2015), approximately 10-fold

slower. Accordingly, studies that use the TerpSH microsatellites consistently recover

higher levels of diversity (Hauswaldt and Glenn 2005; Drabeck et al. 2014) than do

studies that use the Gmu microsatellites (Hart et al 2014; Drabeck et al. 2014). Inference

of terrapin phylogeographic structure is also affected by marker choice, with mtDNA

recovering two phylogeographic groups (the Gulf/Atlantic divide; Lamb and Avise 1992;

Hauswaldt 2004), and microsatellites recovering four population clusters (Gulf, Atlantic,

MA, FL; Hart et al. 2014). While incongruence between mitochondrial and nuclear loci is

not uncommon, terrapin studies have routinely recycled the same markers. The

development of new high-resolution genetic markers is sorely needed.

Evidence for population bottlenecks is highly variable among studies, even

though there is good historical evidence for demographic bottlenecks across the terrapin’s

range (Garber 1990). Louisiana terrapins are an excellent example of the discordance

among studies. Hart et al. (2014) and Coleman (2011) report bottlenecks in this region,

Page 32: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

32

Petre et al. (2015) found evidence that varied by site, and Drabeck et al. (2014) found no

evidence of a bottleneck in this region. In some cases, outbreeding due to terrapin

translocations among populations may have obfuscated genetic evidence of a bottleneck.

More work on population bottlenecks and their detection is warranted.

Finally, while many ecological studies have shown terrapins to exhibit high site

fidelity and low dispersal, which should promote the evolution of population genetic

structure, population genetic studies have consistently recovered low levels of genetic

structure. Several hypotheses have been put forth to explain this discrepancy between

ecology and genetics. Hauswaldt and Glenn (2005) hypothesized that terrapins form

mating aggregations in the early spring and disperse far from their home range to

participate. This behavior would homogenize local population structure such as generated

by nest site fidelity. Ecological work by Seigel (1980) and Butler (2002) support this

interpretation. However, work by Sheridan (2010) suggests that terrapins do not travel

long distances to form mating aggregations. Converse et al. (2015) hypothesize that

mating aggregations homogenize populations at the local level, while male-biased

dispersal homogenizes populations at larger spatial scales. Since most populations are

strongly female-biased, low levels of male migration can result in high levels of

admixture, and thereby reduce population structure. I recommend that future ecological

work focus on population sex ratios, juvenile dispersal, and adult dispersal in relation to

mating aggregations, as these elements of terrapin biology have the potential to

harmonize ecological and molecular findings.

Page 33: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

33

It is important that future genetic work also take full advantage of advances in

genomics to more accurately test models of population structure and population history.

Thankfully, a large volume of high quality loci is on the horizon, as scientists at the

University of Maryland have recently sequenced the entire mitochondrial and nuclear

genomes of the terrapin (Mihai Pop, Center for Bioinformatics and Computational

Biology, University of Maryland, pers. comm.). With a reference genome, a study could

now easily include tens of thousands of single nucleotide polymorphisms or

microsatellite loci. Hyper-variable sequence regions unique to terrapins also could be

identified and used. Indeed, it is now possible to sequence whole genomes from

throughout the range of the terrapin.

The Diamondback Terrapin has a complicated history, and it is important that

future studies incorporate the latest advances in genomics, computational resources, and

population genetic theory. Our understanding of the history of the terrapin is incomplete,

and a range-wide phylogeographic study that takes advantage of next generation

sequencing resources is needed to define management units, identify possible Pleistocene

refugia, and better document population history. Populations in Florida are of particular

interest as they are genetically distinct from both Gulf and Atlantic populations, perhaps

as a consequence of the Suwanne Seaway. On a finer scale, landscape genetic studies that

document contemporary movement patterns would inform conservation efforts. For

instance, models of gene flow that include electrical circuit theory (McRae and Beier

2007) could be used to investigate the effects of crab pots on population connectivity.

Additionally, studies of contemporary gene flow could include historical estimates, as

Page 34: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

34

these provide important context for interpreting patterns of connectivity (Epps et al.

2013; Converse et al. 2015). Finally, more studies on natal philopatry and juvenile

dispersal are needed, because these elements of terrapin biology will have an overarching

impact on population genetic structure. By taking advantage of the latest developments in

DNA sequencing and population genetic theory, and by conducting work on natural

history and ecology in parallel, our understanding of the evolutionary history of the

Diamondback Terrapin stands to become re detailed than ever.

Figure 1.1: Approximate sampling localities for range-wide, regional, and local

Diamondback Terrapin studies. Each noted locality may represent multiple individuals

sampled and/or multiple sampling sites within the same general area.

Page 35: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

35

CHAPTER 2: SPATIOTEMPORAL ANALYSIS OF GENE FLOW IN CHESAPEAKE

BAY DIAMONDACK TERRAPINS

Introduction

The current genetic structure among a set of populations is the product of

contemporary and historical processes, and distinguishing between the two is paramount

for effective population management (Epps et al. 2013). Around the world, the

fragmentation of habitats is a ubiquitous threat to biodiversity because it decreases

population connectivity (dispersal and gene flow) relative to historical levels, thereby

impacting metapopulation dynamics (Hanski and Gilpen 1997; Frankham et al. 2002).

Reductions in gene flow and small effective population size (Ne) caused by habitat

fragmentation diminish metapopulation viability by decreasing genetic diversity and

increasing inbreeding (Lande 1995; Templeton et al. 2001; Bowler and Benton 2005;

Banks et al. 2013; Barr et al. 2015). However, the extent to which habitat fragmentation

decreases population connectivity, however, is dependent upon the interaction between

landscape features and organismal dispersal behavior (Gu et al. 2002; Caizergues et al.

2003; Braunisch et al. 2010; Callens et al. 2011; Crispo et al. 2011). In many cases,

populations that are currently isolated by habitat fragmentation may not have been

isolated in the past (Newmark 2008; Chiucchi and Gibbs 2010; Epps et al. 2013;

Husemann et al. 2015). In contrast with habitat fragmentation, the anthropogenic

translocation of individuals between populations reduces genetic differentiation,

increases diversity within populations, and may obscure estimates of genetic connectivity

(Templeton et al. 1986; Moritz 1999; Weeks et al. 2011). Disentangling how historical

Page 36: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

36

and contemporary processes affect current patterns of genetic diversity is a formidable

challenge, but can be achieved by temporal sampling (Husemann et al. 2015), or by

separately estimating contemporary and historical processes (Chiucchi and Gibbs 2010;

Epps et al. 2013).

In this chapter, I examine population structure and connectivity in the

Diamondback Terrapin (Malaclemys terrapin). Terrapins inhabit North American coastal

and brackish waters, with a range that extends from Texas to Massachusetts (Ernst and

Barbour 1989). During the nineteenth and early twentieth centuries terrapins were

unsustainably harvested, resulting in severe population contractions and local extirpations

(Garber 1998; 1990). To help preserve dwindling populations and supplement terrapin

harvests, governmental and private entities constructed terrapin breeding farms (Coker

1906; Barney 1924; Hildebrand and Hatsel 1926; Hildebrand 1929). Terrapins from

Chesapeake Bay were the preferred variety for consumption (Hay 1917; Hildebrand

1929) and the demand for “Chesapeakes” resulted in terrapins from North Carolina and

possibly other populations to be imported into Chesapeake Bay terrapin farms (Coker

1920). Terrapin meat eventually fell out of favor, and as breeding farms closed terrapins

were reportedly released into local waters. The amount of admixture from these

translocated terrapins is unknown.

While terrapin harvesting in Maryland has been discontinued (Roosenburg et al.

2008), terrapins still face a myriad of threats, including mortality from boat strikes

(Roosenburg 1991, Cecala et al. 2008), drowning in crab and eel pots (Roosenburg et al.

1997; Radzio and Roosenburg 2005; Dorcas et al. 2007; Grosse et al. 2009), habitat loss

Page 37: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

37

and fragmentation (Roosenburg 1991; Wood and Herlands 1997), and predator

introductions (Feinberg and Burke 2003). As male terrapins are smaller and disperse

longer distances than do females (Sheridan 2010), they are particularly vulnerable to

dispersal related mortality. Terrapin populations in Chesapeake Bay exhibit highly

skewed sex ratios in favor of females (Roosenburg 1991; Roosenburg unpubl. data),

making successful male dispersal important for maintaining genetic connectivity.

The consequences of habitat fragmentation and increased mortality on

connectivity and population genetic structure are not entirely clear, however, and

ecological and molecular findings are discordant with respect to levels of connectivity

(Converse and Kuchta, 2017). Ecological data show terrapins reside in small home ranges

of 0.54 – 3.05 km2 (Spivey 1998; Butler 2002) and can remain in the same study site for

over a decade (Lovich and Gibbons 1990; Gibbons et al. 2001). Furthermore, ecological

data suggest terrapins form structured breeding assemblages, with females returning to

the same nesting beach each season (Auger 1989; Roosenburg 1994; Mitro 2003) and

hatchlings demonstrating natal philopatry (Sheridan et al. 2010). In contrast with these

studies, genetic studies indicate that terrapin populations are weakly differentiated (Hart

et al. 2014), with limited structure at both regional and local scales (Hauswaldt and Glenn

2005; Sheridan et al. 2010; Glenos 2013; Drabeck et al. 2014; Petre 2014).

The complex history terrapins share with humans in Chesapeake Bay makes it

important to quantify levels of population genetic structure, including a comparison of

contemporary and historical levels of connectivity. In this chapter, I report on a study of

metapopulation dynamics of the Diamondback Terrapin in Chesapeake Bay. Specifically,

Page 38: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

38

I estimate the following: i) the number of genetic populations in Chesapeake Bay; ii)

levels of genetic diversity within and among populations; iii) effective population sizes;

and iv) levels of contemporary and historical gene flow among populations. In addition, I

identify possible instances of terrapin translocation. By comparing historical and

contemporary levels of genetic connectivity, I examine the impact of habitat

fragmentation and population translocations on patterns of genetic variation, and help

resolve the discordance between ecological and molecular studies (Converse and Kuchta,

2017).

Materials and Methods

Sampling Localities and Microsatellite Genotyping

A total of 617 terrapins from 15 localities throughout Chesapeake Bay and nearby

coastal bays were captured between 2003-2005 (Figure 2.1; Table 2.1), a dataset

provided by Timothy L. King at the U.S. Geological Survey. Terrapins were captured

using fyke nets or collected in winter refugia during hibernation (Haramis et al. 2011)

and were marked with Passive Integrated Transponder tags to prevent resampling.

Twelve microsatellite loci were assayed that were developed for the bog turtle

(Glyptemys muhlenbergii) but also amplify in other Emydid turtles (King and Julian

2004). PCR chemistry and protocols for the loci are detailed in King and Julian (2004). I

used MICRO-CHECKER version 2.2.3 (Van Oosterhout et al. 2004), including 10,000

Monte Carlo simulations to test for the presence of null alleles and estimate 95%

confidence intervals. No evidence of null alleles was detected at any locus.

Page 39: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

39

Population Structure and Genetic Diversity

I used STRUCTURE version 2.3 (Pritchard et al. 2000) to infer the number of

genotypic clusters in the Chesapeake Bay region. STRUCTURE identifies populations by

maximizing conformity to Hardy-Weinberg equilibrium (HWE) while simultaneously

minimizing linkage disequilibrium within K user-defined clusters. I ran STRUCTURE

from K = 1 to K = 15 populations, with each value of K repeated ten times with randomly

generated starting seeds. Each Markov Chain Monte Carlo (MCMC) run consisted of

550,000 iterations, with the first 250,000 discarded as burn-in. I used the admixture

model, the correlated allele frequencies prior, the LOCprior, the LOCISPOP prior, fixed

λ, and inferred α. I used sampling localities (Figure 2.1) as priors for the LOCprior.

STRUCTURE results were collated and ∆K computed via the Evanno method (Evanno et

al. 2005) using STRUCTURE HARVESTER web version 0.6.94 (Earl and vonHoldt 2011).

Label switching and multimodality on preferred values of ∆K were addressed using

CLUMPP version 1.1.2 (Jakobsson and Rosenberg 2007), and the final results were

visualized using DISTRUCT version 1.1 (Rosenberg 2003). I repeated this procedure

within STRUCTURE clusters to detect substructure. I also partitioned genetic variance

using an analysis of molecular variance (AMOVA) in the software ARLEQUIN version

3.5.1.3 (Excoffier and Lischer 2010). Populations were partitioned by landscape features

(river, bay, and coast), sampling locality, and STRUCTURE clusters. Significance was

assessed using 1,000 permutations. I further estimated population differentiation by

quantifying Dest (Jost 2008) in the R package DEMEtics version 0.8-7 (Gerlach et al.

2010) between all STRUCTURE clusters. I determined significance and estimated 95%

Page 40: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

40

confidence intervals using 1,000 bootstrap replicates and Bonferroni correction (Dunn

1961). I used FSTAT version 2.9.3 (Goudet 1995) to estimate allelic richness, allele

count, and linkage disequilibrium for all sampling localities, and ARLEQUIN to estimate

heterozygosity and deviations from HWE.

Mutation Rate

Since coalescent estimates of historical gene flow and effective population sizes

are scaled by mutation rate, I estimated a mutation rate (µ) using approximate Bayesian

computation (ABC) in popABC version 1.0 (Lopes et al. 2009). Mutations were modeled

using the stepwise-mutation model (SMM; Kimura and Ohta 1978) and were measured in

mutations site-1 generation-1. Demographic parameters were estimated under the isolation-

migration model (Nielsen and Wakeley 2001; Hey and Nielsen 2004). Priors for this

analysis are summarized in Table 2.2. I modeled my mutation rate prior using a log-

normal distribution centered at 1 x 10-3 (SD = 0.5; Hedrick 1996; Whittaker et al. 2003).

Genetic tree topology was modeled under a uniform prior. I simulated 2,500,000 genetic

trees and ran the ABC-rejection algorithm with a tolerance of 0.0004, retaining the 1,000

closest simulated points. I did not run an ABC-regression analysis as some of the

summary statistics exhibited multicollinearity, violating the assumptions of local linear

regression (Beaumont et al. 2002). Following the rejection step, I estimated the mode,

2.5% quantile, and 97.5% quantile for µ in R.

Effective Population Size

I used MIGRATE version 3.6.5 (Beerli 2008) to jointly estimate θ (= 4Neµ) while

estimating M (see below) and used the mutation rate estimated by popABC to convert θ

Page 41: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

41

into Ne. I also estimated effective population sizes using ONeSAMP v. 1.2 (Tallmon et al.

2008), which employs ABC and eight common summary statistics (e.g. observed

heterozygosity, Wright’s FIS) to estimate Ne for a single population (Tallmon et al. 2008).

For these analyses, I ran each STRUCTURE population individually and set lower and

upper boundaries for Ne to 2 and 1,000, respectively.

Historical Gene Flow

I used MIGRATE to estimate gene flow levels in Chesapeake Bay prior to

European colonization (historical gene flow, M: proportion of migrants per generation,

scaled by mutation rate). Since MIGRATE operates in a coalescent framework, it

estimates gene flow over long periods of time, up to ~4Ne generations (thousands of

years) for larger populations (Beerli 2009). I used populations demarcated by

STRUCTURE as a priori population assignments in MIGRATE. To improve speed, I used

a Brownian motion model to approximate a stepwise- mutation model. Using slice

sampling, I ran four statically heated parallel chains (heated at 1.0, 1.5, 3.0 and

1,000,000) for 30,000,000 iterations, sampling every 3,000 iterations, and excluded

7,500,000 iterations as burn-in. MCMC estimates of M were modeled with a uniform

prior containing lower and upper boundaries of 0 and 2,000. FST values were used for

initial estimates of M. A full migration model was used, which facilitates comparisons

with gene flow estimates made in BAYESASS. I considered parameter estimates to be

accurate if an effective sample size (ESS) of 1,000 or greater was observed (Beerli, pers.

comm.)

Page 42: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

42

Contemporary Gene Flow

Contemporary rates of gene flow (m: proportion of migrants per generation) in

Chesapeake Bay were estimated using BAYESASS version 3.0 (Wilson and Rannala

2003). BAYESASS estimates all pairwise migration rates among populations. According

to Wilson and Rannala (2003), BAYESASS estimates gene flow “…over the last several

generations.” Following Chiucchi and Gibbs (2010), I assumed this to mean

approximately five generations. Using a generation time of 12 years (Roosenburg,

unpubl. data) BAYESASS is quantifying gene flow within the last 60 years or so, a time

period characterized by extensive anthropogenic influences, including habitat loss and

fragmentation. I used populations demarcated by STRUCTURE as a priori population

assignments. I ran 10 MCMC simulations (Faubet et al. 2007) with different starting

seeds for 20,000,000 iterations, sampling every 2,000 iterations; 10,000,000 iterations

were excluded as burn-in. Chain mixing delta parameters were adjusted in pilot runs to

maintain a MCMC state-change acceptance ratio of 20-40%, the empirically

recommended window (Rannala 2011). I diagnosed MCMC stationarity for each run in

TRACER version 1.5. (Rambaut and Drummond 2007) and used a Bayesian deviancy

measure (Spiegelhalter et al. 2002) to determine which run best fit the data with R

(Meirmans 2014). I took the starting seed from this best-fit run and ran a MCMC for

50,000,000 iterations, sampling every 2,000, with the first 20,000,000 iterations excluded

as burn-in. I visualized MCMC stationarity for this final run in TRACER. The ESS for all

parameters was >200.

Page 43: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

43

Comparison of Historical and Contemporary Gene Flow

I tested for a relationship between historical and contemporary gene flow by

conducting a Mantel test in the R package VEGAN version 2.2-1 (Oksanen et al. 2013)

using 100,000 permutations. To compare historical estimates of gene flow generated by

MIGRATE (M = mh/µ) to contemporary estimates of gene flow from BAYESASS, I

multiplied the M-values generated by MIGRATE by the mutation rate estimated in

popABC. I then subtracted these values from the contemporary estimates of gene flow

from BAYESASS (∆m = m - mh). The resulting value, ∆m, denotes temporal changes in

gene flow. Negative values of ∆m indicate reduced gene flow in the present, positive

values increased gene flow, and values near zero no change.

Population Bottlenecks

Since estimates of θ and M are sensitive to fluctuations in effective population

size (Beerli 2009), I conducted tests to detect bottlenecks. I tested for bottlenecks at two

generational time scales. First, I tested for bottlenecks using a mode-shift test, which is

capable of detecting bottlenecks “…within the past few dozen generations,” (Luikart and

Cornuet 1998). Older bottlenecks were tested for using a Wilcoxon’s sign-rank test,

which detects bottlenecks 25-250 generations in the past (Cornuet and Luikart 1996).

Bottleneck tests were conducted in the program BOTTLENECK version 1.2.02 (Piry et al.

1999). I ran BOTTLENECK under the SMM and the two-phase model (TPM) and tested

for heterozygosity excess. Under the TPM, I set 95% of all mutations to be single-step

with 12% variance within multi-step mutations, following the recommendation of Piry et

al. (1999). All tests were conducted using 50,000 permutations and analyzed by

Page 44: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

44

STRUCTURE cluster. Because small sample sizes can lead to low statistical power in

detecting bottlenecks (Peery et al. 2012), I also pooled all samples together and reran all

tests for the entire Chesapeake Bay (n = 617).

Results

Population Structure and Genetic Diversity

No loci were found to be in linkage disequilibrium, and across all populations and

loci, 7 of 165 loci were out of HWE at α = 0.05, but only one locus in one population

(MD1) was out of HWE after Bonferroni correction. Population genetic metrics are

provided in Table 2.3.

Preliminary runs of STRUCTURE did not detect genetic structure within

Chesapeake Bay. This was because locus D21 was nearly monomorphic. Removing this

locus ameliorated the problem, and all results presented have locus D21 removed. In

total, I identified four terrapin populations (Figure 2.1): the Patuxent River, Kent Island,

the coastal bays, and inner Chesapeake Bay. Initial runs of STRUCTURE found the

Patuxent River (MD3, MD4) to form the first cluster, with the remaining localities

forming a second cluster (Figure 2.1, ∆K = 2). Analysis of the second cluster (Figure 2.1,

∆K = 3) revealed it was composed of three subclusters: Kent Island (MD2, MD8), the

coastal bays (VA2, VA3, VA4), and inner Chesapeake Bay (MD1,5,6,7,9,10,11, VA1).

AMOVAs indicated that most of the genetic diversity in Chesapeake Bay is found

within populations. STRUCTURE clusters explained the most genetic variation (0.96% P

= 0.0031), while landscape features (0.88% P = 0.0154) and sampling locality (0.88% P

< 0.001) explained slightly less. Estimates of Dest among STRUCTURE clusters identified

Page 45: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

45

significant levels of population differentiation among all clusters (Table 2.4). Kent Island

and inner Chesapeake Bay were estimated to be the most similar (Dest = 0.0155) while the

Patuxent River and the coastal bays were estimated to be the most dissimilar (Dest =

0.0654).

Mutation Rate

ABC posterior estimates of µ solved at a mode of 4.3 x 10-4 mutations site-1

generation-1 (Figure 2.2). This estimate of µ is similar to a commonly assumed

microsatellite mutation rate of 5.0 x 10-4 (e.g. Estoup et al. 2002; Faubet et al. 2007;

Chiuchhi and Gibbs 2010).

Effective Population Size

Estimates of Ne from MIGRATE produced a range of effective population sizes in

Chesapeake Bay (Figure 2.3A). Inner Chesapeake Bay was estimated to have the largest

effective population size, at Ne = 758 (95% CI: 476 - 1398). The Patuxent River was the

next largest, at Ne = 144 (8 - 261), followed by Kent Island, at Ne = 109 (19 - 261), and

the coastal bays, at Ne = 98 (0 - 238). ONeSAMP generated the same order of population

sizes, but differed in its estimates. Inner Chesapeake Bay was estimated to have an

effective size of 302 (265 - 361), the Patuxent River = 254 (195 - 467), Kent Island =154

(139 - 188), and the coastal bays = 100 (78 - 158).

