Historical and Contemporary Genetic Perspectives on New World Monk Seals ( )

Alicia Nicole Mihnovets

Submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in the Graduate School of Arts and Sciences

COLUMBIA UNIVERSITY

2017

© 2017

Alicia Nicole Mihnovets

All Rights Reserved

ABSTRACT

Historical and Contemporary Genetic Perspectives on New World Monk Seals (Genus Neomonachus)

Alicia Nicole Mihnovets

Through common descent, closely-related taxa share many life history traits, some of which can influence -proneness. Thus, examining historical and contemporary genetic patterns is valuable in accounting for evolutionary and ecological processes that may be critical to the successful conservation of threatened .

Unsustainable harvesting of monk seals (tribe Monachini) until the late nineteenth century caused the recent extinction of monk seals (Neomonachus tropicalis) and critically low population sizes for Hawaiian and Mediterranean monk seals

(Neomonachus schauinslandi and Monachus monachus, respectively). Having lost one branch of its evolutionary lineage, and with a second branch threatened by extinction, the genus Neomonachus can serve as a valuable case for examining evolutionary and ecological linkages that are sensitive to non-random anthropogenic selection pressure.

An important foundation for such pursuits is the understanding of evolutionary sequences of speciation and diversification that gave rise to common traits shared by extinct and vulnerable species. Further consideration of the phylogenetic non- randomness of species vulnerability requires examination of genetic variation at the population level to infer the presence of fundamental processes (e.g., migration and reproduction) that directly influence population viability.

This dissertation includes three individual studies that make use of molecular systematic and population genetic techniques to address these topics. First, a complete mitochondrial genome sequence of the extinct Caribbean (N. tropicalis) was assembled and used to resolve long-standing phylogenetic questions regarding the sequence of divergence among monk seal species and sister taxa. Second, novel microsatellite marker assays were developed and used to characterize the extent of population-level variation across 24 polymorphic microsatellite loci of 1192 endangered

Hawaiian monk seals (N. schauinslandi) that were sampled during a longitudinal study spanning three decades. Third, resulting genotypes from a subset of individuals (N= 785) were integrated with previously reported genotypes consisting of 18 other loci for the largest ever population-level assessment of N. schauinslandi genetic diversity and population differentiation throughout the Hawaiian archipelago. The new microsatellite data will be of particular value for future individual-level assessment of parentage and relatedness in N. schauinslandi, which will help managers better infer the reproductive mechanisms that factor into population persistence and recovery.

Results of this study expand understanding of the evolutionary and conservation genetic status of monk seals, as well as molecular genetic capacity, for future research regarding a unique and highly imperiled New World lineage.