Historical Gene Flow

Estimates of historical gene flow rates revealed similar but low levels of gene

flow among all populations (Figure 2.3A). Historical gene flow levels from the coastal

Page 46: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

46

bays to inner Chesapeake Bay were the lowest of all routes (m = 0.0082), while levels of

gene flow from the Patuxent River to Kent Island were the highest (m = 0.0181).

Contemporary Gene Flow

Contemporary levels of gene flow in Chesapeake Bay showed much more

variation than did historical levels (Figure 2.3B). Gene flow leaving inner Chesapeake

Bay was the lowest of all contemporary levels (m = 0.0010), while gene flow emigrating

from the coastal bays into inner Chesapeake Bay was the highest (m = 0.3201). Gene

flow between Kent Island and the Patuxent River was also markedly higher than gene

flow between other populations (m = 0.2958 and m = 0.1584, respectively). Of the six

paired gene flow routes, only two were found to be asymmetrical (non-overlapping 95%

CI): gene flow between Kent Island and the Patuxent River, and gene flow between the

coastal bays to inner Chesapeake Bay.

Comparison of Historical and Contemporary Gene Flow

A Mantel test did not detect a relationship between historical and contemporary

gene flow (r = 0.86, P = 0.16667). Three rates were found to increase substantially

through time (Figure 2.3C, solid lines). Contemporary gene flow levels from Kent Island

to the Patuxent River and gene flow from the Patuxent River to Kent Island were much

higher than historical levels (∆m = +0.2831 and ∆m = +0.1403, respectively), as were

gene flow levels from the coastal bays to inner Chesapeake Bay (∆m = + 0.3119). I

removed these routes and performed another Mantel test, but found no significant

relationship (r = 0.86, P = 0.125), as all other routes showed varying degrees of gene

flow reduction, approximately ∆m ≈ -0.01. The Patuxent River to inner Chesapeake Bay

Page 47: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

47

(∆m = -0.0059) and the coastal bays to the Patuxent River (∆m = -0.0084) showed the

least change in gene flow levels over time.

Population Bottlenecks

BOTTLENECK tests failed to detect any signatures of heterozygosity excess in

Chesapeake Bay. Similarly, a mode-shift test failed to detect any bottlenecks, with all

populations assuming a normal “L-shaped” distribution. When the data were pooled to

correct for low power associated with small sample sizes (Peery et al. 2012), the same

results were recovered.

Discussion

Given the extent of habitat fragmentation and its contribution to the ongoing

biodiversity crisis, conservation efforts are often aimed at evaluating and ameliorating

levels of connectivity between populations (Wilcox and Murphy 1985; Hanski and

Gilpen 1997; Beier et al. 2008; Newmark 2008). Many such studies assume that

population connectivity was higher prior to anthropogenic changes, but this is not always

the case, and there is commonly a disconnect between ecological estimates of dispersal

and levels of genetic fragmentation (Kuchta and Tan, 2006; Epps et al. 2013).

I delineated four terrapin populations in the Chesapeake Bay region that exhibited

high levels of heterozygosity and allelic diversity and weak structure (Figure 2.1).

Historical estimates of migration indicate that gene flow was limited among all

populations (mh ≈ 0.01; Figure 2.3A). By contrast, contemporary estimates of migration

were more variable (Figure 2.3B). While most populations remained connected by low

levels of gene flow, substantial increases in gene flow were detected between Kent Island

Page 48: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

48

and the Patuxent River (∆m = +0.2831 and ∆m = +0.1403) and from the coastal bays into

inner Chesapeake Bay (∆m = + 0.3119; Figure 2.3C).

The documented increases in contemporary gene flow may have been human-

mediated, as terrapins are known to have been translocated into and around Chesapeake

Bay to supplement terrapin farms. Terrapins were first brought into Chesapeake Bay

from North Carolina around 1909, and were reportedly released when the terrapin farms

closed (Hildebrand and Hatsel 1926; Hildebrand 1933). This transport of terrapins could

explain my substantial increase in gene flow from the coastal bays to inner Chesapeake

Bay (∆m = +0.3119; Figure 2.3C). Alternatively, the increase in gene flow could be due

to natural processes. Hauswaldt and Glenn (2005) demonstrated that 75% of Chesapeake

Bay terrapins could be correctly assigned to their population of origin using only six

microsatellite loci, and that Chesapeake Bay populations have higher numbers of private

alleles than neighboring populations. If translocations from North Carolina represented a

large influx of genetic variation, one might expect assignment tests to confound

Chesapeake Bay terrapins with terrapins from North Carolina, which was not the case.

Increased spatial sampling is needed to determine whether my documented increase in

contemporary gene flow into Chesapeake Bay is due to the natural movement of

individuals or is due to translocation from North Carolina or another source.

The Patuxent River and Kent Island also exhibited large temporal increases in

gene flow between them (Figure 2.3C). This too may be caused by translocation, as the

largest terrapin farm in Chesapeake Bay was located on the Patuxent River at one time,

and reportedly “…consist[ed] of a large salt water lake, which could accommodate

Page 49: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

49

thousands of terrapins…” (Carpenter 1891). So far as I know, this farm was stocked prior

to imports of terrapins from North Carolina, and thus the terrapins farmed on the

Patuxent River were most likely from Chesapeake Bay. Populations located near Kent

Island and the Patuxent River represent nearby sources for the farm. Thus, anthropogenic

translocation could be the cause of the detected contemporary increases in gene flow.

Alternatively, it is possible that the increases in contemporary gene flow are the product

of natural increases in genetic connectivity, despite the high levels of habitat

fragmentation in the region. Relative to the eastern shoreline, the western shore of

Chesapeake Bay lacks jutting peninsulas (Figure 2.1). This lack of peninsulas may act as

a conduit of gene flow between Kent Island and the Patuxent River, as movement

between these populations is not as circuitous as dispersal along the eastern shore.

However, this requires dispersal distances that are not commonly documented in

ecological studies.

In contrast with the increases in contemporary gene flow discussed above, the

majority of populations exhibited decreased contemporary gene flow. Within Chesapeake

Bay, substantial habitat modification has occurred within the last century. In particular,

shorelines have become reinforced with riprap to prevent erosion. Female terrapins prefer

to nest on sandy beaches (Roosenburg 1994) and usually return to the same location each

nesting season (Szerlag and McRobert 2006). Moreover, studies show that offspring

exhibit natal philopatry (Sheridan et al. 2010). As sandy beaches are lost to shoreline

development, females are restricted to fewer nesting locations, increasing population

fragmentation. In addition, even where terrapins can nest, the mortality risk for eggs,

Page 50: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

50

hatchlings, and adult females has increased, as raccoons (Procyon lotor) and other

mesopredators thrive in human-modified landscapes (Crooks and Soulé 1999). Some

nesting locations suffer mortality rates as high as 92% for nests (Feinberg and Burke

2003) and 10% for adult females (Seigel 1980).

While gravid females face higher mortality during terrestrial nesting excursions,

juvenile females and males of all age classes experience increased mortality in aquatic

habitats as fisheries bycatch. Crab pots are used to harvest Blue Crabs (Callinectes

sapidus), and males and juvenile females (both of which are smaller than adult females)

commonly become entrapped in crab pots and drown (Roosenburg et al. 1997;

Roosenburg and Green 2000). While several states now require a bycatch reduction

device (BRD) to exclude terrapins, Maryland only requires them in recreational crab pots.

However, BRD compliance for recreational crab pots in Maryland is under 35% (Radzio

et al. 2013). Terrapins are characterized by male-biased dispersal (Sheridan et al. 2010),

and thus males in particular experience an increased risk of mortality due to fishery

activities. I suggest that the increased mortality risk of dispersing males has lowered

contemporary gene flow rates among populations.

While it is well documented that terrapins underwent population contractions due

to overharvesting and other factors (Garber 1989; 1990), I failed to find evidence of a

population bottleneck in Chesapeake Bay. This surprising result may be a consequence of

translocation, which would have subsidized populations by reintroducing genetic

diversity, and may have confounded efforts to detect a genetic bottleneck (Hauswaldt and

Glenn 2005; this study). Indeed, natural populations can be greatly effected by

Page 51: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

51

translocation, with potential benefits (Westemeier et al. 1998) or unforeseen

consequences (Frankham et al. 2002). More work on bottleneck detection using genetic

data is badly needed as bottleneck tests using heterozygosity excess may fail to detect

bottlenecks in populations known to have experienced substantial declines (Peery et al.

2012; Funk et al. 2010).

The Diamondback Terrapin has been the focus of much conservation attention,

and a number of ecological studies indicate that terrapins generally exhibit limited

dispersal and high levels of philopatry, which over time would lead to the buildup of

genetic structure (Roosenburg 1994; Spivey 1998; Gibbons et al. 2001; Butler 2002;

Sheridan et al. 2010; Converse and Kuchta, 2017). By contrast, genetic studies find that

terrapin populations are weakly differentiated, even at regional scales (Hauswaldt and

Glenn 2005; Sheridan et al. 2010; Drabeck et al. 2014; Hart et al. 2014; Petre et al. 2015).

One hypothesis to reconcile these data is that terrapins migrate large distances (several

kilometers) to mating aggregations, which would prevent the buildup of genetic structure

among populations (Hauswaldt and Glenn 2005). However, work by Sheridan (2010)

suggests that terrapins do not travel long distances to mating aggregations. I propose a

modification to Hauswaldt and Glenn’s (2005) hypothesis: that mating behavior and

population sex ratios jointly function to limit genetic structure. Under this hypothesis,

terrapins form mating aggregations near their home ranges, which homogenizes

populations at the local level. Similarly, dispersal by male terrapins promotes admixture

among mating aggregations. Male-biased dispersal has large genetic consequences

because populations in Chesapeake Bay contain highly unequal sex ratios. For example,

Page 52: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

52

at Poplar Island and the Patuxent River, female terrapins outnumber males five to one

(5:1) and three to one (3:1), respectively (Roosenburg, unpubl. data). Biased sex ratios

allow dispersing males to disproportionately contribute their genetic material to host

populations. Furthermore, female terrapins mate multiple times, store sperm, and lay

clutches of mixed paternity (Hauswaldt 2004; Sheridan 2010), increasing the odds of

mating with immigrant males. It also remains possible that ecological studies document

structure that is genetically nascent (Landguth et al. 2010).

My results have important implications for the management of species in heavily

modified landscapes. Anthropogenic habitat fragmentation is an ongoing contributor to

the biodiversity crisis, and the study of metapopulation connectivity is crucial for setting

appropriate conservation targets (Wilcox and Murphy 1985). However, current

population genetic structure is the product of the joint influence of contemporary and

historical processes, and thus to assess contemporary changes in connectivity it is

necessary to consider the historical context. Contrary to my initial hypothesis of

substantial decreases in contemporary gene flow among terrapin populations as a

consequence of habitat loss and fragmentation, I documented enormous increases in gene

flow into Chesapeake Bay and between two populations within Chesapeake Bay. I

hypothesize that this is due to translocation events associated with terrapin farming.

Without an estimate of historical levels of connectivity, however, it would not have been

clear that the high contemporary gene flow estimates were a recent phenomenon; indeed,

I may have interpreted the relatively low estimates of contemporary gene flow among

most other populations as indicative of reduced dispersal! Incorporating historical

Page 53: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

53

processes greatly improves interpretation of contemporary processes (Vandergast et al.

2007; Hansen et al. 2009; Pavlacky et al. 2009; Epps et al. 2013; Huseman et al. 2015).

My results confirm the importance of taking historical factors into account when

quantifying genetic connectivity in highly impacted landscapes.

Page 54: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

54

Table 2.1: Summary of sampling locations and sample sizes.

Locality Sample Size Latitude Longitude

Herring Bay, MD (MD1) 25 38.754362 -76.550857

Kent Island, MD (MD2) 38 38.936302 -76.363836

Patuxent River, MD (MD3) 63 38.444332 -76.607811

Buzzard’s Island, Patuxent River, MD (MD4) 60 38.489969 -76.652694

Sandy Island Cove, Nanticoke River, MD (MD5) 55 38.260110 -75.947966

Back Cove, Smith Island, MD (MD6) 17 38.021666 -75.998875

Janes Island, MD (MD7) 56 38.007513 -75.849861

Marshy Creek, Kent Island, MD (MD8) 64 38.954972 -76.227814

Northeast Cove, Bloodsworth Island, MD (MD9) 43 38.167177 -76.062002

Tylerton, Smith Island, MD (MD10) 64 37.964927 -76.020185

St. Jerome’s Creek, MD (MD11) 16 38.134924 -76.347049

Mobjack Bay, VA (VA1) 45 37.325137 -76.350877

Wachapreague, VA (VA2) 38 37.602862 -75.686380

Page 55: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

55

Table 2.1: continued

Metompkin Island, VA (VA3) 20 37.752026 -75.546442

Cedar Island, VA (VA4) 13 37.633496 -75.612748

Page 56: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

56

Table 2.2: Summary of the parameters and priors used in popABC. 2,500,000 genetic trees were simulated and a tolerance of 0.0004

was applied, resulting in 1,000 simulated data points. ICB = inner Chesapeake Bay, Patuxent = Patuxent River, Kent = Kent Island,

and CoB = coastal bays.

Parameter Description Prior

µ Mutation Rate (site-1 generation-1) Lognormal (-3.0,0.5,0.5,0.5)

Ne1 Effective Population Size, Kent Island (individuals) Uniform (0, 5000)

Ne2 Effective Population Size, Patuxent (individuals) Uniform (0, 5000)

Ne3 Effective Population Size, ICB (individuals) Uniform (0, 5000)

Ne4 Effective Population Size, CoB (individuals) Uniform (0, 5000)

NeA1 Ancestral Population Size (individuals) Uniform (0, 10000)

NeA2 Ancestral Population Size (individuals) Uniform (0, 10000)

NeA3 Ancestral Population Size (individuals) Uniform (0, 10000)

m1 Kent -> Patuxent Migration Rate (fraction of immigrants) Uniform (0, 0.1)

m2 Patuxent -> Kent Migration Rate (fraction of immigrants) Uniform (0, 0.1)

m3 ICB -> CoB Migration Rate (fraction of immigrants) Uniform (0, 0.1)

m4 CoB -> ICB Migration Rate (fraction of immigrants) Uniform (0, 0.1)

mA1 Ancestral Migration Rate (fraction of immigrants) Uniform (0, 0.1)

Page 57: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

57

Table 2.2: continued

mA2 Ancestral Migration Rate (fraction of immigrants) Uniform (0, 0.1)

Tev1 Splitting Event 1 (years) Uniform (0, 5000)

Tev2 Splitting Event 2 (years) Tev1 + Uniform (0, 5000)

Tev3 Splitting Event 3 (years) Tev2 + Uniform (0, 5000)

T Generation Time 12 years (constant)

Top Tree Topology (18 possible arrangements) Uniform (0, 18)

Page 58: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

58

Table 2.3: Mean observed (Ho) and expected (He) heterozygosities and number of

alleles (NA) for each locus.

Locus Ho He NA

B91 0.49 0.53 2.29

B08 0.84 0.84 7.71

D93 0.52 0.53 4.57

A18 0.76 0.73 6.07

D87 0.84 0.84 10.57

B67 0.40 0.42 2.07

D90 0.87 0.83 9.00

D55 0.86 0.84 10.50

D114 0.76 0.73 8.14

D121 0.90 0.89 11.86

D62 0.81 0.82 8.71

Page 59: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

59

Table 2.4: Estimates of population differentiation. Values below the diagonal are Dest values, with 95% CIs in brackets. Values above

the diagonal represent P-values, with Bonferroni corrections shown in parentheses.

Kent Island Patuxent River Coastal Bays Inner CB

Kent Island - 0.001

(0.006)

0.001

(0.006)

0.001

(0.006)

Patuxent River 0.0443

[0.0380 – 0.0523]

- 0.001

(0.006)

0.001

(0.006)

Coastal Bays 0.0374

[0.0320 – 0.0486]

0.0654

[0.0576 – 0.0747]

- 0.001

(0.006)

Inner CB 0.0155

[0.0116 – 0.0219]

0.0291

[0.0254 – 0.0343]

0.0333

[0.0274 – 0.0414]

-

Page 60: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

60

Figure 2.1: Sampling localities and STRUCTURE results. Top left: the distribution of

Terrapins (shaded red), and the location of the study (black box). Main figure: four

terrapin populations were demarcated. Triangles indicate the Patuxent River, stars

represent Kent Island, squares represent the coastal bays, and circles represent inner

Chesapeake Bay. At ∆K = 2, the Patuxent River forms a cluster while the remaining

sampling localities form a second cluster. This second cluster is composed of three

subclusters (∆K = 3), which contains Kent Island, inner Chesapeake Bay, and the coastal

bays.

Page 61: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

61

Figure 2.2: Approximate Bayesian computation posterior (solid line) and hyperprior

(dotted line; log-normal) distributions for l, the mutation rate used to convert θ into Ne

and M into mh (MIGRATE), for comparisons with m from BAYESASS. The mode is

1.9 x 10_3.36, or 4.39 x 10

_4 mutations site_1 generation

_1.

−4.0 −3.5 −3.0 −2.5

0.0

0.5

1.0

1.5

Mutation Rate µ (1 x 10x),

Dens

ity

mode = 1 x 10-3.3618

2.5% = 1 x 10-2.7063

97.5% = 1 x 10-3.7690

Page 62: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

62

Figure 2.3: Gene flow in Chesapeake Bay. KI = Kent Island, PR = the Patuxent River,

ICB = inner Chesapeake Bay, and CoB = the coastal bays. Thin lines represent estimates

of m or ∆m of <0.01, intermediate lines represent estimates of 0.01–0.05, and thick lines

represent estimates of >0.05. (A) Results from MIGRATE. Numbers within boxes denote

Ne, while values above arrows indicate proportion of immigrants (mh). (B)

Contemporary gene flow rates determined by BAYESASS. Numbers within boxes

indicate the proportion of individuals to remain within the population, and values above

arrows indicate the proportion of immigrants (m) to their respective populations. (C)

Historical gene flow rates subtracted from contemporary rates (m – mh = ∆m). Dashed

arrows indicate gene flow routes that have reduced contemporary gene flow (-∆m) and

solid arrows indicate routes that have increased contemporary gene flow (+∆m).

Page 63: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

63

CHAPTER 3: EFFECTIVE POPULATION SIZE IN DIAMONDBACK TERRAPINS:

INCONGRUENCE BETWEEN MOLECULAR AND ECOLOGICAL ESTIMATES

Introduction

Effective Population Size

Effective population size (Ne) is a fundamental concept in evolutionary biology,

and is used in many fields from population genetics and phylogenetics to conservation

genetics and molecular ecology. Proposed by Wright (1931), the effective population size

was defined as "the number of breeding individuals in an idealized population that would

show the same amount of dispersion of allele frequencies under random genetic drift or

the same amount of inbreeding as the population under consideration.” Frankham et al.

(2010) recently defined the effective population size as “The size of an ideal population

that loses heterozygosity (or becomes inbred or drifts) at the same rate as the observed

population.” The concept of an “ideal population” is central to any definition of effective

population size.

A commonly used model of the “ideal population” is the Wright-Fisher model

(Fisher 1922; Wright 1931; Crow 1987, 2010). In this model, generations do not overlap,

each individual reproduces a mean of one offspring (to maintain constant population

size), the population is panmictic, two alleles with frequencies p and q are present, and

selection and mutation are absent. These assumptions were made to evaluate how allele

frequencies behave solely under the process of genetic drift (Tran et al. 2013). The

original Wright-Fisher model has since been expanded to incorporate overlapping

generations and multiple alleles across loci (Moran 1958; Tran et al. 2013). More

recently, effective population has been scaled by gene genealogies (the coalescent),

Page 64: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

64

allowing population genetic processes to be incorporated into higher-level (phylogenetic)

analyses (Nordberg and Krone 2002; Sjödin et al. 2005; Wakeley 2009). Indeed, the

concept of the effective population has permeated into many fields of biology, including

species demarcation (de Queiroz 2007; Kuchta and Wake 2016).

Although the effective population size is a fundamental evolutionary concept and

assumes a simple population model, there is no universal statistic for its estimation

(Frankham et al. 2010). Rather, several statistics exist to estimate Ne, which can be

derived from sex-ratios, organismal ploidy, allele frequencies, levels of inbreeding, and

other approaches, as well as combinations of these statistics (Fisher 1922; Wright 1931,

1938; Crow 2010). A sample of these statistics is found in Table 3.1.

The estimation of Ne from multiple methods can create confusion with respect to

which approach is best suited to a specific study system. This confusion is compounded

by reference to “the” effective population size, which implies a population can only

exhibit a single effective population size (Sjödin et al. 2005). Moreover, Ne is sometimes

incorrectly associated with the census population size (Nc), which has a singular value at

one point in time (Nordberg and Krone 2002). Thus, the complexities in estimating Ne

can lead to misconceptions. In addition, the assumptions made by “idealized” populations

are violated by all populations to varying degrees (Frankham et al. 2010). For example,

Elephant Seals (Mirounga angustirostris) are extremely polygynous, and consequently

deviate from assumptions of Fisherian sex-ratios and panmixia (Fabiani et al. 2004).

Thus, unless there is a clear biological basis for choosing a particular model, which is

Page 65: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

65

often not the case, it is useful for studies investigating effective population size to utilize

a variety of approaches.

Study System

Diamondback Terrapins (Malaclemys terrapin) inhabit brackish water habitats

along the Gulf and eastern coastlines of the United States (Ernst and Barbour 1989).

Terrapin populations typically exhibit weak structure at the local and regional scales, and

genetic diversity tends to be higher in Atlantic populations than in Gulf populations

(Hauswaldt and Glenn 2005; Hart et al. 2014; Converse and Kuchta 2017). Near

Chesapeake Bay, four genotypic clusters have been delimited (Converse et al. 2015):

Kent Island (KI), the Patuxent River (PR), inner Chesapeake Bay (ICB) and the coastal

bays (CoB) constitute the known genotypic clusters. However, the degree of population

structure is weak and exhibit high levels of admixture. All populations are characterized

by high levels of heterozygosity, high mean number of alleles, and high allelic richness

(Figure 3.1; Converse et al. 2015). Gene flow is highest between KI and the PR, and from

the CoB to ICB (Converse et al. 2015).

Poplar Island (PI) is situated in Chesapeake Bay approximately 26 km south of

the Bay Bridge and is located within the ICB genotypic cluster (Figure 3.1; Converse et

al. 2015). Poplar Island is actively under construction as an ecological restoration site,

with wetlands and upland forests in different stages of completion. Because female

terrapins nest on sandy beaches (Roosenburg 1994), PI has vast areas with suitable

nesting habitat. Female terrapins lay hundreds of nests on PI each year, and their nests

enjoy relatively low levels of mortality due to the absence of mammalian nest predators

Page 66: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

66

(Feinberg and Burke 2003; Roosenburg et al. 2014). Furthermore, the construction of

artificial wetlands has created habitat capable of sustaining terrapin populations on PI for

the foreseeable future. The combination of optimal nesting habitat, low nest mortality,

suitable habitat for hatchling and adult terrapins, and a lack of predators creates unique

circumstances that allow biologists to study terrapin population dynamics in fine detail.

Nesting and demographic studies as well as survivorship experiments among cohorts

have been ongoing on PI for over a decade (Roosenburg et al., unpublished data).

Thus, terrapin populations in Chesapeake Bay provide an excellent opportunity to

estimate effective population sizes using molecular and ecological approaches, especially

for populations that exhibit weak structure. In this article, I compare estimates of

effective population size generated from molecular and ecological data, and assess

congruence between the two approaches. I further estimate Ne from molecular data using

a variety of methods, including approximate Bayesian computation (ABC), model testing

in a Bayesian framework, and frequentist-based methods, to test for levels of

concordance and discordance when estimating Ne with different methodologies.

Methods

Molecular Data

I used 11 polymorphic microsatellite loci developed in a species that is closely

related to the Diamondback Terrapin, the Bog Turtle (Glyptemys muhlenbergii), for all

molecular estimates of effective population size (King and Julian 2004), a dataset given

to me by Timothy King. I used the genotypic clusters demarcated by Converse et al.

(2015) (KI, PR, ICB, CoB) as a priori population assignments for all molecular estimates

Page 67: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

67

of effective population size. PCR chemistry and protocols for these loci are detailed in

King and Julian (2004).

To estimate effective population sizes in a Bayesian model-testing framework, I

used MIGRATE v. 3.6.11 (Beerli 2008) and compared 12 models of population structure

(Figure 3.2). Each model used uniform distributions to model ϴ (=4Neµ) and M (=m/µ),

bounded between 0.001 and 100, and 0 and 2000, respectively. Individual models were

explored using a Metropolis coupled Markov chain Monte Carlo (MC3) algorithm

composed of 20000 retained steps, with a 50% burn-in and 500 step sampling interval. To

improve speed, I ran a Brownian motion model to approximate a stepwise-mutation

model, and randomly subsampled 40 individuals at each locus. To facilitate model

comparisons, I used an identical starting seed for each model, whereas each model was

composed of the same random genotypic subsample at each locus. I tested for the best-fit

model using approximate Bézier scores (Beerli and Palczewski 2010), which require the

posterior be explored with at least four heated parallel chains (heated at 1.0, 1.5, 3.0, and

1,000,000, respectively). For comparative purposes, I also used harmonic means to select

the best-fit model. I replicated each model five times, for a total of 100000 recorded steps

per model. To probe model selection consistency, I reran these analyses independently

for a second time. I then took the best-fit models from the first and second analyses as

determined by Bézier scores and harmonic means and reran the analyses using 25

replicates (Figure 3.3). To convert ϴ into Ne, I multiplied ϴ values from the best-fit model

by a mutation rate of 4.34 x 10-4 mutations site-1 generation-1 (Converse et al. 2015).

I further estimated effective population sizes using approximate Bayesian

Page 68: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

68

computation (ABC; Beaumont et al. 2002) with the softwares ONeSAMP v. 1.2 (Tallmon

et al. 2008) and DIYABC v. 2.04 (Cornuet et al. 2014). In contrast with MIGRATE,

ONeSAMP estimates Ne one population at a time, rather than jointly estimating ϴ values.

ONeSAMP uses a suite of eight common summary statistics (e.g. Wright’s FIS, observed

heterozygosity) to infer Ne. For these analyses, I set the lower boundary of Ne to 2 and the

upper boundary to 1000, the minimum and maximum values allowed.

For DIYABC, I investigated a single scenario where terrapins colonized

Chesapeake Bay from south to north (northward range expansion). Since ABC models

are computed relatively quickly, I analyzed the entire genotypic dataset as opposed to a

subsample. I modeled individual locus mutation rates with a gamma distribution under

shape 2, with lower and upper boundaries of 1 x 10-5 and 1 x 10-2, respectively. The

overall mean mutation rate was modeled with a uniform distribution bounded between 1

x 10-4 and 1 x 10-3. A generalized stepwise model was used for microsatellite mutation,

which allows repeats to increase or decrease by one or more motifs under a geometric

distribution (Cornuet et al. 2008). Model parameters for Ne and divergence dates are

provided in Table 3.2. I used 34 summary statistics to generate 1000000 simulated data

points, retained the closest 0.1% of simulated data, and then applied polychotomous

logistic regression to the retained data (Beaumont et al. 2002). Effective population sizes

and divergence date estimates were logit transformed and observed under original model

parameters. To test goodness-of-fit, I used the model checking function in DIYABC

(Cornuet et al. 2014). This function checks the distribution of simulated data points under

the selected model and displays its estimates, which must fall within the observed

Page 69: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

69

summary statistics to be deemed congruent. Following the recommendations of Cornuet

et al. (2010), I used a new set of summary statistics to evaluate goodness-of-fit and did

not use summary statistics applied to parameter estimates or model selection (Table 3.2).

Finally, I estimated effective population sizes in NeEstimator v. 2.0 (Do et al.

2014) using heterozygosity excess (Zhdanova and Pudovkin 2008), linkage

disequilibrium (Waples and Do 2010), and molecular coancestry (Nomura 2008). I used a

Pcrit value of 0.05 for all analyses. The Pcrit value is the lowest frequency at which an

allele occurs. If the Pcrit value is too low, estimates of effective population size may

become biased, as the analysis may incorporate alleles found in a single heterozygote in

the population (Do et al. 2014).

Ecological Approaches

I captured terrapins and collected mark-recapture data on Poplar Island (Figure

3.1; 38.760886, -76384044) from 2005-2015. Terrapins were captured with fyke nets,

modified crab pots, dip nets, and by hand (Figure 3.1), from May-August each summer. I

deployed 2-4 fyke nets regularly and used other trapping methods opportunistically. Fyke

nets were deployed at 07:30 on Mondays, checked at 07:30 Tuesday through Friday, and

removed Friday. Trapping locations were rotated every two weeks to ensure thorough

sampling of wetland cells, Poplar Harbor, as well other areas adjacent to Poplar Island.

Captured terrapins were marked with two systems. The first was a passive

integrated transponder (PIT) tag (Biomark, Boise, ID, USA), which emits a unique 10

digit alphanumeric code. This code is read by a PIT tag reader (Biomark, Boise, ID,

USA) allowing for reliable individual identification. Betadine (Purdue Products L.P.,

Page 70: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

70

Stamford, CT, USA) was used to sterilize the integument in the right femoral pocket, and

PIT tags were injected subcutaneously. NewSkin (Prestige Brands, Tarrytown, NY, USA)

was used to seal the PIT tag injection site.

Wing bands (“PI-tags,” National Band & Tag Company, Newport, KY, USA)

were employed as the second system to identify individuals. PI-tags were engraved with a

four digit numeric code proceeded by “PI” (e.g., PI 1234). A small hole was drilled in the

ninth right marginal scute of captured individuals and the PI-tag secured in place with

needle-nose pliers. Tissue displaced during shell boring was collected and stored in a

marked tube containing 95% ethanol. Although PI-tags have fewer combinations relative

to PIT tags, they do not require an additional reader to be interpreted, which facilitates

quick identification of captured individuals in the field.

In addition to individual markings, I measured (in mm) several morphological

traits for all captured terrapins: head width (HW), plastron length (PL), plastron width

(PW), carapace height (CH), carapace length (CL), and right pectoral plastron scute (RP),

as well as mass (in grams), and age (in years). Individuals were sexed, with individuals

that were sexually ambiguous or too young to sex scored as juveniles. If an individual

had 10 or more visible annuli reflecting clearly visible growth in recent years, I

documented its age as >10 because growth lines in older individuals become unreliable. I

did not age younger individuals if the growth rings were too obscured to provide and

accurate estimate of age. I assessed the accuracy of aging by confirming the age of

recaptures of marked individuals in addition to the recaptures of more than 400 marked

hatchlings originally marked at time of hatching.

Page 71: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

71

To estimate the census population size at Poplar Island, I evaluated several

models in MARK v. 8.0 (White and Burnham 1999). I created two versions of my mark-

recapture dataset: the first dataset included males and females as separate groups, while

the second dataset scored males, females, and juveniles as separate groups. For my

MARK analyses, I estimated Nc with models assuming an open population.

I estimated Nc with POPAN (Schwarz and Arnason 1996) and Burnham Jolly-

Seber models (Jolly 1965; Seber 1965; Burnham 1991). Open population models use

survival rates and capture probabilities (p) to estimate Nc (Cooch and White 2014). They

also allow individuals to enter or leave the population in question. However, because

permanent emigration and individual death result in the same observation (no further

recaptures), they are assumed under a single parameter, Φ (apparent survivorship). Open

population models further assume that capture and survival probabilities are equal

between marked and unmarked individuals, and that the area of the study does not change

over time.

POPAN models estimate Nc from Φ, p, and the probability of entering the

population (PENT). I used the logit link function for parameters Φ and p, the log link

function for N (the number of groups for which to estimate Nc), and the multinomial logit

(mlogit) link function for all PENT parameters, as recommend by Cooch and White

(2014). I tested several POPAN models assuming time-dependent or constant

survivorship (e.g., Φ(t) , Φ(.) ), constant and time-dependent capture probabilities (e.g., p(.) ,

p(t) ), and constant and time-dependent entrance probabilities (e.g., PENT(.) , PENT(t)). I

also tested models by sex and with combinations of these assumptions. To facilitate

Page 72: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

72

model selection, I used AICc scores (Akaike 1973, 1974) and selected the best-fit model

for each dataset.

Burnham Jolly-Seber models use the parameters Φ, p, and λ (growth rate) to

estimate Nc, and add an initial population size parameter (No) at the beginning of the

analysis. The PENT parameter from the POPAN model is incorporated in the λ parameter

in the Burnham Jolly-Seber model (Cooch and White 2014). I used the logit link function

for Φ, p, and λ, and the log link function for N, as recommend by Cooch and White

(2014). To assess model fit, I used AICc scores for all analyses.

Results

Molecular Estimates of Ne

Model testing in MIGRATE indicates that the Against Flow (Conduit) model is the

best-fit model for Chesapeake Bay population structure (Bézier = -167283.87, Figure 3.2,

Table 3.3). All population models greatly outperformed the No Flow model (Figure 3.2,

Table 3.3). From MIGRATE, Kent Island was found to have the smallest effective

population size (Table 3.4; mean = 516, median = 678, mode = 330; Figure 3.4),

followed by the Patuxent River (5670, 4399, 1105), the coastal bays (17210, 17616,

17733), and inner Chesapeake Bay (36095, 36609, 33430).

Effective population size estimates in DIYABC were less variable within

populations (Table 3.4; Figure 3.4). DIYABC estimated the Patuxent River to have the

smallest effective population size (mean = 4650, median = 4390, mode = 3660), followed

by Kent Island (7100, 7510, 9220), the coastal bays (13500, 13300, 13200), and inner

Chesapeake Bay (32700, 33100, 35300). The other ABC software, ONeSAMP, generated

Page 73: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

73

effective population size estimates that were approximately 5-10% of those from

DIYABC and MIGRATE (Table 3.4; Figure 3.4). The coastal bays were found to have the

smallest effective size (mean = 98, mode = 96), followed by Kent Island (156, 154), the

Patuxent River (263, 254), and inner Chesapeake Bay (305, 302).

Effective population size estimates from NeEstimator varied by analysis. Effective

population size estimates using linkage disequilibrium estimated Kent Island to be the

smallest (mean = 266), followed by the coastal bays (551), the Patuxent River (1697) and

inner Chesapeake Bay (+∞). Estimates of effective population size using molecular

coancestry estimated Kent Island to be the smallest (36), followed by inner Chesapeake

Bay (67). Molecular coancestry found the effective sizes of the Patuxent River and the

costal bays to both be infinity (+∞). Finally, effective population sizes estimates derived

from heterozygosity excess were uninformative, as all populations were estimated to have

Ne = +∞ (Table 3.4; Figure 3.4).

Ecological Estimates of Ne

From 2006 to 2015, there were 3576 female captures, 532 males, and 1224

juveniles, for a total of 5332 unique captures (Table 3.5). Mark-recapture data suggest

that the terrapin population on PI is growing, as the number and proportion of juveniles

captured increased steadily (Table 3.5). Using non-Fisherian sex-ratios and raw mark-

recapture data, the effective population size near PI is approximately 1822.

My estimates of Nc using Burnham Jolly-Seber open population models

consistently failed to reach numerical convergence (Cooch and White 2014). During

likelihood estimation, Burnham Jolly-Seber models apply penalty constraints to keep

Page 74: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

74

parameters consistent relative to one another (Cooch and White 2014). A majority of my

models failed to converge, and thus did not generate results. However, several simpler

models did reach numerical convergence, but their estimates were not reliable (e.g. Nc of

9.0 x 1013).

AICc scores from the POPAN analyses show that the Φ(t,s)p(t,s)PENT(t)N(s) model is

the best-fit model for both the male-female and male-female-juvenile datasets (Table

3.6). POPAN estimates of Nc with the male-female dataset indicate there are

approximately 943 males in the population [812 – 1080 95% CI; 59.95 SE] and 4533

females [4375 – 4697; 81.98], for a total (Nc) of 5476 adults [5208 – 5765] (Table 3.7).

When juveniles are included, the male population is estimated to be 936 [833 – 1068;

68.36], the female population is 4680 [4511 – 4856; 88.00], and juveniles 1673 [1621 –

1727; 26.98], for a total (Nc) of 7289 individuals [6944 – 7663] (Table 3.7). These

estimates do not reflect the number of hatchlings in the population.

Using the Nc estimates from POPAN and non-Fisherian sex-ratios, the PI

population has an effective size of 3120 [2753 –3534] when juveniles are accounted for,

and 3122 [2799 – 3481] when they are not (Table 3.7).

Discussion

Molecular Approaches

I found considerable variation in effective population size among my molecular

estimates using the same dataset (Table 3.4; Figure 3.4). ONeSAMP and NeEstimator

generated small estimates of Ne relative to MIGRATE and DIYABC for all genotypic

clusters (Table 3.4; Figure 3.4). More congruence was observed between MIGRATE and

Page 75: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

75

DIYABC for the CoB and ICB clusters, but these methods diverged in their estimates for

the PR and KI clusters (Table 3.4; Figure 3.4). For NeEstimator, I could not determine Ne

estimates from molecular coancestry and linkage disequilibrium, or for any Ne estimates

from heterozygosity excess.

There are several explanations for the observed incongruence among molecular

approaches. The time-scale over which Ne was estimated varied greatly by analysis. For

example, MIGRATE operates in a coalescent framework, while NeEstimator does not

(Beerli 2008; Do et al. 2010). Consequently, MIGRATE estimates effective population

sizes over evolutionary time, which can approach 4Ne generations (thousands of years) if

the population is sufficiently large (Beerli 2009). In contrast, NeEstimator estimates the

effective population size over “contemporary time,” which Do et al. (2014) define as

“…the time periods encompassed by the samples.” Thus, MIGRATE’s estimates of Ne

predate European colonization of Chesapeake Bay, while NeEstimator’s estimates of Ne

encompass 2003 - 2005, the time window my genetic samples were collected (Converse

et al. 2015).

Differences in prior distributions also affect estimates of Ne. MIGRATE and

ONeSAMP both estimate Ne from ϴ (=4Neµ), but differ greatly in how the ϴ prior can be

set. MIGRATE allows ϴ to assume several prior distributions, while ONeSAMP only

estimates ϴ under a uniform distribution. Furthermore, the uniform distributions used by

these two programs are not the same. MIGRATE allows the lower and upper boundaries

of the uniform prior to be customized, while ONeSAMP does not. For MIGRATE, I set the

ϴ prior to be uniform between 0.001 and 100 (P. Beerli, personal communication), while

Page 76: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

76

ONeSAMP bounded the distribution between 2 and 12 (Tallmon et al. 2008). Combined

with summary statistics, ONeSAMP’s estimates of Ne cannot exceed 1000 (Tallmon et al.

2008), well below MIGRATE’s upper estimate of ~57500 given the mutation rate

(Converse et al. 2015). Likewise, differences in summary statistics may influence

estimates of Ne for ABC approaches. DIYABC allows up to 12 summary statistics for

microsatellites, while ONeSAMP uses the same eight summary statistics for all analyses

(Tallmon et al. 2008; Cornuet et al. 2014).

The population model assumed for each analysis also influences estimates of Ne.

Gene flow, inbreeding, mutation rates, and recombination rates all affect Ne (Frankham et

al. 2010). MIGRATE incorporates gene flow into its Ne estimates, while DIYABC,

ONeSAMP, and NeEstimator do not (Table 3.8; Beerli 2008; Tallmon et al. 2008; Do et al.

2010; Cornuet et al. 2014). MIGRATE, DIYABC, and ONeSAMP incorporate mutation

rates, but differ in how they are modeled. MIGRATE uses either a stepwise-mutation

model (Kimora and Ohta 1978) or a Brownian motion model to approximate a stepwise-

mutation model, while DIYABC assumes a geometric distribution of motif mutation

(Cornuet et al. 2008). ONeSAMP does not allow the mutation rate to be modeled (Tallmon

et al. 2008). Only DIYABC and NeEstimator (linkage disequilibrium) incorporate

recombination rates into Ne estimates, and only ONeSAMP incorporates inbreeding into its

Ne estimates (Table 3.8).

Harmonic means did not agree with Bézier scores for the best-fit model (Table

3.8). Harmonic means are known to poorly approximate likelihood calculations using

MCMC (Beerli and Palczewski 2010), and MCMC methods are often used in population

Page 77: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

77

genetics. As computation power increases, other likelihood approximations (e.g. Bézier

scores) are more feasible.

Weak population structure and high admixture levels may have impacted my

estimates of Ne and was a cause of incongruence among the methods. High admixture is

usually a result of high gene flow (Wakeley 2009). In my study, MIGRATE was the only

software that quantified gene flow when estimating Ne. Thus, in weakly structured

populations, molecular estimates of Ne that ignore gene flow may be greatly impacted.

Although DIYABC does not estimate gene flow, it was most similar to MIGRATE (Table

3.4; Figure 3.4). This congruence is most likely attributable to the population models of

both analyses. Ne estimates from MIGRATE assumed the Against Flow (Conduit) model

(Figure 3.2), while DIYABC assumed a one way, south to north colonization model.

Ecological Approaches

Burnham Jolly-Seber open population models often failed to generate estimates of

Nc, most likely not due to a disconnect between model assumptions and biological reality.

For the PI population, the assumptions of the Burnham Jolly-Seber model are reasonable

(and similar to POPAN models). The complexity of the likelihood function, coupled with

penalty constraints, prevented my models from generating estimates of Nc.

POPAN estimates of Nc from both terrapin datasets recovered the same model

(Φ(t,s)p(t,s)PENT(t)N(s)) using AICc scores (Table 3.6). In these models, all parameters vary

by time, and all parameters vary by sex, with the exception of PENT. Previous work

(Sheridan et al. 2010) suggest that males disperse further than females, but my data

suggests males and females (and juveniles) have the same probability of entering the PI

Page 78: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

78

population. While it is often stated that effective population sizes are ~10% of the census

size (Frankham et al. 2010), my mark-recapture data with ecological Ne estimates

suggests the ratio of Ne to Nc on PI is 56% [52 – 59%] when juveniles are not factored in,

and 43% [40 – 45%] when they are (Table 3.7). As the PI population continues to grow,

this percentage may decrease, especially if sex-ratios deviate further from Fisherian

ratios.

Incongruence Between Molecular and Ecological Estimates of Ne

For the ICB genotypic cluster, I observed incongruence between ecological and

molecular estimates of Ne (Table 3.4, Table 3.7; Figure 3.4). Differences in temporal and

geographic scale are most likely responsible for this incongruence. MIGRATE, DIYABC,

and ONeSAMP incorporate mutation rates to estimate Ne, and thus their estimates are over

coalescent time (Wakeley 2009). Recent changes in population dynamics are difficult to

detect with these methods. Indeed, precipitous and recent changes in population

dynamics can introduce ambiguity in coalescent histories, and can result in specious

parameter estimates (Beerli 2009). By contrast, ecological mark-recapture methods

operate over contemporary time, and are sensitive to recent population changes (Besbeas

et al. 2002; Baker et al. 2004). Thus, my genetic approaches may be estimating historical

effective population sizes, while my ecological approaches estimate contemporary

effective population sizes. However, simultaneous differences in geographic scale

complicate comparison between the molecular and ecological estimates of Ne.

Terrapin populations are known to exhibit weak genetic structure but demonstrate

nest-site and home range fidelity (Auger 1989; Lovich and Gibbons 1990; Spivey 1998;

Page 79: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

79

Gibbons et al. 2001; Hauswaldt and Glenn 2005; Sheridan et al. 2010; Hart et al. 2014;

Petre et al. 2015; Converse et al. 2015; Converse and Kuchta 2017). Over regional scales,

this may be of large consequence for estimates of Ne. All of the molecular approaches

assumed the ICB cluster exhibited no substructure, and thus estimates of Ne reflect the

entire ICB genotypic cluster. On the other hand, the open population models and mark-

recapture data do not make this assumption. The POPAN and Burnham Jolly-Seber

models explicitly state they assume the population in question experiences emigration

and immigration (Jolly 1965; Seber 1965; Burnham 1991; Schwarz and Arnason 1996).