TABLE OF CONTENTS

List of Figures ...... v

List of Tables ...... vii

List of Abbreviations ...... ix iv

Acknowledgements ...... xi vi

Dedication ...... xiv ix

CHAPTER 1. INTRODUCTION ...... 1

The Phylogenetic Relationships of Monk Seals ...... 3

The Population Status of Monk Seals ...... 3

The Conservation Genetics of Monk Seals ...... 7

Research Focus ...... 8

References ...... 10

CHAPTER 2. A RECONSTRUCTION OF THE MITOCHONDRIAL GENOME

OF THE CARRIBEAN MONK SEAL ...... 14

Abstract ...... 15

Introduction ...... 17

Materials and Methods ...... 20

Tissue sampling and aDNA extraction ...... 20

Mitochondrial genome sequencing and assembly ...... 20

Mitochondrial genome annotation and sequence analysis ...... 21

Phylogenetic analysis...... 21

Results and Discussion ...... 23

Mitochondrial genome sequencing and assembly ...... 23

Genome organization and nucleotide composition ...... 23

i

Protein coding regions ...... 24 tRNA and rRNA genes ...... 24

Non-coding regions ...... 25 Phylogenetic analysis...... 26

Conclusion ...... 27

Acknowledgements ...... 28

References ...... 29

Supplementary Information ...... 48

Ancient DNA extraction methods ...... 48

CHAPTER 3. A NOVEL MICROSATELLITE MULTIPLEX ASSAY FOR

THE ENDANGERED ...... 49

Abstract ...... 50

Introduction ...... 51

Methods and Results ...... 52

Acknowledgements ...... 54

Author’s Contributions ...... 54

References ...... 56

Supplementary Material ...... 61

CHAPTER 4. REVISITING THE GENETIC DIVERSITY AND CONNECTIVITY

OF THE ENDANGERED HAWAIIAN MONK SEAL ...... 63

Abstract ...... 64

Introduction ...... 67

ii

Methods and Materials ...... 69

Study system ...... 69

Sample collection, DNA extraction, and microsatellite genotyping ...... 70

Study design ...... 71

Data analysis ...... 72

Results ...... 75

DNA sample summary...... 75

Genetic diversity ...... 76

Genetic differentiation ...... 79

Inter-island movement ...... 83

Discussion ...... 84

Genetic diversity ...... 84

Genetic differentiation ...... 87

Inter-island movement ...... 91

Conclusion ...... 93

Acknowledgements ...... 94

References ...... 95

iii

Supplementary Material ...... 147

Data management and quality control for consensus genotyping ...... 147

CHAPTER 5. CONCLUSION ...... 150

Contributions and Implications ...... 152

Next Steps ...... 154

Dispersal ...... 155

Kinship ...... 155

Sexual selection ...... 156

Emerging opportunities ...... 156

References ...... 158

iv

LIST OF FIGURES

Figure 2-1. N. tropicalis illustration ...... 33

Figure 2-2. Geographical distribution of tribes Monachini, Miroungini, and ...... 34

Figure 2-3. N. tropicalis pelt ...... 35

Figure 2-4. Mitochondrial genome sequence map of N. tropicalis ...... 36

Figure 2-5. Inferred secondary cloverleaf structures for the 22 tRNA genes of N. tropicalis ...... 37

Figure 2-6. Cladogram ...... 39

Figure 3-1. Comparison of the cumulative probability of identity for unrelated individuals and siblings...... 58

Figure 4-1. Map of the Northwestern Hawaiian Islands (NWHI) and main Hawaiian Islands (MHI) ...... 101

Figure 4-2. Comparison of the probability of distinguishing between unrelated individuals, and between siblings in a given population ...... 102

Figure 4-3. Calculations of mean expected heterozygosity plotted against mean observed heterozygosity for data partitions ...... 103

Figure 4-4. P-values of HWE exact tests for homozygosity and heterozygosity excess for data partitions ...... 103

Figure 4-5. Inbreeding coefficient (Fis) plotted relative to sample sizes of data partitions examined ...... 104

Figure 4-6. Observed and expected heterozygosity and inbreeding coefficient plotted for each time increment and for combined time increments (all years) ...... 104

Figure 4-7. Unbiased global estimates for Fst and four analogues ...... 105

Figure 4-8. Results of genetic individual assignment tests for five data partitions ...... 106

Figure 4-9. Principal component analysis (PCoA) of the “All years” data partition ...... 107

Figure 4-10. PCoA plots for data partitions of samples collected before 1991 and between 1991 and 1996 ...... 108

v

Figure 4-11. PCoA plots for data partitions of samples collected between 1997 and 2002 and between 2003 and 2007 ...... 109

Figure 4-12. Isolation-by-distance in Hawaiian monk seals ...... 110

Figure 4-13. Spatial autocorrelation analysis in Hawaiian monk seals (HMS) ...... 113

Figure 4-14. Relative migration networks generated based on estimates of Nm for all data partitions ...... 118

Figure 4-15. Classification by Schmelzer et al. of Hawaiian monk seal breeding areas into biogeographical regions according to differential environmental variables ...... 119

Figure 4S-1. Configuration of relational database created in Microsoft Access to manage genetic and population demography data ...... 149

vi

LIST OF TABLES

Table 2-1. Mitochondrial genome GenBank accession information for species examined in phylogenetic analyses ...... 40