Consequently, the mark-recapture data may be estimating Ne for a sub-population within

the ICB cluster that cannot be detected genetically due to a complete lack of population

structure. Thus, the two approaches (molecular and ecological) may be estimating Ne

over two geographic scales as well as two time scales.

Congruence between Ecological and Molecular Estimates

Weak population structure may be a cause of incongruence between the molecular

estimates of Ne, adding uncertainty as to which inferences are to be trusted. High

admixture and weak population structure likely impact estimates of Ne that do not

incorporate population connectivity, and Chesapeake Bay terrapin populations exhibit

these characteristics (Converse et al. 2015). Since these populations are connected by

high levels of gene flow, the Ne estimates from MIGRATE (~35000) are likely the most

accurate. DIYABC inferences of Ne for this genotypic cluster are similar to MIGRATE

(~33000). Indeed, the mark-recapture data suggest the terrapin population near PI harbors

at least ~6900 individuals, and an Ne of at least ~2700. This result renders estimates of Ne

Page 80: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

80

from ONeSAMP (~300) and NeEstimator (~67) for the entire ICB cluster suspect. My data

suggest the ICB cluster has an historic size of ~35000, and the local population near PI

has a contemporary Ne of ~3100.

Effective Population Size in Chelonians

Studies of effective population size in Chelonians have revealed considerable

variation within and among species. Within Chesapeake Bay, Hart et al. (2014) recovered

95% confidence intervals for KI terrapins at Ne = 7533 – +∞. This interval falls near or

within my estimates of 7100 (mean), 7500 (median), and 9220 (mode) from DIYABC

(Table 3.4; Figure 3.4). Likewise, Hart et al. (2014) estimated a 95% CI for PR to be 247

– +∞, consistent with my estimates from ONeSAMP (mean = 263, median = 254; Table

3.4; Figure 3.4). Among species of turtles, estimates of Ne vary widely. Pacific Green

Turtles (Chelonia mydas) near Michoacan, Mexico exhibit a female effective size of

1900 – 2300 (Chassin-Noria et al. 2004); in Australasia, Ne estimates range from 300 –

125000 (Dethmers et al. 2006). Flatback Turtles (Nanator depressus) in Australia exhibit

an effective size of 1200 – 4200 (Theissinger et al. 2009). In Michigan, USA, Eastern

Box Turtles (Terrapene carolina) populations range from Ne = 6675 – 9516 (Marsack

and Swanson 2009). Northwest Atlantic Leatherback Turtles (Dermochelys coriacea) are

estimated to have undergone precipitous drops in effective population size, from Ne ~

3500000 to Ne = 70 (Molfetti et al. 2013). My estimates of Ne in terrapins generally fall

within estimates of effective size for other Chelonians.

My study documented incongruence among molecular approaches for estimates

of Ne, and suggests that studies should use more than one metric to evaluate Ne. The

Page 81: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

81

genetic diversity within a population is the result of many complex biological phenomena

over varying time scales, and currently no single population genetic model can

encompass this complexity. Consequently, populations are studied with several simpler

models, which may make different underlying assumptions. My study demonstrates that

ecological data can help rule out spurious estimates of Ne when molecular data conflict.

Table 3.1: A sample of equations used to estimate effective population sizes from

different types of data. Variables shown are allele frequency (p), the number of males

(Nm), the number of females (Nf), and time (t).

Type Equation

Allele frequencies

Non-Fisherian sex-ratios

Inbreeding

Temporal variation (harmonic mean)

Page 82: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

82

Table 3.2: Parameters used to determine Ne in DIYABC. Ne is measured in individuals

and Tn in years.

Parameter Distribution Conditions

Ne1 (Kent) U(10, 10000) None

Ne2 (Pat) U(10, 10000) None

Ne3 (ICB) U(10, 50000) None

Ne4 (CoB) U(10, 25000) None

T1 (Pat, Kent) U(0, 5400) T2 >= T1

T2 (ICB, (Pat, Kent) T1 + U(0, 5400) T3 >= T2

T3 (CoB, (ICB, Pat, Kent) T2 + U(0, 5400) T3 > T1

Page 83: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

83

Table 3.3: Population models and their respective Bézier scores and harmonic means (HM). Models are found in Figure 3.

Scores marked with an asterisk (*) denote the best-fit model from each run. These best-fit models were compared among each other a

third time, with 25 replicates. The Against Flow (Conduit) model was found to fit the data best.

Model Bézier 1 HM 1 Bézier 2 HM 2 Bézier 3

Against Flow (Conduit) -696,050.01*

-2,900.31

-741,885.21

-12,516.79

-167,283.87*

Against Flow -1,126,892.35

-13,185.13

-1,131,493.31

-3,672.51

NA

Coastal Escape (Conduit) -1,403,164.66

2,426.40

-1,378,049.31

-2,626.78

NA

Central Conduit -1,406,497.91

-2,570.59

-1,430,924.77

-2,415.99

NA

With Flow -1,532,529.78

-22,073.41

-454,854.51*

-3,122.33

-940,968.69

Tributary Trap (Conduit) -1,863,853.60

-2,225.91

-1,188,523.09

-2,344.78

NA

Full Matrix -1,918,529.84 -2,857.01 -1,757,393.44 -2556.35 NA

Page 84: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

84

Table 3.3: continued

Coastal Escape -2,057,439.59

-2,382.84

-621,170.64

-2,889.32

NA

Central Sink -2,295,251.84

-2,337.15

-1,839,373.89

-2,180.14*

-1,468,248.55

Tributary Trap -2,386,171.33

-2,138.72*

2,062,235.94

-2,739.67

-1,576,769.90

With Flow (Conduit) -6,136,159.11

-969,642.15

-4,999,347. 08

-608,126.09

NA

No Flow -27,133,017.74 -18,303,896.36

-49,470,297.80

-39,417,145.05

NA

Page 85: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

85

Table 3.4: Effective population size estimates from molecular data. Effective population sizes are listed under mean/median/mode.

Values in brackets denote 95% confidence intervals. ONeSAMP does not report modes while

NeEstimator only reports means. LD = linkage disequilibrium, HE = heterozygosity excess, and MoCo = molecular coancestry.

Method Ne ICB Ne KI Ne PR Ne CoB

MIGRATE 36095/36609/33430 516/678/330 5670/4399/1105 17210/17616/17733

[15000 – 58140] [0 – 1473] [0 – 44380] [16318 – 19147]

DIYABC 32700/33100/35300 7100/7500/9220 4650/4390/3660 13500/13300/13200

[15600 – 46700] [3010 – 9720] [1650 – 8700] [7350 – 19800]

ONeSAMP 305/302/NA 156/154/NA 263/254/NA 98/96/NA

[265 – 361] [139 – 188] [195 – 467] [78 – 157]

NeEstimator

LD +∞/NA/NA 266/NA/NA 1697/NA/NA 551/NA/NA

[+∞ – +∞] [149 – 861] [375 – +∞] [182 – +∞]

Page 86: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

86

Table 3.4: continued

HE +∞/NA/NA +∞/NA/NA +∞/NA/NA +∞/NA/NA

[73 – +∞] [23 – +∞] [23 – +∞] [27 – +∞]

MoCo 67/NA/NA 36/NA/NA +∞/NA/NA +∞/NA/NA

[5 – 205] [27 – 668] [+∞ – +∞] [+∞ – +∞]

Page 87: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

87

Table 3.5: Unique individual captures by sex and year.

Females Males Juveniles Total

2006 62 27 30 119

2007 117 36 14 167

2008 180 18 26 224

2009 402 35 32 469

2010 290 29 195 514

2011 439 40 144 623

2012 552 109 105 766

2013 696 102 164 962

2014 475 56 211 742

2015 363 80 303 746

Total 3576 532 1224 5332

Page 88: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

88

Table 3.6: Top: the five best POPAN models according to AICc scores for the male-female only dataset. Bottom: the five best POPAN

models according to AICc scores for the male-female-juvenile dataset

Model AICc ∆AICc AICc Weight

M+F dataset

Φt,s pt,s PENTt Ns 8953.64 0.00 1.00

Φt,s pt PENTt Ns 9024.03 70.39 0.00

Φ(.),s pt PENTt Ns 9145.55 191.91 0.00

Φt,s pt,s PENTt,s Ns 9309.62 355.98 0.00

Φ(.),s pt PENTt Ns 9374.05 420.41 0.00

M+F+J dataset

Φt,s pt,s PENTt Ns 11691.04 0.00 1.00

Φ(.),s pt,s PENTt Ns 11967.82 276.74 0.00

Φ(.) pt,s PENTt Ns 12135.00 443.97 0.00

Φt,s pt PENTt Ns 12295.70 604.67 0.00

Φt,s pt,s PENTs Ns 12333.59 642.56 0.00

Page 89: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

89

Table 3.7: Top: Census population estimates (Nc) from POPAN using both datasets. M = male, F = female, J = juvenile. Values in

brackets denote 95% confidence intervals. Bottom: Effective population size estimates (Ne) for the male-female and male-female-

juvenile datasets.

Dataset Nc Males Nc Females Nc Juveniles Nc Total

M+F 943 [833 – 1068] 4533 [4375 -4697] NA 5476 [5208 – 5765]

M+F+J 936 [812 – 1080] 4680 [4511 – 4856] 1673 [1621 – 1727] 7289 [6944 – 7663]

Ne 95% CI

M+F 3120 2753 - 3534

M+F+J 3122 2799 - 3481

Page 90: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

90

Table 3.8: Summary of evolutionary processes and population dynamics incorporated by population genetic software. “Custom”

indicates the software allows the user to model the parameter in question.

* For estimates of Ne derived from linkage-disequilibrium.

Gene Flow Recombination Inbreeding Coalescent Mutation Population Models

MIGRATE Y/custom N N Y Y/custom Y/custom

DIYABC N Y/custom N Y Y/custom Y/custom

ONeSAMP N N Y Y Y N

NeEstimator N Y* N N N N

Page 91: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

91

Figure 3.1: Left: Sampling localities (diamonds) by genotypic clusters from Converse et

al (2015). Sampling localities located the PR genotypic cluster are listed in yellow font,

KI in blue, ICB in purple, and CoB in red. Poplar Island (PI) is found within the and is

denoted with a star. Right: Arial view of Poplar Island (PI) in 2013. Wetland cells are

located to the east and planned upland forests to the west. Poplar Harbor is outlined in

yellow, with Jefferson Island located to the north and Coaches Island to the south. Fyke

net trapping locations are shown in red.

Page 92: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

92

Figure 3.2: Twelve (12) MIGRATE models of Chesapeake Bay population structure.

ICBPat

CoB

KI

Full Matrix Central Conduit Central Sink

Tributary Trap Against FlowWith Flow

Coastal Escape Against Flow (Conduit)With Flow (Conduit)

Coastal Escape (Conduit) Tributary Trap (Conduit) No Flow

Page 93: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

93

Figure 3.3: Analysis flow for testing MIGRATE models. Each model was replicated five

times, and evaluated with Bézier scores, harmonic means, and log Bayes factor (LBF).

This initial analysis step was replicated a second time. The best models from these two

runs were then replicated 25 times and tested against one another using Bézier scores and

LBF.

Page 94: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

94

Figure 3.4: Effective population sizes in Chesapeake Bay. MIGRATE = blue; DIYABC =

red; ONeSAMP = purple; NeEstimator—linkage disequilibrium = black, sold;

NeEstimator—heterozygosity excess = black, long-dash; NeEstimator—molecular

coancestry = black, short-dash.

010

000

2000

030

000

4000

0

KIICB PR CoB

Effe

ctiv

e P

opul

atio

n S

ize

(Ne)

Population

Page 95: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

95

CHAPTER 4: TURTLE SOUP, PROHIBITION, AND THE POPULATION GENETIC

STRUCTURE OF DIAMONDBACK TERRAPINS

Introduction

Turtle soup was a popular food item in the United States during late nineteenth

and early twentieth centuries. Although many turtle species were consumed,

diamondback terrapins (Malaclemys terrapin) were considered a delicacy and were

highly sought. The historical market price for terrapins demonstrates their popularity: a

dozen larger terrapin sold for $70.00 during 1915-1920 (Coker 1920), or ~$830 today.

Recipes for turtle soup varied, but many contained sherry. However, in 1920 the United

States ratified the Eighteenth Amendment, banning the production, sale, and transport of

alcoholic beverages (Prohibition). The sherry used to make turtle soup became difficult to

procure, and demand for turtle soup plummeted. After the market crashed, several

terrapin farms purportedly dumped their stocks into local waters.

Prior to Prohibition, intense demand for turtle soup resulted in the decline of

terrapins across large portions of their range, particularly along the Atlantic seaboard

(Kennedy 2017). Due to their larger size, female terrapins were preferred (Coker 1920).

To combat extirpation and supplement the turtle soup market, terrapin farms were

established by private businesses and governmental agencies to explore domestication

and cultivation (Hay 1917; Hildebrand & Hatsel 1926; Hildebrand 1929, 1933).

Terrapins from the mid- and north Atlantic were preferred for their superior taste and

size. In particular, terrapins from Chesapeake Bay were the favorite and were sold with

the moniker “Chesapeakes” (Hay 1917; Coker 1920). In 1891, approximately 40,000 kg

Page 96: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

96

of terrapin were harvested from Chesapeake Bay alone, but by 1920 terrapin harvests

plummeted to ~370 kg (Cook 1989). As wild stocks dwindled, terrapins were

translocated from other regions in an effort to maintain stock solvency (Coker 1920).

Terrapins from North and South Carolina (“Carolinas”) were frequently shipped to other

states, while terrapins from Florida were derided as too small and insipid in flavor (Hay

1917). Terrapins from Texas achieved a desirable size but lacked the flavor of

“Chesapeakes” and “Carolinas,” and some terrapin farms hybridized terrapins with the

goal of creating a quick-growing, yet flavorful, terrapin (Hildebrand 1933). Hybridization

experiments between terrapins from Texas and the Carolinas began in 1914 at the U.S.

Fisheries Biological Station in Beaufort, North Carolina, and the first hybrid individuals

were confirmed in 1919 (Hildebrand 1933). However, Prohibition followed these

hybridization experiments. As terrapin farms along the Atlantic coast closed, they

purportedly released their mixed stocks into local waters. The population admixture that

resulted is poorly known.

Previous work has detected hints of the influence of translocation on patterns of

terrapin genetic diversity (Converse and Kuchta 2017). Terrapins from South Carolina

were shown to be more similar to Texas populations, while Florida populations were

identified as genetically distinct (Hauswaldt & Glenn 2005; Hart et al. 2014; Drabeck et

al. 2014). In addition, dramatic increases in contemporary gene flow into Chesapeake

Bay were interpreted to be the result of translocation (Converse et al. 2015).

Alternatively, documented patterns of genetic diversity could be due to natural

movements. For example, the Suwannee Seaway is a hypothetical embayment that

Page 97: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

97

existed during the early Miocene and late Pliocene (Bert 1986; Webb 1990; Randazzo &

Jones 1997; Lee & Ó Foighil 2003). This seaway ran through the northern part of

present-day Florida, potentially facilitating gene flow between the Gulf and the Atlantic

while isolating populations in peninsular Florida (Converse and Kuchta 2017).

Here I use microsatellite data to examine patterns of genetic variation across the

terrapins’ range and report evidence of human-mediated gene flow that is consistent with

historical accounts of terrapin translocation. I accomplish this by: i) inferring population

structure in a Bayesian framework, as well as with discriminate analysis of principal

components (DAPC); ii) estimating both historical and contemporary levels of gene flow

and comparing them to infer changes in gene flow over time; iii) testing alternative

models of population structure, including stepping-stone models and models that include

translocation events; and iv) surveying for population bottlenecks. Based on previous

studies and historical documentation, I hypothesized: i) admixture between Atlantic and

Gulf populations, but not with Florida populations; ii) some non-adjacent populations

(e.g. Texas and New Jersey) would show increases in contemporary gene flow, consistent

with translocation; iii) population genetic models incorporating translocation would

outperform other models; and iv) genetic tests for bottlenecks would fail to detect

population contractions in populations known to have experienced serious decline, due to

human-mediated influxes of genetic diversity.

Page 98: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

98

Methods

Sampling and Population Structure

I analyzed six polymorphic microsatellite loci (Hauswaldt and Glenn 2003) from

12 sampling localities throughout the Gulf and Atlantic seaboards (Figure 4.1A), a

dataset donated by Susanne Hauswaldt. This dataset includes sampling from Texas (TX),

Florida (FL), South Carolina (SC), North Carolina (NC), Maryland (MD), New Jersey

(NJ), and New York (NY) during 2001 (Figure 4.1A). Tests of allelic richness,

heterozygosity, and number of alleles, as well as lab procedures and PCR protocols are

available in Hauswaldt and Glenn (2005).

I demarcated coastal population structure with two methods. The first was a

Bayesian analysis in the program STRUCTURE v. 2.3 (Pritchard et al. 2000). The second

was a discriminate analysis of principal components (DAPC) in the R package adegenet

v. 1.4-2 (Jombart 2008). STRUCTURE delineates genetic clusters by maximizing

conformity to Hardy-Weinberg equilibrium while simultaneously minimizing linkage

disequilibrium among loci for K user-defined clusters, while DAPC utilizes k-means to

maximize variation between groups after PCA transformation. For STRUCTURE, I ran K

from 1 to 12 (sampling locations), replicating each value of K ten times, each with a

random starting seed. Individual runs were composed of a Markov chain Monte Carlo

(MCMC) algorithm of 700,000 steps, with the first 50% removed as burn-in. I used the

admixture model, the LOCprior, and LOCISPOP prior for all runs. Preferred values of

∆K were computed with the Evanno method (Evanno et al. 2005) in STRUCTURE

HARVESTER v. 0.6.94 (Earl & vonHoldt 2011). Output from STRUCTURE HARVESTER

Page 99: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

99

was processed in CLUMPP v. 1.1.2 (Jakobsson & Rosenberg 2007) to control for labeling

switching and multimodality. DISTRUCT v. 1.0 (Rosenberg 2003) was used to visualize

the data. I ran STRUCTURE in a hierarchical fashion: first I ran an initial analysis to

detect basal levels of structure (Evanno et al. 2005), and then I searched for additional

structure within each identified population individually.

For DAPC, I first optimized the number of PC axes to avoid under- or over-fitting

my population genetic models. I accomplished this with cross-validation, which uses

stratified random sampling and divides the dataset into a training set and a validation set.

I partitioned the training set to be 90% of the data and the validation set to 10%, and

employed 100 replicates. The cross-validation method estimated 60 axes should be

retained. The Bayesian information criterion (BIC) was used to choose the optimum

number of clusters for DAPC analysis. Four discriminate functions (DF) were retained

for each analysis.

Historical and Contemporary Gene Flow

I used MIGRATE v. 3.6.5 (Beerli 2009) to estimate historical gene flow levels (M

= mh/µ: proportion of migrants per generation, scaled by mutation rate). For these

analyses I treated sampling localities as populations with the exception of South Carolina.

Because South Carolina lacked structure among its sampling localities (Figure 4.1G; see

results), I treated all South Carolina sampling localities as a single population. I randomly

subsampled 40 individuals from each population and slice sampled the posterior

distribution with a MCMC of 5,000,000 steps with the first 50% removed as burn-in.

Each MCMC consisted of four statically heated parallel chains sampled every 500

Page 100: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

100

iterations. Five independent replicates were run, totaling 25,000,000 MCMC steps.

Estimates of θ (=4Neµ: effective population size, scaled by mutation rate) were modeled

under a uniform distribution bounded between 0.0001 and 100, and historical gene flow

estimates were bounded between 0 and 2000.

For estimates of contemporary gene flow (m: proportion of migrants per

generation), I used BAYESASS v. 3.0 (Wilson & Rannala 2003). I ran 10 independent

MCMC simulations with random starting seeds (Faubet et al. 2007) for 30,000,000 steps,

and sampled every 3,000 steps. I discarded the first 50% as burn-in. I used TRACER v.

1.5 (Rambaut & Drummond 2007) to visualize MCMC simulations, and used R scripts to

calculate a Bayesian deviancy measure to determine the run that best fit the data

(Spiegelhalter et al. 2002; Meirmans 2014).

Temporal Changes in Gene Flow

Since MIGRATE uses the coalescent, it estimates gene flow over long periods of

time, approximately 4Ne generations (several thousand years) in the past (Beerli 2009).

Thus, its gene flow estimates pre-date translocation events in the early twentieth century.

By contrast, BAYESASS estimates gene flow “…over the last several generations”

(Wilson & Rannala 2003), where “several” is commonly interpreted as roughly five

generations (e.g. Chiucchi & Gibbs 2010; Converse et al. 2015). With a generation time

of 12 years (Roosenburg, unpublished data), the contemporary gene flow estimates of

BAYESASS are within the last 60 years or so. These estimates post-date known terrapin

translocation events.

Page 101: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

101

To estimate temporal changes in gene flow, I compared historical and

contemporary estimates. First, I multiplied the historical gene flow estimates from

MIGRATE (M = mh/µ) by a mutation rate (µ) of 2.72 x 10-3. This mutation rate was

estimated explicitly for the microsatellites used by Hauswaldt (2004). The resulting

values are subtracted from the contemporary gene flow estimates (m) generated by

BAYESASS. Thus, ∆m = m - mh. Positive values of ∆m indicate increased levels of

contemporary gene flow, negative values indicate reduced contemporary gene flow, and

values near zero indicate no temporal change in gene flow.

Testing Coastal Population Structure

I compared eight models of gene flow in MIGRATE, generated with the methods

described above (Figure 4.2). My initial model was a linear stepping-stone (Model A),

which restricted gene flow to adjacent populations. I then added or removed gene flow

routes from Model A based on results from STRUCTURE, DAPC, and my ∆m estimates.

Models B-F incorporated gene flow from translocation events with varying connectivity

to Florida. Models G and H modeled the Suwannee Seaway. I used approximate Bézier

scores and log Bayes factors (LBF) to determine which model best explained coastal

population structure.