Table 2-2. Characteristics of genes and other elements in the mitochondrial genome of N. tropicalis...... 42

Table 2-3. Base composition of mitochondrial genome regions for N. tropicalis ...... 44

Table 2-4. Codon composition based on an average of 5,575 codons in 13 protein-coding genes of N. tropicalis ...... 46

Table 3-1. Summary information for microsatellite multiplex assay comprised of 24 polymorphic loci...... 59

Table 4-1. Overall sample size and partitioned according to sex, location where individuals where first observed, and year when first identified ...... 120

Table 4-2. Genotypes integrated in this study ...... 121

Table 4-3. Distribution of private alleles identified at 17 loci in multiple time partitions ...... 124

Table 4-4. Summary of genetic diversity estimates for all loci of entire population, according to partitioned time increments and over entire time of study ...... 125

Tables 4-5 (A-E). Genetic diversity summary per island population for all data partitions ...... 126

Tables 4-6 (A-E). Genetic diversity values for pairwise comparisons of Fst and analogous estimators between populations in the “All years” data set ...... 131

Tables 4-7 (A-E). Matrices of P-values for pairwise ’s Exact test comparisons of genotype distributions between island populations ...... 133

Table 4-8. First generation migrant (F0) assignments across islands ...... 134

Table 4-9 (A-E). Results of bidirectional estimates of relative migration rates using the divMigrate function in the diveRsity R package ...... 136

Tables 4A-1 (A-D). Pairwise comparisons of Fst and analogous genetic diversity estimators with lower and upper bounds of 95% confidence intervals for all data partitions ...... 141

vii

Table 5-1. Common assumptions underlying the use of translocation as a management tool for imperiled populations, and associated questions of relevance to management and genetic research of the Hawaiian monk seal ...... 160

viii

LIST OF ABBREVIATIONS

A adenine aDNA ancient DNA bp base pair

C cytosine

DNA deoxyribonucleic acid et al. et alia g gram

G guanine

M moles mg milligram min minute mL milliliter nM nanomoles ori origin of replication

PCR polymerase chain reaction

RNA ribonucleic acid rRNA ribosomal RNA s second t time

T thymine tRNA transfer RNA

ix

µg microgram

µL microliter

µM micromolar

x

ACKNOWLEDGMENTS

I am thankful for the support and encouragement of many mentors, colleagues, friends, and family. First and foremost, I thank my advisor, Dr. George Amato for his steadfast support, encouragement, and optimism. He generously ensured that I had access to all of the resources and moral support necessary to pursue my research, and I appreciate his capacity to infuse rigorous scientific thinking with wit, candor, and humor.

I would also like to give special thanks to the members of my dissertation committee—

Dr. Don Melnick, Dr. Rob DeSalle, Dr. Howard Rosenbaum, and Dr. Eleanor Sterling— for their thoughtful feedback through various phases of my dissertation research. I am also thankful to Dr. Sterling and Dr. Josh Ginsberg for all that I learned in the years I spent as a member of their lab group. I extend many thanks to Dr. Susan Perkins, who generously provided mentorship when I was considering my options for dissertation research, as well as office space during my time at the American Museum of Natural

History (AMNH). I also thank Dr. Claudia Wultsch for her mentorship, support, and friendship.

I am deeply in debt to the members of the Hawaiian Monk Seal Research Program for the opportunity to study Hawaiian monk seals; for their insights, support, and great company in the field and lab; and for warmly welcoming me during my time in .

It was a privilege to work with Dr. Charles Littnan, Dr. Jennifer Schultz, Dr. Michelle

Barbieri, Angie Kaufman, Dr. Stacie Robinson, Mark Sullivan, Dr. Jason Baker, Dr.

Albert Harting, Thea Johanos, Jessie Lopez, Whitney Taylor, the crew of the Oscar Elton

Sette, and all of the volunteers who were collecting valuable data in the field.

xi

I am also very grateful to Dr. Alan Scott and his team at the Johns Hopkins

Genome Core—Beth Morosy, David Mohr, and Peng Zhang—who were generously willing to take a crack at sequencing the mitchondrial genome with old and degraded DNA.

I am indebted to Maria Rios and Mohammad Faiz at AMNH, and Lourdes

Gautier, Maria Estrada, Jae McFadden, Dr. Amy Kohn, Alexandra Vamanu, and Thomas

Tarduagno at Columbia University for their myriad forms of administrative and academic support, and for always sparing time to share their smiles and positive energy. Regardless of the business at hand, it is always enjoyable interacting with them.

I also want to express my gratitude to Dr. Francine Kershaw, Dr. Megan Cattau,

Dr. Rae Wynn-Grant, Dr. Natalia Rossi, Yili Lim, Sara Lizzo, Dr. Bryan Falk, Dr.

Antonia Florio, Stephen Gaughran, Dr. Marcia Macedo, Dr. Robert Muscarella, Silvia

Muscarella, Dr. Georgina Cullman, Dr. Mary Blair, Dr. Dan Flynn, Mika Cheng, Dr. Kari

Schmidt, Dr. Camilo Sanin, Dr. Jonathan , Dr. Bernardo Santos, Yiru Cheng, Lais

Coelho, Sarah Guindre-Parker, Stephanie Schmiege, Dr. Danielle Tufts, and the rest of the E3B PhD student community for their friendship and camaraderie during my tenure at

Columbia and AMNH. I feel very lucky to have crossed paths with these fantastic human beings and I look forward to keeping in touch as our careers and lives proceed. I also want to acknowledge my dearest friends outside of school: Alex, Chris, Dave, Devon,

Drew, Gaelin, Jen, Jennie, Julia, Kari, Lia, Linda, Mandy, Melinda, Stephanie, and

Tommy, all of whom have put up with long distances, long stretches of time between getting together, and/or equally long phone and Skype conversations. I am so grateful for your friendship.

xii

Finally, there aren’t enough expressions of appreciation to extend to my parents for their love and support, and for encouraging me to live passionately and courageously.

So I will just have to settle for one, great big, THANK YOU.

xiii

DEDICATION

I dedicate this dissertation to my grandparents and step-grandfather: Vivienne L.

“Baba” Swanton, George J. “Jack” Arnold, Pelageia F. Michnowiec, Peter S.

Michnowiec, and Joseph S. “Smokey” Swanton.

xiv

CHAPTER 1. INTRODUCTION

Biodiversity loss has been occurring at a pace far greater than background rates inferred from pre-human fossil records, largely as a result of human activities (Pimm et al., 1995; Wilson, 2001). There is mounting evidence that suggests that these anthropogenic activities can induce rapid evolutionary change (Dulvy et al., 2003; Mace and Purvis, 2008; Stockwell et al., 2003), thus presenting scientists in the field with one of the greatest challenges ever encountered (Myers, 1996).

Studies of species extinction patterns have found that tends to be phylogenetically non-random, and have suggested that entire clades are likely to be disproportionately lost in the future due to shared life history traits that render closely- related species vulnerable to extinction (Cardillo et al., 2008; Mace and Purvis, 2008;

McKinney, 2006; Purvis et al., 2000b). Associated with biodiversity loss are negative consequences such as the loss of genetic variation, and the disruption or degradation of fundamental evolutionary processes (e.g., adaptation, speciation, extinction) (Myers,

1996). To begin to fully understand the implications of these trends and adequately inform conservation action, it is necessary to look more closely at fine-scale factors contributing to extinction risk at the individual, population, and species levels.

Factors that contribute to extinction include small population size, constricted or isolated geographic distribution, reduced effective population size, demographic and environmental stochasticity, allee effects mediated by behavior (e.g., aggression, mate choice), anthropogenic impacts (e.g., fragmentation, unsustainable consumption, climate change), , reduced genetic variation, and inbreeding depression.