Bottlenecks

BOTTLENECK v. 1.2.02 (Piry et al. 1999) was used to test for bottlenecks. Locus

mutation was modeled with the stepwise-mutation model (SMM) and the two-phase

model (TPM). The TPM modeled 95% of mutations as single-step while multi-step

mutations were modeled with 12% variance (Piry et al. 1999). Each test consisted of

Page 102: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

102

20,000 permutations. I tested for bottlenecks by sampling locality and by STRUCTURE

cluster. I used the Wilcoxon signed-rank test, which detects bottlenecks 25 - 250

generations in the past (Cornuet & Luikart 1996), and a mode-shift test, which detects

bottlenecks “…within the past few dozen generations” (Luikart & Cornuet 1998).

Results

Population Structure — STRUCTURE

Initial runs of STRUCTURE found ∆K = 2 best describes coastal population

structure (Figure 4.1B). However, ∆K scores of 4 and 7 are comparable, and I also report

them for comparative purposes (Figure 4.1C,D). As ∆K identifies basal levels of

hierarchical structure (Evanno et al. 2005), I also tested for substructure within groups for

∆K = 2, which included northern and southern groups. Inference of population

substructure within the northern group (NC, MD, NJ, NY) revealed K = 2 or K = 3

subclusters (Figure 4.1E,F). This is due to two solutions of K having overlapping

likelihood scores. If K = 2 is adopted, NJ and NY constitute a cluster (Figure 4.1E), while

if K = 3 is adopted, they form separate subclusters (Figure 4.1F). The southern group

(TX, FL, SC) included ∆K = 3 subclusters (Figure 4.1G), one for each state. I did not

detect substructure among SC sampling localities. Thus, in total there are either five or

six genotypic clusters, dependent upon K = 2 or K = 3 for north Atlantic terrapins.

Population Structure — DAPC

BIC indicated the optimum number of clusters is six (Figure 4.3). DAPC for the

six-cluster analysis is shown in Figure 4.4, and membership probabilities are provided in

Figure 4.5. Mid- and north Atlantic terrapins form clusters 1, 3, and 6; cluster 4 diagnoses

Page 103: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

103

Chesapeake Bay (MD) terrapins, and cluster 2 includes populations located in the Gulf

(TX, FL). Cluster 5 is composed of SC terrapins and links cluster 4 with clusters 1, 3, and

6. Cluster 2 exhibits no overlap with any cluster.

If DAPC is run on sampling localities (Figure 4.1A), an alternative population

genetic structure is delimited that resembles the spatial distribution of my sampling

localities (Figures 4.6, 4.7). Clusters 1-6 (SC1-6) overlap heavily. Cluster 7 (NC)

deviates slightly from clusters 1-6. Cluster 8 (MD) overlaps with cluster 7 and clusters 1-

6. Cluster 9 (NJ) falls above cluster 8, and cluster 10 (NY) falls above cluster 9. Cluster

11 (FL) exhibits no overlap with any cluster, while cluster 12 (TX) overlaps with SC

populations.

Historical and Contemporary Gene Flow

The highest levels of historical gene flow were from NJ to NC (Table 4.1; mh =

0.0775) and the lowest were from NY to FL (mh = 0.0128). Contemporary gene flow

estimates show SC to NC exhibited the highest levels of gene flow (Table 4.2; m =

0.1497), while SC to FL demonstrated the lowest levels (m = 0.0037).

Of the 42 gene flow routes, six exhibited increased levels of contemporary gene

flow (Table 4.3; +∆m > 0.010), 22 demonstrated reduced levels of contemporary gene

flow (-∆m < -0.010), and 14 were relatively stable over time (-0.010 < ∆m < 0.010). Four

of the six gene flow routes estimated to have increased contemporary gene flow are found

between adjacent populations. The two routes between non-adjacent populations to show

higher levels of contemporary gene flow are from TX to SC (∆m = +0.0567) and from

NC to NY (∆m = +0.0320).

Page 104: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

104

Models of Population Structure

As the diamondback terrapin has a linear distribution (Figure 4.1), I used a

stepping stone process as my null model of population connectivity (Figure 4.2A). To

model translocation events, I modified the stepping stone model to include unidirectional

gene flow between TX and SC, and NC and NY; these models differed only in their

relative isolation of FL (Figure 4.2B-F). I modeled the Suwannee Seaway by modeling

bidirectional gene flow between TX and SC, with FL either completely isolated (Figure

4.2G) or a sink population (Figure 4.2H). I found that the Atlantic Exchange model

(Figure 4.2E), which modeled translocation events from TX to SC and NC to NY (gene

flow between non-adjacent populations), and which included bidirectional gene flow

between SC and FL but unidirectional gene flow between TX and FL, vastly

outperformed all other models, including the linear stepping-stone model (Bézier scores =

-13,724.32 vs. -72,456.30). Converting approximate Bézier scores into posterior

probabilities confirmed the Atlantic Exchange model as the best-fit model (PP = 100%).

All other models had estimated posterior probabilities near 0%.

Bottlenecks

I recovered no evidence for genetic bottlenecks for any sampling locality with the

TPM or SMM, although the TPM for SC6 approached significance (Table 4.4; P =

0.055). Mode-shift tests also failed to detect bottlenecks for any sampling locality. In

addition, neither the TPM, SMM, nor a mode-shift test detected bottlenecks for any

STRUCTURE cluster (Table 4.4).

Page 105: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

105

Discussion

Over the last two centuries, the relationship between terrapins and humans has

been complex, and the terrapin’s population genetic structure reflects this relationship.

The demand for turtle soup resulted in historical population contractions and extirpations,

and culminated in the construction of terrapin farms (Hay 1917; Hildebrand and Hatsel

1926; Hildebrand 1929, 1933). To get flavorful terrapins to market quickly, Texas and

Carolina terrapins were hybridized at the North Carolina terrapin farm (Hildebrand

1933). Then, in 1920, the enactment of Prohibition restricted access to sherry, which

drastically cut demand for turtle soup. Consequently, many terrapins were released into

local waters, which promoted population admixture and may have resulted in the

reintroduction of genetic diversity.

I documented population genetic structure that is consistent with historical

accounts of terrapin translocation during the twentieth century (Hay 1917; Coker 1920). I

recovered two or three genotypic clusters in the mid- and north Atlantic (MD, NJ-NY)

and three genotypic clusters in the Gulf and southern populations (TX, FL SC), for a total

of six genotypic clusters (Figure 4.1B-G; Figure 4.3). I also recovered the North

American Gulf/Atlantic phylogeographic divide previously described in the terrapin and

other taxa (Figure 4.1C; Avise et al. 1992; Lamb & Avise 1992; Hauswaldt 2004; Soltis

et al. 2006; Converse and Kuchta 2017).

In addition to delineating population structure and quantifying population

connectivity, I also found that a modified stepping-stone model with genetic exchange

along the Atlantic seaboard and unidirectional gene flow from TX to SC and from NC to

Page 106: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

106

NY best describes terrapin population structure (Figure 4.2E). This model of population

connectivity outperformed a linear stepping-stone model, as well as models of the

Suwannee Seaway, a natural conduit of gene flow between Gulf and Atlantic populations

(Figure 4.2). In addition, I found Florida populations to be divergent from neighboring

populations (Figure 4.1B-D, G, Figures 4.4, 4.6, Table 4.3), which complements accounts

that Florida terrapins were not translocated due to their inferior size and taste (Hay 1917),

as well as other studies that reported FL terrapins to be genetically distinct (Hauswaldt

and Glenn 2005; Hart et al. 2014; Drabeck et al. 2014).

My Bayesian model comparisons supported bidirectional gene flow between SC

and FL, and unidirectional gene flow from TX to FL (the Atlantic Exchange model,

Figure 4.2E), but did not support alternative models of connectivity, such as the

Suwannee Seaway (Figure 4.2).

I found that NC is highly admixed with other mid- north Atlantic populations

(Figure 4.1C-F), which could be the result of human-mediated gene flow. For example,

relative to historical levels of gene flow, I documented increased contemporary

connectivity from NC to NY (Table 4.3), which is consistent with accounts of

translocation (Coker 1920). North Carolina was the location of a terrapin breeding

operation (Hildebrand 1926, 1929, 1933), while NY was the location of a large terrapin

market (Coker 1920). With the volume of terrapins brought to market, it is possible some

terrapins escaped or were released into local waters. North Carolina also exhibited

increased contemporary gene flow into SC and Chesapeake Bay, MD (Table 4.3),

consistent with a previous study that detected large increases of contemporary gene flow

Page 107: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

107

into Chesapeake Bay (Converse et al. 2015). Finally, I observed admixture between TX

and SC populations (Figure 4.1D,G; Figure 4.6), as well as increased levels of

contemporary gene flow from TX into SC (Table 4.3), consistent with chronicled

translocation and hybridization experiments (Coker 1920; Hildebrand 1933; Converse

and Kuchta 2017).

Although I documented population genetic evidence of translocation between

some non-adjacent populations, my study failed to find genetic evidence for some known

instances of translocation. In particular, I did not detect increases in contemporary gene

flow or admixture between TX and NC (Table 4.3, Figure 4.1B-D, Figures 4.4, 4.6).

There are several possible explanations: translocation can fail (Short et al. 1992; Dodd &

Seigel 2002), and released terrapins from TX may not have successfully interbred with

local populations in NC. Alternatively, my sampling in NC may have not included

admixed localities (Figure 4.1A). I also did not find increased contemporary gene flow

from MD to NY, although I detected admixture between these populations (Table 4.3,

Figure 4.1B-F, Figures 4.4, 4.6). Terrapins from MD may not have been released into

local waters in NY, or they may have failed to interbreed. Despite current weak demand

for turtle soup, terrapins from the Chesapeake Bay region continue to be sold at markets

in New York, often illegally (Lester 2007; Roosenburg et al. 2008).

It is well documented that terrapin populations historically underwent severe

contractions (Kennedy 2017), but I failed to detect any population bottlenecks in any

region (Table 4.4). For example, the decline of terrapins from Chesapeake Bay is

documented by shrinking harvests after decades of overexploitation (Cook 1989), but I

Page 108: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

108

did not detect bottlenecks in this region. Indeed, previous work in Chesapeake Bay

indicated terrapins exhibit relatively high levels of genetic diversity (Hauswaldt and

Glenn 2005; Converse et al. 2015). It is possible that I failed to detect population

bottlenecks because tests of heterozygosity excess have weak statistical power (Peery et

al. 2012). Alternatively, the failure to detect bottlenecks may be the consequence of

terrapin translocation events reintroducing genetic diversity into populations. If true, the

enactment of Prohibition may have inadvertently benefited the terrapin in two ways. The

first was the collapse of the turtle soup market, which slowed the harvesting of natural

populations. The second was the closure of terrapin farms and the release of translocated

individuals into local populations, which may have reintroduced genetic diversity and

increased population viability (Frankham et al. 2010).

Thus, my study shows that population genetic structure in the diamondback

terrapin possesses the signature of historical translocation events. Translocation among

natural populations is known to increase levels of population admixture and genetic

diversity (Puckett et al. 2013; Wagner et al. 2015). At least in some cases, increased

levels of genetic diversity save populations from the negative consequences of inbreeding

depression and lowered mean population fitness (Westemeier et al. 1998; Moritz 1999;

Storfer 1999; Souty-Grosset and Grandjean 2008; Doody et al. 2009). However,

translocated populations may exhibit high levels of genetic diversity but have low

effective population sizes, suggesting several population genetic metrics are required to

judge the efficacy of translocation (Olson et al. 2013). An alternative outcome of

translocation is that it may harm populations by causing outbreeding depression, harm

Page 109: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

109

locally adapted populations by moving them away from an adaptive peak (Laikre et al.

2010; Calsbeek et al. 2012), or introduce diseases (Maddocks et al. 2015). Nonetheless,

given the extent of the current biodiversity crisis (Worm et al. 2006; Pongsiri et al. 2009;

Bellard et al. 2012; Assis et al. 2016), including increasing population fragmentation

(Templeton et al. 2001; Epps et al. 2005; Banks et al. 2013; Barr et al. 2015),

conservation-oriented translocation has become increasingly pivotal to maintaining

population viability (Pérez et al. 2012).

My study shows that convoluted genetic histories can be disentangled with

modern population genetic tools and that translocation can leave an indelible fingerprint

in populations. Anthropogenic influences increasingly disrupt population dynamics

(Templeton et al. 2001; Epps et al. 2005; Epps and Keyghobad 2015; Converse et al.

2015); however, the indirect consequences of social and political activities are not always

predictable. My study suggests the population genetic structure in the Diamondback

Terrapin may be the byproduct of an interaction between market demand for turtle soup

during the late nineteenth and early twentieth centuries, followed by the enactment of

Prohibition in 1920, which resulted in the large scale release of captive terrapins into

local waters.

Page 110: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

110

Table 4.1: Historical gene flow (mh) estimates among coastal populations. Gene flow is

measured by the proportion of migrants per generation, ranging from 0.0 – 1.0

Table 4.2: Contemporary gene flow (m) estimates among coastal populations. Gene flow

is measured by the proportion of migrants per generation, ranging from 0.0 – 1.0

TX FL SC NC MD NJ NY

TX - 0.0295 0.0282 0.0235 0.0272 0.0228 0.0520

FL 0.0502 - 0.0224 0.0300 0.0243 0.0232 0.0262

SC 0.0424 0.0263 - 0.0489 0.0287 0.0294 0.0729

NC 0.0354 0.0207 0.0215 - 0.0221 0.0438 0.0254

MD 0.0180 0.0277 0.0719 0.0249 - 0.0451 0.0583

NJ 0.0606 0.0202 0.0421 0.0775 0.0493 - 0.0388

NY 0.0336 0.0128 0.0335 0.0374 0.0280 0.0214 -

TX FL SC NC MD NJ NY

TX - 0.0268 0.0849 0.0184 0.0191 0.0207 0.0197

FL 0.0207 - 0.0233 0.0181 0.0192 0.0172 0.0195

SC 0.0042 0.0037 - 0.0047 0.0054 0.0077 0.0051

NC 0.0139 0.0133 0.1497 - 0.0511 0.0245 0.0574

MD 0.0169 0.0080 0.0518 0.0101 - 0.0145 0.0129

NJ 0.0108 0.0109 0.0389 0.0154 0.0630 - 0.1125

NY 0.0142 0.0128 0.0324 0.0157 0.0286 0.0303 -

Page 111: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

111

Table 4.3: Temporal changes in gene flow (∆m) among all populations; bolded values denote increased gene flow between disjunct

populations.

TX FL SC NC MD NJ NY

TX - -0.0027 +0.0567 -0.0051 -0.0081 -0.0021 -0.0323

FL -0.0209 - +0.0009 -0.0118 -0.0051 -0.0060 -0.0067

SC -0.0382 -0.0026 - -0.0441 -0.0233 -0.0217 -0.0678

NC -0.0215 -0.0074 +0.1282 - +0.0290 -0.0193 +0.0320

MD -0.0911 -0.0197 -0.0201 -0.0148 - -0.0306 -0.0453

NJ -0.0498 -0.0093 -0.0032 -0.0621 +0.0137 - +0.0738

NY -0.0194 -0.0172 -0.0011 -0.0217 +0.0006 +0.0089 -

Page 112: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

112

Table 4.4: BOTTLENECK results by sampling locality and STRUCTURE cluster, and their

respective P-values; an “L-Shaped” distribution under a Mode-Shift test indicates a

bottleneck was not detected.

SMM TPM Mode-Shift

Sampling

Locality

SC1 0.578 0.344 L-Shaped

SC2 0.281 0.922 L-Shaped

SC3 0.422 0.945 L-Shaped

SC4 0.500 0.281 L-Shaped

SC5 0.656 0.500 L-Shaped

SC6 0.281 0.055 L-Shaped

NC 0.422 0.219 L-Shaped

MD 1.000 1.000 L-Shaped

NJ 0.922 0.719 L-Shaped

NY 0.945 0.922 L-Shaped

TX 0.719 0.781 L-Shaped

FL 0.961 0.961 L-Shaped

STRUCTURE

Cluster

SC 0.961 0.422 L-Shaped

MD 1.000 1.000 L-Shaped

NJ/NY/NC 0.961 0.719 L-Shaped

Page 113: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

113

Table 4.4: continued

TX 0.719 0.781 L-Shaped

FL 0.961 0.961 L-Shaped

Page 114: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

114

Figure 4.1: A) Black box: the distribution of the terrapin (red) along the North American

coastline. Main: sampling localities and number of samples for each location. B) All

terrapin populations under ∆K = 2. C) All terrapin populations under ∆K = 4. D) All

terrapin populations under ∆K = 7. Subsections B,C, and D depict the same analysis,

explored under different genotypic partitions. E) Mid- and north Atlantic terrapins under

K = 2. In this scenario, NY and NJ terrapins constitute a single population. F) Mid- and

north Atlantic terrapins under K = 3. In this scenario, NY and NJ form separate

populations. Subsections F and G depict the same analysis under two values of K with

similar likelihood scores. G) SC terrapins investigated with Gulf populations. TX and FL

form separate populations and SC1-6 constitute a single cluster. No substructure is

present within SC1-6. TX, FL, SC, and MD are diagnosable clusters. If NY and NJ form

a single cluster (subsection E), there are five total clusters. If NY and NJ form separate

clusters (subsection F), there are six total clusters.

TX

FL

SC1SC3SC2,SC4-6

NC

TX

FL

NY

NJ

MD

NC

SC6

SC5

SC4

SC3

SC2

SC1

∆K = 2 ∆K = 4 ∆K = 7

K = 2 K = 3

∆K = 3

MDNJ

NY

A B C D E F G

(n = 21)

(n = 26)(n = 56)

(n = 20)

(n = 14)

(n = 12) N

500 km

(n = 20)

(n = 42)

(n = 21)(n = 34, 26, 28)

Page 115: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

115

Figure 4.2: The eight MIGRATE models of population structure derived from

STRUCTURE, DAPC, and ∆m results. Solid lines denote naturally occurring gene flow

while dashed lines indicate gene flow possibly arising from translocation. Bézier scores

are found below model names. Model A is a linear stepping-stone; Model B added gene

flow from TX to SC and from NC to NY; Model C removed all gene flow to and from FL

while retaining translocation; Model D treated FL as a sink population; Model E allowed

gene flow with FL on the Atlantic coast; Model F allowed gene flow with FL on the Gulf

Coast; Model G depicts the Suwannee Seaway with completely isolated FL populations;

Model H depicts the Suwannee Seaway with FL as a sink population.

Page 116: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

116

Figure 4.3. BIC results indicating DAPC analyses are optimized at six clusters.

2 4 6 8 10 12

305

310

315

320

Value of BIC versus number of clusters

Number of clusters

BIC

Page 117: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

117

Figure 4.4: DAPC for 60 retained axes and four discriminate functions. Six clusters are

recovered with this model. The top half contains populations along the Atlantic coast

while Gulf populations are found on the right half of the diagram. SC terrapins are

present within each cluster.

Page 118: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

118

Figure 4.5: Assignment probabilities for individuals under K =6 (Figure 4.4). Warmer

colors indicate higher assignment probabilities for clusters.

SC1

SC2

SC3

SC4

SC5

SC6

NC

MD

NJ

NY

FLTX

SC

1 2 3 4 5 6

Cluster

Page 119: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

119

Figure 4.6: DAPC for 60 retained axes on sampling localities and four discriminate

functions. Clusters in the model resemble the spatial distribution of sampling localities

along the Gulf and Atlantic seaboards. (Figure 4.1). FL exhibits no admixture with any

cluster while TX shows overlap with populations found in SC.

Page 120: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

120

Figure 4.7: Assignment probabilities for individuals under K = 12 (Figure 4.6). Warmer

colors indicate higher assignment probabilities for clusters. Clusters 9 and 10 exhibit

uncertainty when assigning individuals to either NJ or NY, consistent with STRUCTURE

results (Figure 4.1E,F).

SC1

SC2

SC3

SC4

SC5

SC6

NC

MD

NJ

NY

FLTX

SC

1 2 3 4 5 6 7 8 9 10 11 12Cluster

Page 121: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

121

CHAPTER 5: SPLITTING DIAMONDS: DATING PHYLOGEOGRAPHIC

DIVERGENCE IN THE DIAMONDBACK TERRAPIN USING APPROXIMATE

BAYESIAN COMPUTATION.

Introduction

Divergence dating is an essential component of biogeographic inference. Because

most taxa lack a complete evolutionary chronicle in the fossil record, divergence dates

are commonly estimated using a molecular clock. However, divergence estimates from

molecular data are sensitive to the rate of nucleotide substitution (Hillis et al. 1996), and

rates of nucleotide substitution vary by gene and taxon, and the rate of evolution can

itself evolve (Thorne 1998; Kuchta et al. 2016). Additionally, different molecular clock

models generate different divergence date estimates, even within the same taxon.

Diamondback Terrapins (Malaclemys terrapin) are turtles that are found in

brackish water habitats along the eastern North American continent from Texas to

Massachusetts (Ernst and Barbour 1989). Terrapins are characterized by male-biased

dispersal and prefer shallow water, near shore environments (Roosenburg 1991; 1994;

Spivey 1998; Sheridan et al. 2010). Although seven subspecies are currently recognized,

terrapins lack strong genetic structure at the local and regional scales (Hauswaldt and

Glenn 2005; Glenos 2013; McCafferty et al. 2013; Hart et al. 2014; Drabeck and

Chatfield 2014; Converse et al. 2015; Petre et al. 2015; Converse and Kuchta 2017).

However, terrapins exhibit a genetic break near Cape Canaveral, Florida that coincides

with morphological variation in their carapace shape (Lamb and Avise 1992; Avise et al.

1992; Converse and Kuchta 2017). Moreover, this “Gulf/Atlantic” phylogeographic

Page 122: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

122

divide found in the terrapin is present in over 25 other taxa, including the Horseshoe Crab

(Limulus polyphemus), the Dusky Seaside Sparrow (Ammodramus maritimus), the

Hermit Crab (Pagurus longicarpus), and the Blacktip Shark (Carcharhinus limbatus)

(Saunders et al. 1986; Avise and Nelson 1989; Young et al. 2002; Keeney et al. 2005;

Soltis et al. 2006).

Although the terrapin shares a common phylogeographic pattern with other taxa,

it is divergent in its lack of diversity in its mitochondrial DNA (mtDNA; Lamb and Avise

1992; Avise et al. 1992; Converse and Kuchta 2017). In a classic phylogeographic study,

Avise et al. (1992) produced two hypotheses to explain this paucity of mtDNA variation.