1

Recent pedigree data from wild populations, and advances in molecular and analytical tools for tracing patterns of inbreeding, have enabled success in detecting inbreeding depression within and among populations (Liberg et al., 2005; Raga, 2002; Saccheri et al., 1998). Levels of inbreeding depression in small populations have been found to vary across taxa, populations, and environments, but are often substantial enough to influence factors such as birth weight, survival, reproduction, and resistance to disease, thereby impacting both individual and population performance (Keller and Waller, 2002). This then compromises resilience to environmental and demographic stochasticity as well as anthropogenic threats, thus increasing the vulnerability of the populations to further decline. Such relationships can occur in a mutually-reinforcing feedback loop that may ultimately lead to species extinction, a process known as the ‘extinction vortex’ (Soulé,

1987; Soulé and Mills, 1998).

For many species the extinction vortex is set into motion or exacerbated by unsustainable use by humans (Courchamp et al., 2006; Rosser and Mainka, 2002), which is considered one of the most serious conservation threats in marine environments and is the source of many ongoing conservation problems for marine (Reeves and

Reijnders, 2002). Pinniped species have been especially impacted by heavy harvesting levels due to demands for food, leather, , and blubber, particularly between the seventeenth and early twentieth centuries. Since then, the pressure from harvesting has been significantly curtailed through protective legislation and international treaties regulating harvest and habitat protection, enabling some species (e.g., the northern (Mirounga angustirostris), the California sea ( californianus), and the ( vitulina)) to recover (Bonnell and Selander, 1974; Hoelzel,

2

1999). Yet, other populations and species are failing to recover, perhaps most notably in tribe Monachini, which is comprised of the extinct Caribbean monk seal (Neomonachus tropicalis), the endangered Hawaiian monk seal (Neomonachus schauindslandi), and the endangered (Monachus monachus).

The Phylogenetic Relationships of Monk Seals

Numerous hypotheses regarding the evolutionary history of have been based on morphological, fossil, or molecular evidence (Arnason et al., 2006; Davis et al.,

2004; Fyler et al., 2005; Higdon et al., 2007; Scheel et al., 2014; Wyss, 1988), with varying degrees of agreement. For example, N. schauinslandi and N. tropicalis share more morphological similarities than either does with M. monachus, suggesting a more recent divergence time for N. schauinslandi and N. tropicalis (Higdon et al., 2007).

Molecular studies have identified a large divergence time between M. monachus and N. schauinslandi (Higdon et al., 2007). Yet, the oldest monachine fossil, dating to approximately 14.5 MYA, was found along the Atlantic coast of the southeastern United

States (the historical range of N. tropicalis), and exceeds the age of the oldest monachine fossils found in Europe, thus suggesting that N. tropicalis may be more basal than either of its two sister species (assuming that no other common ancestors remain to be discovered in the fossil record).

The Population Status of Monk Seals

Tribe Monachini (genus Monachus and genus Neomonachus) is the only group of modern phocid seals that inhabits temperate subtropical waters (Higdon et al., 2007).

Historical records of M. monachus observations date back to 2,500 years ago, when the species range was continuous from the Black and Mediterranean seas to the eastern

Atlantic Ocean (McClenachan and Cooper, 2008). The first records of a severe

3

population decline in the eastern Mediterranean appeared during the Roman era, when monk seals were a valuable source of oil, fur, and medical products (Johnson and

Lavigne, 1999). Declines followed for populations in the western Mediterranean and eastern Atlantic, due to sealing activity by the Spanish and Portuguese that continued until the sixteenth century (International Union for the Conservation of Nature, 2016).

Deliberate killing by fishermen due to perceived losses of catch and fishing gear damage has also been common since the Roman era and continues today (Gucu et al., 2004). The two main remaining populations are distributed at the extreme limits of the historical range and are both at the brink of extinction. Along the coast, approximately 130 individuals comprise what is thought to be the only colony that retains a social structure believed to be natural to M. monachus (Forcada et al., 1999). Over

4,000 km away, the second population is composed of 250-300 individuals, which inhabit the eastern along the coast of and , as well as several small fragmented groups that can be found along islands in the Ionian and Aegean Seas

(International Union for the Conservation of Nature, 2016).