The first hypothesis proposed that a slow rate of evolution in mtDNA resulted in less

accumulation of genetic diversity relative to other taxa in the Gulf region (Lamb and

Avise 1992; Avise et al. 1992). Under this hypothesis, terrapin mtDNA exhibits a 14-fold

decrease in evolutionary rate relative to other vertebrate taxa, and terrapin populations

diverged into Gulf and Atlantic clusters when most other taxa did, approximately 350,000

- 1,100,000 years ago. (Lamb and Avise 1992; Avise et al. 1992). A slow rate of

molecular evolution has been documented in the Painted Turtle (Chrysemys picta) using

complete genome sequencing (Shaffer et al. 2013) and initial phylogeographic work on

the terrapin revealed extremely low nucleotide diversity (Hauswaldt 2004). Thus, the

hypothesis of slow molecular evolution in the terrapin (Lamb and Avise 1992; Avise et

al. 1992) has received recent empirical support. The second hypothesis by Avise et al.

(1992) proposed that terrapin mtDNA evolved at the “conventional” vertebrate rate, and

terrapin populations thus diverged approximately 50,000 years ago. The estimates from

Page 123: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

123

these hypotheses are based on a shared vertebrate mtDNA clock of 2% myr-1, which has

been much debated (Avise et al. 1992; Hillis et al. 1996; Yang 2014). In this article, I

revisit the competing phylogeographic scenarios for the terrapin postulated by Lamb and

Avise (1992) and Avise et al. (1992). I use microsatellite data and approximate Bayesian

computation (ABC) to evaluate three models of terrapin phylogeographic divergence.

From these models, I estimate divergence dates and effective population sizes, and

address hypotheses of slower rates of mtDNA evolution in the Diamondback Terrapin.

Materials

Sample Collection and Microsatellite Preparation

Susanne Hauswaldt donated a dataset composed of 12 localities along the Atlantic

and Gulf seaboards of the United States, including a total of 320 terrapins (Figure 5.1).

Terrapins were collected from ACE Basin, South Carolina (SC1), the Ashley River

(SC2), Charleston Harbor (SC3), the Wando River (SC4), the Cooper River (SC5), Cape

Romain (SC6), Beaufort, North Carolina (NC), the Patuxent River, Maryland (MD),

Stone Harbor, New Jersey (NJ), Oyster Bay, New York (NY), the Florida Keys (FL), and

Nueces Bay, Texas (TX). DNA from blood, tail tips, or leg muscle was extracted as

described in Hauswaldt and Glenn (2005). I then screened samples for six polymorphic

microsatellite loci using protocols described in Hauswaldt and Glenn (2003) and

Hauswaldt and Glenn (2005). Population genetic metrics including allelic richness,

heterozygosity, and number of alleles are detailed in Hauswaldt and Glenn (2005).

Page 124: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

124

Coastal Population Structure

STRUCTURE version 2.3 (Pritchard et al. 2000) was used to demarcate genotypic

clusters (K) among my sampling localities. I ran STRUCTURE from K = 1 to K = 12, with

each value of K replicated 10 times with different and random starting seeds. Individual

runs were composed of a Markov Chain Monte Carlo (MCMC) of 700,000 iterations,

with 350,000 MCMC iterations discarded as burn-in. I used the admixture model,

correlated allele frequencies model, inferred α, LOCprior, and LOCISPOP prior for all

runs. STRUCTURE HARVESTER web version 0.6.94 (Earl and vonHoldt 2011) was used

collate STRUCTURE results and compute ∆K using the Evanno method (Evanno et al.

2005). Preferred values of ∆K were analyzed in CLUMPP version 1.1.2 (Jakobsson and

Rosenberg 2007), to account for label switching and multimodality, and its output was

visualized in DISTRUCT version 1.1 (Rosenberg 2003).

Divergence Dating — Microsatellite Modeling

I used approximate Bayesian computation (ABC; Beaumont et al. 2002) to

estimate divergence dates between Gulf and Atlantic populations using DIYABC version

2.04 (Cornuet et al. 2014). I modeled individual locus mutation rates with gamma

distributions bounded between 1 x 10-5 and 1 x 10-2 mutations site-1 generation-1 with

shape 2, and modeled the mean microsatellite mutation rate with a uniform distribution

bounded between 1 x 10-4 and 1 x 10-3. A generalized stepwise model was used for

microsatellite mutation, which allows repeats to increase or decrease by one or more

motifs under a geometric distribution (Cornuet et al. 2008). For parameter estimates,

Page 125: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

125

scenario selection, and confidence testing, I used 12 one-sample and nine two-sample

microsatellite summary statistics.

Divergence Dating — Range Scenarios and Priors

To test the phylogeographic hypotheses from Lamb and Avise (1992) and Avise

et al. (1992), I used the three most discrete genotypic clusters recovered by STRUCTURE

(MD, FL, TX; Figure 5.1) to simplify phylogeographic scenarios and decrease

computation time. These locations were also selected based on sample size and lower

levels of population admixture (Haudwaldt and Glenn 2005; this study). I converted

divergence date estimates from generations to years with a generation time of 12 years

(Converse et al. 2015; Roosenburg unpubl. data). I estimated all parameters under the

ABC-regression algorithm (Beaumont et al. 2002). In total, I ran two separate

phylogeographic analyses using DIYABC. The first analysis was composed of a single

scenario with the Gulf/Atlantic divergence date modeled with a uniform prior bounded

between 0 and 1,100,000 years (Figure 5.2A; Avise et al. 1992). Effective population

sizes were modeled under uniform distributions bounded between 10 and 40,000, as

determined by preliminary analyses. From this initial model, I generated 1,000,000

simulated data points, retained the closest 0.01% of simulated points, and applied

polychotomous logistic regression to the retained data. Estimates for divergence dates

and effective population sizes were logit transformed and quantified under the original

parameter choice option (Cornuet et al. 2014).

For my second analysis, I took the estimates of the mode (“intermediate

divergence”) and 5% and 95% confidence intervals (“young” and “old” divergences,

Page 126: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

126

respectively) from the first analysis and centered Gaussian distributions around each, for

a total of three models (Figure 5.2B). I parameterized standard deviations to be equal to

20% of each scenario’s divergence estimate and bounded each distribution by two

standard deviations. The divergence date between Texas and Florida was modeled with a

uniform prior bounded between 0 and the divergence estimate from each scenario (Figure

5.2B). For these analyses, I simulated 3,000,000 data points and modeled all remaining

parameters as indicated in the inital analysis. I estimated the posterior probabilities of the

scenarios and their model parameters by retaining the closest 0.01% of simulated data.

Divergence Dating — Scenario Confidence and Scenario Checking

To evaluate the effectiveness of my ABC analyses to discern among competing

scenarios, I generated pseudo-observed datasets (pods). Each analysis consisted of 1,000

pods with parameter values drawn from prior distributions in previous ABC analyses.

These pods generate relative posterior probabilities amongst the three scenarios. Using

these relative posterior probabilities, I estimated type 1 and type 2 errors amongst

competing scenarios. Type 1 error arises when the “true” scenario is rejected in favor of a

“false” scenario, while type 2 error arises when a “false” scenario is selected over the

“true” scenario. I estimated type 1 and type 2 errors for all possible model comparisons.

Mean type 2 error was calculated over each respective scenario.

Finally, I used the model checking function in DIYABC to evaluate the goodness-

of-fit for the preferred scenario. This function evaluates the consistency of the posterior

parameter distributions from the preferred scenario with the observed dataset. For the

model to be deemed accurate, the observed statistics must fall within the distribution of

Page 127: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

127

simulated points under the preferred scenario (Cornuet et al. 2010). Following the

recommendations of Cornuet et al. (2010), I used a new suite of summary statistics to

evaluate my model’s goodness-of-fit, and avoided summary statistics from previous

analyses for parameter estimation or model selection.

Results

Coastal Population Structure

STRUCTURE recovered two major clusters (K=2) along the Gulf and Atlantic

coasts (Figure 5.1). The “Atlantic” cluster consisted of populations in the mid- and north

Atlantic, while the “Gulf” cluster was composed of populations from South Carolina and

the Gulf of Mexico. Analysis of substructure within the Atlantic cluster recovered either

two or three clusters, with Maryland as the most discrete (Figure 5.1). Evaluation of the

Gulf cluster populations recovered Texas, Florida, and South Carolina are discrete

entities. No substructure was present within South Carolina populations. Based on these

results, I selected Maryland (MD), Texas (TX), and Florida (FL) as target populations for

downstream analyses. Thus, Atlantic populations are represented by Maryland, and Gulf

populations by Florida and Texas.

Divergence Dating and Effective Population Size

My initial ABC scenario (Figure 5.2A) estimated a Gulf/Atlantic divergence of

157,200 years ago (95% CI = 65,760 - 775,200; Figure 5.2A). Texas and Florida are

estimated to have diverged approximately 39,840 years ago [26,040 - 442,800]. For my

second analysis (Figure 5.2B), I constructed three models and respectively centered

Page 128: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

128

Gaussian distributions around 157,200 years ago (mode), 65,760 (5% CI), and 775,200

(95% CI) as described in the methods.

Results from my second analysis are presented in Table 5.1. Of the three

competing scenarios, ABC posteriors indicated the intermediate divergence scenario was

the preferred model (PP = 66.30%). The young divergence scenario was the next best (PP

= 30.85%) followed by the old divergence scenario (PP = 2.85%). Under the intermediate

divergence scenario, the Gulf/Atlantic split occurred 138,000 years ago [106,800 -

195,600]. The young divergence scenario estimated the split at 63,960 years ago [36,600

- 86,520], and the old divergence scenario estimated the split at 669,600 years ago

[512,400 - 938,400]. Under the preferred scenario, Maryland exhibits an effective

population size of 30,300 individuals [17,900 - 38,800], Texas exhibits an effective size

of 15,700 individuals [6,040 - 35,600], and Florida 4,770 individuals [2,710 - 11,700].

Type 1 and Type 2 error summaries are found in Table 5.2. Error was highest

between the intermediate and young divergence scenarios, as these models were closer to

one another in parameter space. Error was lowest when comparing the young and old

divergence scenarios, as these models were distant in parameter space.

Microsatellite Mutation

ABC estimates for the preferred intermediate divergence scenario recovered an

overall mutation rate of 2.70 x 10-4 mutations site-1 generation-1 [1.66 - 7.61 x 10-4]. The

young divergence scenario estimated the mutation rate to be 5.24 x 10-4 [2.75 - 9.44 x 10-

4], and the old divergence scenario estimated the rate to be 1.26 x 10-4 [1.08 - 4.08 x 10-4].

Page 129: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

129

Discussion

Lamb and Avise (1992) and Avise et al. (1992) recovered unusually low levels of

mtDNA diversity in turtles and used phylogeographic inference to estimate slower rates

of molecular evolution in Testudines. These seminal contributions delineated a link

between molecular level processes and range-wide phylogeographic structure among

disparate taxa. Moreover, these studies helped demonstrate that observation at the

phylogeographic level could contribute to understanding molecular level phenomena.

Using molecular data and phylogeographic models evaluated using ABC, my study

suggests that Gulf and Atlantic terrapin populations did not diverge 350,000 - 1,100,000

years ago, as hypothesized by Lamb and Avise (1992) and Avise et al. (1992).

Consequently, my study casts further doubt on the postulated 14-fold slowdown in

terrapin mtDNA molecular rate (Avise et al. 1992). However, my results suggest the

terrapin has some degree of decreased rate of molecular evolution, which is consistent

with a sister taxon (Painted Turtle) that exhibits a slower rate of molecular evolution

(Shaffer et al. 2013). In addition, the terrapin’s nuclear and mitochondrial genomes have

been sequenced as of late 2015 (Mihai Pop, pers. comm.). Once available, these data will

help shed light on the terrapin’s rate of molecular evolution relative to other vertebrate

taxa and help elucidate its phylogeographic history.

There are several explanations for the Gulf and Atlantic phylogeographic divide.

According to historical estimates, sea levels were approximately 80 meters lower than

contemporary levels 150,000 years ago (Grant et al. 2012). This period of lower sea

levels coincides with the “intermediate divergence” model. Similarly, sea levels were

Page 130: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

130

approximately 80 meters below contemporary levels around 65,000 years ago (Grant et

al. 2012). This change in sea level coincides with the “young divergence” model. Lower

sea levels would have increased the perimeter of peninsular Florida, possibly to the

Florida Shelf. As a consequence, the dispersal distance between Gulf and Atlantic

populations would have increased, leading to deceased genetic exchange.

Alternatively, the Gulf/Atlantic genotypic divide may be an indirect consequence

of the Gulf Stream (Avise 2000). North of Cape Canaveral, Florida, the Gulf Stream

deviates away from the Atlantic coast and towards the Atlantic Ocean, which may

impede dispersal between Gulf and Atlantic populations. It has been postulated that this

is the cause of the shared phylogeographic divergence among the terrapin and other

coastal taxa (Saunders et al. 1986; Soltis et al. 2006). A third explanation by Wise et al.

(2004) proposed that a combination of climate, currents, mangrove ecosystems, and

sediments contribute to adversely to prevent dispersal along the eastern Florida coastline.

These explanations are not mutually exclusive. It is possible that fluctuations in sea level

caused Gulf and Atlantic populations to diverge, while the Gulf Stream and features of

Florida’s eastern shoreline maintain a barrier to gene flow.

The incongruences between my divergence date estimates and those of Lamb and

Avise (1992) could be a byproduct of marker choice. Lamb and Avise (1992) assayed

mtDNA using restriction enzymes, while my study used nuclear microsatellites.

Incongruences between maternally-inherited, haploid mtDNA and biparentally-inherited,

diploid nuclear DNA are well documented (Hudson and Coyne 2002; Kuchta and Tan

2005, 2006; Schelly et al. 2006; Linnen and Farrell 2007; Kuchta et al. 2009; Kuchta et

Page 131: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

131

al. 2016). Likewise, variation in marker mutation rate may partially explain my

divergence date estimates. I estimated my loci to have an overall mutation rate of 2.70 x

10-4 mutations site-1 generation-1, much faster than estimates of Chelonian mtDNA

mutation of 2.01 x 10-9 (Eo and DeWoody 2010).

I obtained three divergence date estimates for terrapin phylogeography using

ABC analyses. The “intermediate divergence” model had the strongest support (PP =

66.30%) and suggests that the Gulf and Atlantic divide occurred ~138,000 years ago. The

“young divergence” model was the next best (PP = 30.85%), and estimated the Gulf and

Atlantic divide at 63,960 years ago. The “old divergence” model had the least support

(PP = 2.85%), and estimated the divergence to have occurred 669,600 years ago (Table

5.1).

The estimates generated by “intermediate divergence” model fell in between the

original divergence hypotheses of 50,000 or 350,000 - 1,100,000 years ago postulated by

Lamb and Avise (1992) and Avise et al. (1992). However, the “young divergence” model

estimated the Gulf and Atlantic divergence occurred 63,960 years ago, while the “old

divergence” model recovered a divergence date of 669,000 years ago. Both of these

estimates fell near or within the divergence windows postulated by Lamb and Avise

(1992) and Avise et al. (1992) (Figure 5.2). In addition, ABC estimates of effective

population size were congruent with geographic patterns for number of alleles, allelic

richness, and observed heterozygosity (Hauswaldt and Glenn 2005; Hart et al. 2014).

Maryland exhibited the largest effective population size and highest metrics for genetic

diversity, followed by Texas and Florida (Table 5.1).

Page 132: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

132

Type 1 error estimates indicated that the intermediate divergence scenario was

the most likely to be rejected when it was the preferred model (39.2%), while the old

divergence scenario was the least likely to be rejected when preferred (13.1%; Table 5.2).

This pattern was also true of type 2 error. In my analyses, higher type 1 and 2 error was

associated when models were fit with slightly overlapping priors (e.g. “intermediate

divergence” and “young divergence” models; Figure 5.2B), and lower when priors were

disparate (e.g. “young divergence” and “old divergence” models). While there is some

uncertainty when choosing between “young” and “intermediate” divergence models, the

“old divergence” model had poor support.

The paucity of mtDNA diversity in the Diamondback Terrapin presented a novel

problem for terrapin phylogeography, and resulted in two competing hypotheses with

implications on rates of molecular evolution (Lamb and Avise 1992; Avise et al. 1992;

Converse and Kuchta 2017). Using ABC, my study found that neither of the two original

divergence hypotheses were well supported, and instead indicates a divergence date of

~138,000 years ago. This divergence date suggests a faster rate of molecular evolution in

the terrapin than previously estimated and also reveals the complexity of the terrapin’s

phylogeographic history.

Page 133: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

133

Table 5.1: Summary of the three competing models for Gulf-Atlantic divergence dates. Values in brackets indicate 95% confidence

intervals. The intermediate model was centered on the mode from the initial analysis, the young model on the 5% CI, and the old

model on the 95% CI.

Model Posterior

Probability

Gulf/Atlantic Split

(years)

TX/FL Split (years) Ne MD Ne TX Ne FL

Intermediate 66.30%

[61.71 – 69.89%]

138,000

[106,800 – 195,600]

96,120

[44,000 – 146,400]

30,300

[17,900 - 38,800]

15,700

[6,040 - 35,600]

4,770

[2,710 - 11,700]

Young 30.85%

[23.49 – 38.21%]

63,960

[48,600 – 86,520]

41,880

[18,840 – 62,280]

27,000

[13,500 - 37,700]

6,210

[3,460 - 34,500]

2,420

[1,560 - 6,760]

Old 2.85%

[0 - 9.63%]

669,600

[512,400 – 938,400]

214,800

[88,560 – 648,000]

36,700

[27,600 - 39,600]

33,200

[15,300 - 38,200]

11,800

[5,340 - 21,400]

Page 134: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

134

Table 5.2: Type 1 and type 2 errors for all model comparisons. “among models” indicates

Type 1 and 2 errors when all models are evaluated simultaneously. “between models”

denotes Type 1 and 2 errors between the two models listed. The first model listed for

each comparison is assumed as the “true” model for error calculation.

Model Type 1 Error (%) Type 2 Error (%)

among models

Intermediate (mode)

39.2

16.6

Young (5% CI) 20.6 15.4

Old (95% CI)

between models

13.1 4.5

Intermediate/Young 27.0 24.5

Intermediate/Old 8.5 13.2

Young/Intermediate 24.5 27.0

Old/Intermediate 13.2 8.5

Young/Old 2.3 4.9

Old/Young 4.9 2.3

Page 135: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

135

Figure 5.1: Sampling localities and STRUCTURE results. Sample sizes are found on right

in parentheses and correspond to sampling locality. Populations used in downstream

ABC analyses are indicated in red. Faded into the background is a figure adopted from

Avise et al. (1992), with blue circles denoting their Gulf cluster and white circles their

Atlantic cluster. The genotypic and morphological transition near Cape Canaveral, FL, is

represented by a dashed line. Two genetic clusters (∆K=2) were initially identified, with

South Carolina populations found within the Gulf cluster. Texas, Florida, and South

Carolina were identified as subclusters. Within the Atlantic cluster, Maryland formed a

discrete entity while New Jersey and New York exhibited weak differentiation.

Page 136: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

136

Figure 5.2: ABC models for Gulf/Atlantic divergence scenarios. Time (thousands of

years) is found along the middle x-axis and applies to parts (A) and (B). A) Priors (dotted

lines) and posteriors (solid lines) for the initial model. Blue lines represent the FL/TX

divergence date and red lines denote the Atlantic/Gulf divergence date. Posteriors include

modes (squares) and 5% and 95% confidence intervals (vertical lines). B) The three

divergence scenarios constructed from the results from the initial model. All

Atlantic/Gulf divergence dates (red lines) were modeled using Gaussian distributions.

The top scenario is a Gaussian distribution centered on the mode from the initial analysis

(157,200; sd =31,440), the middle scenario is a Gaussian distribution centered on the 5%

confidence interval (65760; sd = 13,152), and the bottom scenario is a Gaussian

distribution centered on the 95% confidence interval (775,200; sd = 155,040). For each

scenario, the TX/FL divergence date (blue lines) was modeled with a uniform prior

bounded between zero and the mode for each Gaussian distribution. Ranges for the

Atlantic and Gulf divergence dates are indicated in brackets.

Page 137: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

137

LITERATURE CITED

Akaike, H. 1973. Information theory and an extension of the maximum likelihood

principle. In: Springer Series in Statistics – Selected Papers of Hirotuga Akaike

(eds. E. Parzen, K. Tanabe, and G. Kitagawa) pp. 199–213. Springer, New York.

Akaike, H. 1974. A new look at the statistical model identification, IEEE Transactions on

Automatic Control 19: 716–723.

Assis, J., A.V. Lucas, I .Bárbara, and E.Á. Serrão. 2016. Future climate change is

predicted to shift long-term persistence zones in the cold-temperate kelp Laminaria

hyperborea. Marine Environmental Research 113: 174–182.

Auger, P.J. 1989. Sex ratio and nesting behavior in a population of Malaclemys terrapin

displaying temperature-dependent sex-determination. Ph.D. dissertation, Tufts

University, Medford, MA, 174 pp.

Avise, J.C. 2000. Phylogeography: The History and Formation of Species. Harvard

University Press, Cambridge, MA. 464 pages.

Avise, J. C., and W. S. Nelson. 1989. Molecular genetic relationships of the extinct dusky

seaside sparrow. Science 243: 646–648.

Avise, J.C., B.W. Bowen, T. Lamb, A.B. Meylan, and E. Bermingham. 1992.

Mitochondrial DNA evolution at a turtle's pace: evidence for low genetic variability

and reduced microevolutionary rate in the Testudines. Molecular Biology and

Evolution 9: 457–473

Baker A.J., P.M. González, T. Piersma, et al. 2004. Rapid population decline in red

knots: fitness consequences of decreased refuelling rates and late arrival in

Page 138: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

138

Delaware Bay. Proceedings of the Royal Society B: Biological Sciences 271: 875–

882.

Banks, S.C., G.J. Cary, A.L. Smith et al. 2013. How does ecological disturbance

influence genetic diversity? Trends in Ecology & Evolution 28: 670–679.

Barney, R.L. 1924. Further notes on the natural history and artificial propagation of the

diamond-back terrapin. Bulletin US Bureau of Fisheries 38:91–111.

Barr, K.R., B.E. Kus, K.L. Preston, et al. 2015. Habitat fragmentation in coastal southern

California disrupts genetic connectivity in the cactus wren (Campylorhynchus

brunneicapillus). Molecular Ecology 24: 2349–2363.

Beaumont, M.A., W. Zhang, and D.J. Balding. 2002. Approximate Bayesian computation

in population genetics. Genetics 162: 2025–2035.

Beerli, P. 2008. Migrate version 3.6.11: a maximum likelihood and Bayesian estimator of

gene flow using the coalescent. Distributed over the internet at

http://popgen.scs.edu/migrate.html

Beerli, P. 2009. How to use MIGRATE or why are Markov chain Monte Carlo programs

difficult to use. In: Population Genetics for Animal Conservation (eds G. Bertorell,

M.W. Bruford, H.C. Hauffe, A. Rizzoli, and C. Vernesi), pp. 42–79, Cambridge

University Press, Cambridge.

Beerli, P. and M. Palczewski, 2010. Unified framework to evaluate panmixia and

migration direction among multiple sampling locations, Genetics 18: 313–326.

Beier, P., D.R. Majka, and W.D. Spencer. 2008. Forks in the road: choices in procedures

for designing wildlife linkages. Conservation Biology 22: 836–837.

Page 139: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

139

Bellard C, Bertelsmeier C, Leadley P, Thuiller W, Courchamp F (2012) Impacts of

climate change on the future of biodiversity. Ecology Letters, 15, 365–377.

Bert, T.M. 1986. Speciation in western Atlantic stone crabs (genus Menippe): the role of

geological processes and climatic events in the formation and distribution of

species. Marine Biology 93: 157–170.

Besbeas P., S.N. Freeman, B.J.T. Morgan, and E.A. Catchpole. 2002. Integrating mark–

recapture–recovery and census data to estimate animal abundance and demographic

parameters. Biometrics 58: 540–547.

Bowler, D.E., and T.G. Benton. 2005. Causes and consequences of animal dispersal

strategies: relating individual behaviour to spatial dynamics. Biological Reviews of

the Cambridge Philosophical Society 80: 205–225.

Braunisch, V., G. Segelbacher, and A.H. Hirzel. 2010. Modeling functional landscape

connectivity from genetic population structure: a new spatially explicit approach.

Molecular Ecology 19: 3664–3678.

Bromham. L. 2002. Molecular clocks in reptiles: life history influences rate of molecular

evolution. Molecular Biology and Evolution 19: 302–309.

Burnham, K.P. 1991. On a unified theory for release-resampling of animal populations.

In: Proceedings of 1990 Taipei Symposium in Statistics, (eds. M.T. Chao, and P.E.

Cheng), pp.11-36. Institute of Statistical Science, Academia Sinica. Taipei, Taiwan.

Butler, J.A. 2002. Population ecology, home range, and seasonal movements of the

Carolina diamondback terrapin, Malaclemys terrapin centrata in northeastern

Page 140: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

140

Florida. Florida Fish and Wildlife Conservation Commission, Tallahassee, FL, 72

pp.

Caizergues A., A. Bernard-Laurent, J.F. Brenot, L. Ellison, and J.Y. Rasplus. 2003.

Population genetic structure of rock ptarmigan Lagopus mutus in Northern and

Western Europe. Molecular Ecology 12: 2267–2274.

Callens, T., P. Galbuser, E. Matthysens, E.Y. Durand, M. Githiru, J.R. Huyghe and L.

Lens. 2011. Genetic signature of population fragmentation varies with mobility in

seven bird species of a fragmented Kenyan cloud forest. Molecular Ecology 20:

1829–1844.

Carpenter, F.G. 1891. An Age of Terrapin. In: Current Literature: A Magazine of Record

and Review, Volume VI (ed F. Somers), pp. 246–247. The Current Literature

Publishing Company, New York.

Castillo, J.A., C.W. Epps, A.R. Davis, and S.A. Cushman. 2014. Landscape effects on

gene flow for a climate-sensitive montane species, the American pika. Molecular

Ecology 23: 843–856.

Cecala, K.K., J.W. Gibbons, and M.E. Dorcas ME. 2008. Ecological effects of major

injuries in diamondback terrapins: implications for conservation and management.

Aquatic Conservation: Marine and Freshwater Ecosystems 19: 421–427.

Chassin-Noria, O., A. Abreu-Grobois, P.H. Dutton, and K. Oyama. 2004. Conservation

genetics of the east Pacific Green Turtle (Chelonia mydas) in Michoacan, Mexico.

Genetica 121: 195– 206.

Page 141: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

141

Chiucchi, J.E. and H.L. Gibbs 2010. Similarity of contemporary and historical gene flow

among highly fragmented populations of an endangered rattlesnake. Molecular

Ecology 19: 5345–5358.

Coker, R.E. 1906. The natural history and cultivation of the diamondback terrapin with

notes on other forms of turtles. Bulletin North Carolina Geological Survey 14:3–69.

Coker, R.E. 1920. The diamondback terrapin: past, present and future. Scientific Monthly

11: 171–186.

Coleman, A.T. 2011. Biology and conservation of the diamondback terrapin, Malaclemys

terrapin pileata, in Alabama. PhD dissertation, The University of Alabama at

Birmingham. 221 pages.

Converse, P.E., S.R. Kuchta, W.M. Roosenburg, P.F.P. Henry, G.M. Haramis, and T.L.

King. 2015. Spatiotemporal analysis of gene flow in Chesapeake Bay

Diamondback Terrapins (Malaclemys terrapin). Molecular Ecology 24: 5864–

5876.

Converse, P.E. and S.R. Kuchta. 2017. Molecular Ecology and Phylogeography of the

Diamond-backed Terrapin. In: The Ecology and Conservation of Diamond-backed

Terrapins, Malaclemys terrapin (eds W. Kennedy and W.M. Roosenburg). The

Johns Hopkins University Press, Baltimore MD.

Cooch, E. and G. White. 2008. Program MARK: a gentle introduction. Available in. pdf

format for free at http://www.phidot.org/software/mark/docs/book/

Cook, R. 1989. A natural history of the diamondback terrapin. Underwater Naturalist 18:

25–31.

Page 142: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

142

Cornuet, J.M. and G. Luikart. 1996. Description and power analysis of two tests for

detecting recent population bottlenecks from allele frequency data. Genetics 144:

2001–2014.

Cornuet, J. M., F. Santos, M. A. Beaumont, C. P. Robert, J. M. Marin, D. J. Balding, T.

Guillemaud, and A. Estoup. 2008. Inferring population history with DIY ABC: a user-

friendly approach to approximate Bayesian computation. Bioinformatics 24: 2713–

2719.

Cornuet, J. M., V. Ravigné, and A. Estoup. 2010. Inference on population history and

model checking using DNA sequence and microsatellite data with the software

DIYABC (v1. 0). BMC Bioinformatics 11:401.

Cornuet, J.M., P. Pudlo, J. Veyssier, A. Dehne-Garcia, M. Gautier, R. Leblois, J.-M.

Marin, and A. Estoup. 2014. DIYABC v2.0: a software to make approximate

Bayesian computation inferences about population history using single nucleotide

polymorphism, DNA sequence and microsatellite data. Bioinformatics 30:1187–

1189.

Crispo, E., J.-S. Moore, J.A. Lee-Yaw, S.M. Gray, and B.C. Haller. 2011. Broken

barriers: human-induced changes to gene flow and introgression in animals.

Bioassays 33: 508–518.

Crooks, K.R. and M.E. Soulé. 1999. Mesopredator release and avifaunal extinctions in a

fragmented system. Nature 400: 563–566.

Crow, J.F. 1987. Population genetics history: a personal view. Annual Review of

Genetics 21: 1–22.

Page 143: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

143

Crow, J.F. 2010. Wright and Fisher on Inbreeding and Random Drift. Genetics 184: 609–

611.

Csilléry, K., M.G.B. Blum, O.E. Gaggiotti and O. François. 2010. Approximate Bayesian

Computation (ABC) in practice. Trends in Ecology and Evolution 25: 410–418.

de Queiroz, K. 2007. Species concepts and species delimitation. Systematic Biology 56:

879–866.

Dethmers, K.M., D. Broderick, C. Moritz, et al. 2006. The genetic structure of

Australasian Green Turtles (Chelonia mydas): exploring the geographical scale of

genetic exchange. Molecular Ecology 15: 3931–3946.

Do C., R.S. Waples, D. Peel, G.M. Macbeth, B.J. Tillett, and J.R. Ovenden. 2014.

NeEstimator V2: re-implementation of software for the estimation of contemporary

effective population size (Ne) from genetic data. Molecular Ecology Resources 14:

209–214.

Dodd, C.K.J. and R.A. Seigel 2002. Translocations of Amphibians: Proven Management

Method or Experimental Technique? Conservation Biology 16: 552–554.

Doody, J.S., B. Green, D. Rhind, et al. 2009. Population-level declines in Australian

predators caused by an invasive species. Animal Conservation 12: 46–53.

Dorcas, M.E., J.D. Willson, and J.W. Gibbons. 2007. Crab trapping causes population

decline and demographic changes in diamondback terrapins over two decades.

Biological Conservation 137: 334–340.

Dorcas, M.E. and J.D. Willson. 2012. Invasive Pythons in the United States: Ecology of

an Introduced Predator. University of Georgia Press. Athens, GA.

Page 144: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

144

Drabeck, D. H., and M. Chatfield. 2014. The status of Louisiana's Diamondback Terrapin

(Malaclemys terrapin) populations in the wake of the Deepwater Horizon oil spill:

insights from population genetic and contaminant analyses. Journal of Herpetology

48:125–136.

Dunn, O.J. 1961. Multiple comparisons among means. Journal of the American Statistical

Association 56: 52–64.

Earl, D. A., and B. M. vonHoldt. 2011. STRUCTURE HARVESTER: a website and

program for visualizing STRUCTURE output and implementing the Evanno

method. Conservation Genetics Resource 4: 359–361.

Eo, S. H. and J. A. DeWoody. 2010. Evolutionary rates of mitochondrial genomes

correspond to diversification rates and to contemporary species richness in birds

and reptiles. Proceedings of the Royal Society B: Biological Sciences 277: 3587–

3592.

Epps, C.W., P.J. Palsbøll, J.D. Wehausen, et al. 2005. Highways block gene flow and

cause a rapid decline in genetic diversity of desert bighorn sheep. Ecology Letters

8: 1029–1038.

Epps, C.W., S.K. Wasser, J.L. Keim, B.M. Mutayoba and J.S. Brashares. 2013.

Quantifying past and present connectivity illuminates a rapidly changing landscape

for the African elephant. Molecular Ecology 22:1574–1588.

Epps, C.W. and N. Keyghobad 2015. Landscape genetics in a changing world:

disentangling historical and contemporary influences and inferring change.

Molecular Ecology 24: 6021–6040.

Page 145: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

145

Ernst, C. H. and R. W. Barbour, eds. 1989. Turtles of the world. Smithsonian Institution

Scholarly Press.

Estoup, A., P. Jarne, and J.M. Cornuet. 2002. Homoplasy and mutation model at

microsatellite loci and their consequences for population genetics analysis.

Molecular Ecology 11: 1591–1604.

Evanno, G., S. Regnaut, and J. Goudet. 2005. Detecting the number of clusters of

individuals using the software structure: a simulation study. Molecular Ecology 14:

2611–2620.

Excoffier, L., P.E. Smouse and J. M. Quattro. 1992. Analysis of molecular variance

inferred from metric distance among DNA haplotypes: application to human

mitochondrial DNA restriction data. Genetics 131: 479–491.

Excoffier, L. and H.E.L. Lischer. 2010. Arlequin suite ver 3.5: a new series of programs

to perform population genetics analyses under Linux and Windows. Molecular

Ecology Resources 10: 564–567.

Fabiani, A., F. Galimberti, S. Sanvito, and A.R. Hoelzel. 2004. Extreme polygyny

among southern elephant seals on Sea Lion Island, Falkland Islands. Behavioral

Ecology 15: 961–969.

Faubet, P., R.S. Waples, and O.E. Gaggiotti. 2007. Evaluating the performance of a

multilocus Bayesian method for the estimation of migration rates. Molecular

Ecology 16: 1149–1166.

Page 146: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

146

Feinberg, J.A. and R.L. Burke. 2003. Nesting ecology and predation of Diamondback

Terrapins, Malaclemys terrapin, at Gateway National Recreation Area, New York.

Journal of Herpetology 37: 517–526.

Fisher, R.A. 1922. On the dominance ratio. Proceedings of the Royal Society of

Edinburgh 42: 321–341.

Frankham, D., J. Ballou and D. Briscoe. 2010) Conservation Genetics. The Cambridge

University Press, Cambridge, UK.

Freedberg, S., M.A. Ewert, B.J. Ridenhour, M. Neiman and C.E. Nelson. 2005. Nesting

fidelity and molecular evidence for natal homing in the freshwater turtle,

Graptemys kohnii. Proceedings of the Royal Society B: Biological Sciences 272:

1345–1350.

Funk, W.C., E.D. Forsman, M. Johnson, T.D. Mullins, and S.M. Haig. 2010. Evidence

for recent population bottlenecks in northern spotted owls (Strix occidentals

caurina). Conservation Genetics 11: 1013–1021.

Galbraith, D.A., B.N. White, R.J. Brooks and P.T. Boag. 1993. Multiple paternity in

clutches of snapping turtles (Chelydra serpentina) detected using DNA

fingerprints. Canadian Journal of Zoology 71: 318–324.

Garber, S.W. 1988. Diamondback terrapin exploitation. Plastron Papers 17: 18–22.

Garber, S.W. 1990. The ups and downs of the diamondback terrapin. Conservationist 44:

44–47.

Page 147: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

147

Gerlach, G., A. Jueterbock, P. Kraemer, J. Deppermann, and P. Harmand. 2010.

Calculations of population differentiation based on GST and D: forget GST but not

all statistics! Molecular Ecology 19: 3845–3852.

Gibbons, J.W., J.E. Lovich, A.D. Tucker, N.N. FitzSimmons and J.L. Greene. 2001.

Demographic and ecological factors affecting conservation and management of the

diamondback terrapin (Malaclemys terrapin) in South Carolina. Chelonian

Conservation and Biology 4: 66–74.

Glenos, S.M. 2013. A comparative assessment of genetic variation of diamondback

terrapin (Malaclemys terrapin) in Galveston Bay, Texas in relation to other

northern Gulf Coast populations. MS thesis, The University of Houston, Clear Lake

TX. 71 pages.

Goedbloed, D.J., P. van Hooft, H.-J. Megens, K. Langenbeck, W. Lutz, R.P.M.A.

Crooijmans, S.E. van Wieren, R.C. Ydenberg, and H.H.T. Prins. 2013.

Reintroductions and genetic introgression from domestic pigs have shaped the

genetic population structure of Northwest European wild boar. BMC Genetics 14:

1–10.

Grant, K. M., E. J. Rohling, M. Bar-Matthews, A. Ayalon, M. Medina-Elizalde, C. B.

Ramsey, C. Satow, and A. P. Roberts. 2012. Rapid coupling between ice volume

and polar temperature over the past 150,000 years. Nature 491: 744–747.

Grosse, A.M., J.D. Dijk, K.L. Holcomb, and J.C. Maerz. 2009. Diamondback terrapin

mortality in crab pots in a Georgia tidal marsh. Chelonian Conservation and

Biology 8: 98–100.

Page 148: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

148

Gu, W., R. Heikkilä, and I. Hanski. 2002. Estimating the consequences of habitat

fragmentation on extinction risk in dynamic landscapes. Landscape Ecology 17:

699–710.

Hansen, B.D., K.P Harley, D.B. Lindenmayer, and A.C. Taylor. 2009. Population genetic

analysis reveals a long-term decline of a threatened endemic Australian marsupial.

Molecular Ecology 18: 3346–3362.

Hanski, I.H. and M. Gilpin. 1997. Metapopulation Biology: Ecology, Genetics, and

Evolution. Academic Press, San Diego, CA.

Haramis, G.M., P.F. Henry, and D.D. Day. 2011. Using scrape fishing to document

terrapins in hibernacula in Chesapeake Bay. Herpetological Review 42: 170–177.

Harden, L.A., N.A. Diluzio, J.W. Gibbons and M.E. Dorcas. 2007. Spatial and thermal

ecology of the diamondback terrapin (Malaclemys terrapin) in a South Carolina salt

marsh. Journal of the North Carolina Academy of Sciences 123:154–162.

Hart, K.M. 2005. Population biology of diamondback terrapins: defining and reducing

threats across their geographic range. PhD dissertation, Duke University, Durham

NC. 259 pages.

Hart, K.M., M.E. Hunter and T.L. King. 2014. Regional differentiation among

populations of the Diamondback terrapin (Malaclemys terrapin). Conservation

Genetics 15:593–603.

Hauswaldt, J.S. 2004. Population genetics and mating pattern of the diamondback

terrapin (Malaclemys terrapin). PhD dissertation, University of South Carolina,

Columbia SC. 216 pages.

Page 149: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

149

Hauswaldt, J.S, and T.C. Glenn. 2003. Microsatellite DNA loci from the Diamondback

terrapin (Malaclemys terrapin). Molecular Ecology Notes 3:174–176.

Hauswaldt, J. S., and T. C. Glenn. 2005. Population genetics of the diamondback terrapin

(Malaclemys terrapin). Molecular Ecology 14:723–732.

Hay, W.P. 1917. Artificial propagation of the diamond-back terrapin. Bulletin of the

United States Bureau of Fisheries 24: 1–20.

Hedrick, P.W. 1996. Bottlenecks(s) or metapopulation in cheetas. Conservation Biology

10: 897–899.

Hey, J. and R. Nielsen. 2004. Multilocus methods for estimating population sizes,

migration rates and divergence time, with applications to the divergence of

Drosophila pseudoobscura and D. persimilis. Genetics 167: 747–760.

Hildebrand, S.F. 1929. Review of experiments on artificial culture of diamondback

terrapin. Bulletin U.S. Bureau of Fisheries 45: 25–70.

Hildebrand, S.F. 1933. Hybridizing diamond-back terrapins. Journal of Heredity 113:

231–238.

Hildebrand, S.F. and C. Hatsel. 1926. Diamond-back terrapin culture at Beaufort, N.C.

U.S. Bureau of Fisheries Economic Circular 60: 1–20.

Hillis, D.M. and J.J. Bull. 1993. An empirical test of bootstrapping as a method for

assessing confidence in phylogenetic analysis. Systematic Biology 42: 182–192.

Hillis, D. M., B. K. Mable, and C. Moritz, eds. 1996. Application of Molecular

Systematics: The State of the Field and a Look to the Future. Molecular

Systematics. Sinauer Associates, Sunderland, MA.

Page 150: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

150

Hindar, K., N. Ryman, and F. Utter. 1991. Genetic effects of cultured fish on natural fish

populations. Canadian Journal of Fisheries and Aquatic Sciences 48: 945–957.

Ho, S.Y.M. 2007. Calibrating molecular estimates of substitution rates and divergence

times in birds. Journal of Avian Biology 38: 409–414.

Hudson, R. R., and J. A. Coyne. 2002. Mathematical consequences of the genealogical

species concept. Evolution 56: 1557–1565.

Husemann, M., L. Cousseau, T. Callens, E. Matthysen, C. Vangestel, C. Hallmann, and

L. Lens. 2015. Post-fragmentation population structure in a cooperative breeding

Afrotropical cloud forest bird: emergence of a source-sink population network.

Molecular Ecology 24: 1172–1187.

Ireland, J.S., A.C. Broderick, F. Glen, B.J. Godley, G.C. Hays, P.L.M. Lee and D.O.F.

Skibinski. 2003. Multiple paternity assessed using microsatellite markers, in green

turtles Chelonia mydas (Linnaeus, 1758) of Ascension Island, South Atlantic.

Journal of Experimental Marine Biology and Ecology 291:149–160.

Jakobsson, M., and N. A. Rosenberg. 2007. CLUMPP: a cluster matching and

permutation program for dealing with label switching and multimodality in analysis

of population structure. Bioinformatics 23: 1801–1806.

Jolly, G.M. 1965. Explicit estimates from capture-recapture data with both death and

immigration – Stochastic model. Biometrika 52: 225–247.

Jombart, T. 2008. adegenet: a R package for the multivariate analysis of genetic markers.

Bioinformatics 24: 1403–1405.

Page 151: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

151

Jost, L. 2008. GST and its relatives do not measure differentiation. Molecular Ecology

18: 4015–4026.

Keeney, D. B., M. R. Heupel, R. E. Hueter, and E. J. Heist. 2005. Microsatellite and

mitochondrial DNA analyses of the genetic structure of blacktip shark

(Carcharhinus limbatus) nurseries in the northwestern Atlantic, Gulf of Mexico,

and Caribbean Sea. Molecular Ecology 14: 1911–1923.

Kennedy, V. (2017) History of Commercial Fisheries and Artificial Propagation of

Dimondback Terrapins. In: The Ecology and Conservation of Diamond-backed

Terrapins, Malaclemys terrapin (eds V. Kennedy and W.M. Roosenburg). The

Johns Hopkins University Press, Baltimore, MD.

Kimura, M. and T. Ohta. 1978. Stepwise mutation model and distribution of allelic

frequencies in a finite population. Proceedings of the National Academy of

Sciences 75: 2868–2872.

King, T.L. and S.E. Julian. 2004. Conservation of microsatellite DNA flanking sequence

across 13 Emydid genera assayed with novel bog turtle (Glyptemys muhlenbergii)

loci. Conservation Genetics 5:719–725.

King, T.L., J.F. Switzer, C.L. Morrison, M.S. Eackles, C.C. Young, B.A. Lubinski, and

P. Cryan. 2006. Comprehensive genetic analyses reveal evolutionarily distinction

of a mouse (Zapus hudsonius preblei) proposed for delisting from the U.S.

Endangered Species Act. Molecular Ecology 15: 4331– 4359.

Kirkpatrick, M. and N.H. Batron. 1997. Evolution of a species’ range. American

Naturalist 150: 1–23.

Page 152: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

152

Krysko, K.L., K.M. Enge, and F.W. King. 2004. The veiled chameleon, Chamaeleo

calyptratus: A new exotic lizard species in Florida. Florida Scientist, 67, 249–253.

Kuchta, S.R. and A.M. Tan. 2005. Isolation by distance and post-glacial range expansion

in the rough-skinned newt, Taricha granulosa. Molecular Ecology 14: 225–244.

Kuchta, S.R., and A.M. Tan. 2006. Lineage diversification on an evolving landscape:

phylogeography of the California newt, Taricha torosa (Caudata: Salamandridae).

Biological Journal of the Linnaean Society 89: 213–239.

Kuchta, S.R. and A.M. Tan. 2006. Limited genetic variation across the range of the red-

bellied newt, Taricha rivularis. Journal of Herpetology 40: 561–565.

Kuchta, S.R., D.S. Parks, R.L. Mueller, and D.B. Wake. 2009. Closing the ring: historical

biogeography of the salamander ring species Ensatina eschscholtzii. Journal of

Biogeography 36: 982–995

Kuchta, S.R., and D.B. Wake. 2016. Wherefore and whither the ring species? Copeia,

104: 189–201.

Kuchta, S.R., A.D. Brown, P.E. Converse, and R. Highton. 2016. Multilocus

phylogeography of the Cumberland Plateau Salamander, Plethodon kentucki:

species delimitation methods at the phylogeographic scale. PLoS ONE 11(3):

e0150022.

Laikre L., M.K. Schwartz, R.S. Waples, N. Ryman, and The GeM Working Group. 2010.

Compromising genetic diversity in the wild: unmonitored large-scale release of

plants and animals. Trends in Ecology and Evolution 25: 520–529.

Page 153: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

153

Lamb, T. and J.C. Avise. 1992. Molecular and population genetic aspects of

mitochondrial DNA variability in the diamondback terrapin, Malaclemys terrapin.

Journal of Heredity 83: 262–269.

Lande, R. 1995. Mutation and conservation. Conservation Biology 9: 782–791.

Landguth E.L., S.A. Cushman, M.K. Schwartz, M. Murphy, K.S. McKelvey, and G.

Luikart. 2010. Quantifying the lag time to detect barriers in landscape genetics.

Molecular Ecology 19: 4179–4191.

Lazell, J.D. 1976. This Broken Archipelago: Cape Cod and the Islands, Amphibians and

Reptiles. Quadrangle/New York Times Book Co., New York.

Lee T. and D. Ó Foighil. 2003. Phylogenetic structure of the Sphaeriinae, a global clade

of freshwater bivalve molluscs, inferred from nuclear (ITS-1) and mitochondrial

(16S) ribosomal gene sequences. Zoological Journal of the Linnean Society 137:

245–260.

Lester, L.A. 2007. Tracking terrapins through genetic analysis: Multilocus assignment

tests shed light on origin of turtles sold in markets. MEM thesis, Duke University,

Durham NC. 38 pages.

Linnen, C. R., and B. D. Farrell. 2007. Mitonuclear discordance is caused by rampant

mitochondrial introgression in Neodiprion (Hymenoptera: Diprionidae) sawflies.

Evolution 61: 1417–1438.

Lopes, J.S., D. Balding, and M.A. Beaumont. 2009. PopABC: a program to infer

historical demographics. Bioinformatics 25: 2747–2749.

Page 154: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

154

Lovich, J.E. and J.W. Gibbons. 1990. Age at maturity influences adult sex ratio in the

turtle Malaclemys terrapin Oikos 59: 126–134.

Luikart, G. and J.M. Cornuet. 1998. Empirical evaluation of a test for identifying recently

bottlenecked populations from allele frequency data. Conservation Biology 12:

228–237.

Maddocks, C.E., E.T. Nolan, S.W. Feist, et al. 2015. Puffy skin disease (PSD) in rainbow

trout, Oncorhynchus mykiss (Walbaum): a case definition. Journal of Fish Diseases

38: 653–664.

Marsack, K.M. and B.J. Swanson. 2009. A genetic analysis of the impact of generation

time and road-based habitat fragmentation on Eastern Box Turtles (Terrepene c.

carolina). Copeia 2009: 647–652.

McCafferty, S.S., A. Shorette, J. Simundza and B. Brennessel. 2013. Paucity of genetic

variation at an MHC Class I gene in Massachusetts populations of the diamond-

backed terrapin (Malaclemys terrapin): a cause for concern? Journal of

Herpetology 47: 222–226.

McRae, B.H. and P. Beier. 2007. Circuit theory predicts gene flow in plant and animal

populations. Proceedings of the National Academy of Sciences 104: 19885–19890.

Meirmans, P.G. 2014. Nonconvergence in Bayesian estimation of migration rates.

Molecular Ecology Resources 14: 726–733.

Mitro, M.G. 2003. Demography and viability analyses of a diamondback terrapin

population. Canadian Journal of Zoology 81: 716–726.

Page 155: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

155

Molfetti É., S.T. Vilaça, J.Y. Georges, et al. 2013. Recent demographic history and

present fine-scale structure in the Northwest Atlantic Leatherback (Dermoclemys

coriacea) Turtle population. PLOS ONE, 8: e58061.

Moran, P.A.P. 1958. Random processes in genetics. Mathematical Proceedings of the

Cambridge Philosophical Society 54: 60–71.

Moritz, C. 1999. Conservation Units and Translocations: Strategies for Conserving

Evolutionary Processes. Hereditas 130: 217–228.

Naylor, R., K. Hindar, I.A. Fleming, et al. 2005. Fugitive Salmon: Assessing the Risks of

Escaped Fish from Net-Pen Aquaculture. BioScience 55: 427–437.

Newmark, W.D. 2008. Isolation of African protected areas. Frontier in Ecology and the

Environment 6: 321–328.

Nielsen, R. and J. Wakeley. 2001. Distinguishing migration from isolation: a Markov

chain Monte Carlo approach. Genetics 158: 885–896.

Nomura, T. 2008. Estimation of effective number of breeders from molecular coancestry

of single cohort sample. Evolutionary Applications 1: 462–474.

Nordborg, M. and S.M. Krone. 2002. Separation of time scales and convergence to the

coalescent in structured populations. In: Modern Developments in Theoretical

Population Genetics (eds M. SlatkinM and M. Veuille), pp.130–164. Oxford

University Press, Oxford.

Nussear, K.E., C.R. Tracy, P.A. Medica, et al. 2012. Translocation as a Conservation

Tool for Agassiz's Desert Tortoises: Survivorship, Reproduction, and Movements.

The Journal of Wildlife Management 76: 1341–1353.

Page 156: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

156

Oksanen, J., F.G. Blanchet, R. Kindt, et al. 2013 vegan: Community Ecology Package. R

package version 2010 http://CRAN.R-project.org/package=vegan

Oosterhout, C.V., W.F. Hutchinson, D.P.M. Wills, and P. Shipley. 2004. MICRO-

CHECKER: software for identifying and correcting genotyping errors in

microsatellite data. Molecular Ecology Notes 4: 535–538.

Parker, I.D., D.E. Watts, R.R. Lopez, et al. 2008. Evaluation of the Efficacy of Florida

Key Deer Translocations. The Journal of Wildlife Management 72: 1069–1075.

Pavlacky, D.C., A.W. Goldizen, P.J. Prentis et al. 2009. A landscape genetics approach

for quantifying the relative influence of historic and contemporary habitat

heterogeneity on the genetic connectivity of a rainforest bird. Molecular Ecology

18: 2945–2960.

Peery, M.Z., R. Kirby, B.N. Reid, et al. 2012. Reliability of genetic bottleneck tests for

detecting recent population declines. Molecular Ecology 21: 3403–3418.

Pérez, I., J.D. Anadón, M. Díaz, et al. 2012. What is wrong with current translocations?

A review and a decision-making proposal. Frontiers in Ecology and the

Environment 10: 494–501.

Petre, C., W. Selman, B. Kreiser, S. H. Pearson, and J. J. Wiebe. 2015. Population

genetics of the diamondback terrapin, Malaclemys terrapin, in Louisiana.

Conservation Genetics16: 1243–1252.

Piry, S., G. Luikart, and J.M. Cornuet. 1999. BOTTLENECK: A computer program for

detecting recent reductions in the effective population size using allele frequency

data. Journal of Heredity 90: 502--503.

Page 157: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

157

Pongsiri, M.J., J. Roman, V.O. Ezenwa, et al. 2009. Biodiversity Loss Affects Global

Disease Ecology. BioScience 59: 945–954.

Pritchard, J. K., M. Stephens, and P. Donnelly. 2000. Inference of population structure

using multilocus genotype data. Genetics 155: 945–959.

Pritchard, J.K., X. Wen and D. Falush. 2003. Documentation for structure software:

Version 2.3. Retrieved from http://pritch.bsd.uchicago.edu/structure.html

Puckett, E.E., T.V. Kristensen, C.M. Wilton, S.B. Lyda, K.V. Noyce, P.M. Holahan,

D.M. Leslie, J. Beringer, J.L. Belant, D. White, and L.S. Eggert. 2014. Influence of

drift and admixture on population structure of American black bears (Ursus

americanus) in the Central Interior Highlands, USA, 50 years after translocation.

Molecular Ecology 23: 2414–2427.

R Core Team. 2014. R: A language and environment for statistical computing. R

Foundation for Statistical Computing, Vienna, Austria. URL: http://www.R-

project.org/

Radzio, T.A. and W.M. Roosenburg. 2005. Diamondback terrapin mortality in the

American eel pot fishery and evaluation of a bycatch reduction device. Estuaries

and Coasts 28: 620–626.

Radzio, T.A., J.A. Smolinsky, and W.M. Roosenburg. 2013. Low use of required terrapin

bycatch reduction devices in a recreational crab pot fishery. Herpetological

Conservation and Biology 8: 222–227.

Rambaut, A. and A.J. Drummond. 2007. Tracer v1.4. Available from:

http://beast.bio.ed.ac.ul/Tracer

Page 158: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

158

Randazzo, A.F. and D.S. Jones. 1997. The Geology of Florida. University Press of

Florida, Gainesville, FL.

Rannala, B. 2011. BayesAss edition 3.0 user's manual.

Roosenburg, W. M. 1991. The diamondback terrapin: population dynamics, habitat

requirements, and opportunities for conservation. New perspectives in the

Chesapeake system: a research and management partnership. Proceedings of a

Conference. Chesapeake Research Consortium 137: 227–234.

Roosenburg, W. M. 1994. Nesting habitat requirements of the diamondback terrapin: a

geographic comparison. Wetland Journal 6: 8–11.

Roosenburg, W.M. 1996. Maternal condition and nest site choice: an alternative for the

maintenance of environmental sex determination? American Zoologist 36: 157–

168.

Roosenburg, W.M., W. Cresko, M. Modesitte, and M.B. Robbins. 1997. Diamondback

terrapin (Malaclemys terrapin) mortality in crab pots. Conservation Biology 11:

1166–1172.

Roosenburg, W.M., K.L. Haley and S. McGuire. 1999. Habitat selection and movements

of diamondback terrapins, Malaclemys terrapin, in a Maryland estuary. Chelonian

Conservation and Biology 3: 425–429.

Roosenburg, W.M. and J.P. Green. 2000. Impact of a bycatch reduction device on

diamondback terrapins and blue crab capture in crab pots. Ecological Applications

10: 882–889.

Page 159: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

159

Roosenburg, W.M., J. Cover, and P.P. van Dijk. 2008. Legislative Closure of the

Maryland Terrapin Fishery: Perspectives on a Historical Accomplishment. Turtle

and Tortoise Newsletter 12: 27–30.

Roosenburg, W.M., E. Clowes, and P.E. Converse. 2014. Terrapin Monitoring at Poplar

Island, 2013. Final Report Submitted to the Army Corps of Engineers, Baltimore

Office, Baltimore, MD pp. 22.

Rosenberg, N. A. 2003. distruct: a program for the graphical display of population

structure. Molecular Ecology Notes 4: 137–138.

Ryman, N., F. Utter, and L. Laikre. 1995. Protection of intraspecific biodiversity of

exploited fishes. Reviews in Fish Biology and Fisheries 5: 417–446.

Saunders, N. C., L. G. Kessler, and J. C. Avise. 1986. Genetic variation and geographic

differentiation in mitochondrial DNA of the horseshoe crab, Limulus polyphemus.

Genetics 112: 613–627.

Schelly, R., W. Salzburger, S. Koblmüller, N. Duftner, and C. Sturmbauer. 2006.

Phylogenetic relationships of the lamprologine cichlid genus Lepidiolamprologus

(Teleostei: Perciformes) based on mitochondrial and nuclear sequences, suggesting

introgressive hybridization. Molecular Phylogenetics and Evolution 38: 426–438.

Schwarz, C.J. and A.N. Arnason. 1996. A general methodology for the analysis of open-

model capture recapture experiments. Biometrics 52: 860–873.

Seber, G.A.F. 1965. A note on the multiple recapture census. Biometrika 52: 249–259.

Page 160: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

160

Seigel, R.A. 1980. Nesting habitats of diamondback terrapins (Malaclemys terrapin) on

the Atlantic Coast of Florida. Transactions of the Kansas Academy of Science 83:

239–246.

Shaffer, H. B., P. Minx, D. E. Warren, and A. M. Shedlock et al. 2013. The western

painted turtle genome, a model for the evolution of extreme physiological

adaptations in a slowly evolving lineage. Genome Biology 14: R28.

Sheridan, C.M. 2010. Mating system and dispersal patterns in the diamondback terrapin

(Malaclemys terrapin). PhD dissertation, Drexel University, Philadelphia PA. 204

pages.

Sheridan, C. M., J. R. Spotila, W. F. Bien, and H. W. Avery. 2010. Sex-biased dispersal

and natal philopatry in the diamondback terrapin, Malaclemys terrapin. Molecular

Ecology 19: 5497–5510.

Short, J., S.D. Bradshaw, J, Giles, and R. Prince. 1992. Reintroduction of macropods

(Marsupialia: Macropodoidea) in Australia—a review. Biological Conservation 62:

189–204.

Siegel, R.A. 1980. Nesting habits of diamondback terrapins (Malaclemys terrapin) on the

Atlantic coast of Florida. Transactions of the Kansas Academy of Sciences 83:

239–246.

Sjödin P., I. Kaj, S. Krone, M. Lascoux, and M. Nordberg. 2005. On the meaning and

existence of an effective population size. Genetics 169: 1061–1070.

Slater, G. J., L. J. Harmon, and M. E. Alfaro. 2012. Integrating fossils with molecular

phylogenies improves inference of trait evolution. Evolution 66: 3931–3944.

Page 161: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

161

Slatkin, M. 1987. Gene flow and the geographic structure of natural populations. Science

236: 787–792.

Soltis, D.E., A.B. Morris, J.S. McLachlan, P.S. Manos and P.S. Soltis. 2006.

Comparative phylogeography of unglaciated eastern North America. Molecular

Ecology 15: 4261–4293.

Soto, D., F. Jara, and C. Moreno. 2001. Escaped Salmon in the Inner Seas, Southern

Chile: Facing Ecological and Social Conflicts. Ecological Applications 11: 1750–

1762.

Souty-Grosset, C. and F. Grandjean. 2008. Biodiversity Conservation And Habitat

Management (ed. F. Gherardi, M. Gualtieri, and C. Corti). Encyclopedia of Life

Support Systems (EOLSS).

Spiegelhalter, D.J., N.G. Best, and B.P. Carlin. 2002. Bayesian measures of model

complexity and fit. Journal of the Royal Statistical Society. Series B

(Methodological), 64: 583–639.

Spivey, P.B. 1998. Home range, habitat selection, and diet of the diamondback terrapin

(Malaclemys terrapin) in a North Carolina estuary. MS thesis. University of

Georgia, Athens GA. 80 pages.

Storfer, A. 1999. Gene flow and endangered species translocations: a topic revisited.

Biological Conservation 87: 173–180.

Szerlag, S. and S.P. McRobert. 2006. Road occurrence and mortality of the northern

diamondback terrapin. Applied Herpetology 3: 27–37.

Page 162: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

162

Tallmon, D.A., A, Koyuk , G.H. Luikart, and M.A. Beaumon. 2008. ONeSAMP: a

program to estimate effective population size using approximate Bayesian

computation. Molecular Ecology Resources 8: 299–301.

Templeton, A.R., H. Hemmer, G. Mace, U.S. Seal, W.M. Shields, and D.S. Woodruff.

1986. Local adaption, coadaptation, and population boundaries. Zoo Biology 5:

115–125.

Templeton, A.R., R.J. Robertson, J. Brisson, and J. Strasburg. 2001. Disrupting

evolutionary processes: The effect of habitat fragmentation on collared lizards in

the Missouri Ozarks. Proceedings of the National Academy of Sciences 98: 5426–

5432.

Theissinger, K., N.N. FitzSimmons, C.J. Limpus, C.J. Parmenter, and A.D. Phillott.

2009. Mating system, multiple paternity, and effective population size in the

endemic Flatback Turtle (Natator depressus) in Australia. Conservation Genetics

10: 329–346.

Thorne, J.L., H. Kishino, and I.S. Painter. 1998. Estimating the rate of evolution of the

rate of molecular evolution. Molecular Biology and Evolution 15: 1647–1657.

Tran, T.D., J. Hofrichter, and J. Jost. 2013. An introduction to the mathematical structure

of the Wright–Fisher model of population genetics. Theory in Biosciences, 132:

73–82.

Valenzuela, N. 2000. Multiple paternity in side-neck turtles Podocnemis expansa:

evidence from microsatellite DNA data. Molecular Ecology 9: 99–105.

Page 163: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

163

Vandergast, A.G., A.J. Bohonak, D.B. Weissman, et al. 2007. Understanding the genetic

effects of recent habitat fragmentation in the context of evolutionary history:

phylogeography and landscape genetics of a southern California endemic Jerusalem

cricket (Orthoptera: Stenopelmatidae: Stenpelmatus). Molecular Ecology 16: 977–

992.

Wakeley, J., ed.. 2009. Coalescent Theory: An Introduction. Roberts and Company

Publishers, Greenwood Village CO.

Waples, R.S. and C. Do. 2008. LDNE: a program for estimating effective population size

from data on linkage disequilibrium. Molecular Ecology Resources 8: 753–756.

Webb, S.D. 1990. Historical biogeography. In: Ecosystems of Florida (eds R.I. Myers

and J.J. Ewel), pp. 70–100. University of Florida Press. Gainesville, FL.

Weeks, A.R., C.M. Sgro, A.G. Young, R. Frankham, et al. 2011. Assessing the benefits

and risks of translocations in changing environments: a genetic perspective.

Evolutionary Applications 4: 709–725.

Westemeier, R., J. Brawn, S. Simpson, et al. 1998. Tracking the long-term decline and

recovery of an isolated population. Science 282: 1695–1698.

White, G.C., and K.P. Burnham. 1999. Program MARK: Survival estimation from

populations of marked animals. Bird Study S46: S120–S138.

Whittaker, J.C., R.M. Harbord, N. Boxall, et al. 2003. Likelihood-based estimation of

microsatellite mutation rates. Genetics 164: 781–787.

Wiktor, D.M., M. Hill, and R.M. Chambers. 2001. Genetic diversity of diamondback

terrapins (Malaclemys terrapin) from Piermont Marsh, Hudson River, NY. In: (eds

Page 164: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

164

J.R. Waldman and W.C. Nieder) Final Reports of the Tibor T. Polgar Fellowship

Program, 2000. Hudson River Foundation, New York.

Wilcox, B.A. and D.D. Murphy. 1985. Conservation strategy: the effects of

fragmentation on extinction. American Naturalist 125: 879–887.

Wilson, G.A. and B. Rannala. 2003. Bayesian inference of recent migration rates using

multilocus genotypes. Genetics 163: 1177–1191.

Wise, J., M.G. Harasewych, and R.T. Dillon. 2004. Population divergence in the sinistral

whelks of North America, with special reference to the east Florida ecotone.

Marine Biology 145: 167–1179.

Wood, R.C. and R. Herlands. 1997. Turtles and tires: the impact of road kills on northern

diamondback terrapin, Malaclemys terrapin terrapin, populations on the Cape May

peninsula, southern New Jersey. Proceedings: Conservation, Restoration, and

Management of Tortoises and Turtles–An International Conference. New York

Turtle and Tortoise Society, New York, USA, 46–53.

Worm, B., E.B. Barbier, N. Beaumont, et al. 2006. Impacts of Biodiversity Loss on

Ocean Ecosystem Services. Science 314: 787–790.

Wright, S. 1931 Evolution in Mendelian populations. Genetics 16: 97–159.

Wright, S. 1938. Size of population and breeding structure in relation to evolution.

Science 87: 430–431.

Yang, Z., ed. 2014. Molecular Evolution: a statistical approach. Oxford University Press,

Oxford, United Kingdom.

Page 165: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

165

Young, A., C. Torres, J. Mack, and C. Cunningham. 2002. Morphological and genetic

evidence for vicariance and refugium in Atlantic and Gulf of Mexico populations of

the hermit crab Pagurus longicarpus. Marine Biology 140: 1059–1066.

Zhang, J.-B., R.-Q. Li, X.-G. Xiang, S. R. Manchester, L. Lin, W. Wang, J. Wen, and Z.-

D. Chen. 2013. Integrated fossil and molecular data reveal the biogeographic

diversification of the eastern Asian-eastern North American disjunct hickory genus

(Carya Nutt.). PLoS ONE 8: e70449.

Zhdanova, O., and A.I. Pudovkin. 2008. Nb_HetEx: a program to estimate the effective

number of breeders. Journal of Heredity 99: 694–695.

Page 166: The Population Ecology, Molecular Ecology, and ... 2016.pdf · The Population Ecology, Molecular Ecology, and Phylogeography of the Diamondback Terrapin (Malaclemys terrapin) Director

!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!!

!!

Thesis and Dissertation Services