N. tropicalis was once widely distributed throughout the western , the Gulf of , and the . Fossil and sighting records suggest that its range was bound by the southeastern United States to the north, and Venezuela to the south (McClenachan and Cooper, 2008). Population estimates based on a synthesis of historical accounts suggest that between 233,000 and 338,000 seals were once distributed among 13 breeding colonies throughout the West Indies (Figure 1) (McClenachan and

Cooper, 2008). With the European colonization of the West Indies came heavy harvesting pressure for food and oil until the late 1890s, which caused chronic range

4

constriction and population decline. The last individual was observed in the wild in 1952, and in 2008, the species was officially declared extinct.

The range of N. schauinslandi includes the entire Hawaiian Archipelago, which spans 2,500 km. The archipelago is distinguished by two groups of atolls, islets, and islands, which are classified as the Northwest Hawaiian Islands (NWHI, also designated as Papahānaumokuākea Marine National Monument), and the main Hawaiian Islands

(MHI). Although historical counts of total population size are limited (Schultz et al.,

2011), evidence suggests that the population has undergone at least three serious declines.

The first decline occurred in subpopulations inhabiting the MHI, where harvest by

Polynesian colonizers resulted in extirpation during the fifth and sixth centuries (Schultz et al., 2009). A second decline followed shortly after the arrival of European sailors during the nineteenth century, when all six primary monk seal subpopulations occupying the NWHI were hunted nearly to extinction. Despite several decades of population recovery during the first half of the twentieth century, observations recorded since the late 1950s show that the population is once again declining. Low juvenile survival and female reproductive rates are the proximate causes of the contemporary population decline (Schultz et al., 2010), but its ultimate causes remain unclear. Potential influences of genetic factors on disease epidemics, juvenile mortality, and reproductive rates are a top concern (Schultz et al., 2009).

Currently, six primary subpopulations are found in the Northwest Hawaiian

Islands, at , , Lisianski, Pearl and Hermes Reef, Midway

Atoll, and . Each year a small number of individuals are also observed on

Necker and Nihoa Islands, and at Maro Reef, Gardner Pinnacles, and occasionally

5

Johnston Atoll. Documented observations also suggest that recolonization and population growth are occurring on the Main Hawaiian Islands (Baker and Johanos, 2004). The collection of subpopulations throughout the Hawaiian archipelago comprises a panmictic metapopulation (Schultz et al., 2011). Approximately 1250 individuals remain. Models predict that if current population trends continue, the species could become extinct within

50 years (International Union for the Conservation of Nature, 2016).

Long term monitoring of N. schauinslandi across its core range of the

Northwestern Hawaiian Islands has enabled spatially and temporally comprehensive observations of reproduction and survival. More than 85% of the pups weaned since the

1980s have been marked in their birth year, and re-sighted throughout their lifetimes, resulting in coverage spanning four to five generations (Baker and Thompson, 2007).

Approximately 4200 individuals in the NWHI have been tagged, two of which have been re-sighted in the Main Hawaiian Islands (Baker and Johanos, 2004; Baker and

Thompson, 2007). Genotypes have been processed for approximately 2,400 individuals, among which replicate sampling is expected for approximately 250 individuals.

As large-bodied mammals that were once over-harvested, that occupy a high trophic level in the food chain, exhibit low reproductive rates, and have high habitat specificity, extant monk seal species face a high risk of extinction (Mace and Purvis,

2008; Purvis et al., 2000a, 2000b). Both N. schauinslandi and M. monachus are listed as critically endangered on the International Union for the Conservation of Nature (IUCN)

Red List of Threatened Species (International Union for the Conservation of Nature,

2016), and endangered on the list of species protected by the US Act

(US Endangered Species Act of 1973).

6

The Conservation Genetics of Monk Seals

For extant Monachus species, genetic information has been used to explore phylogenetic relationships, determine population structure, infer demographic events, and quantify genetic diversity, using increasingly more powerful molecular tools as they become available. Genetic diversity is the raw material required for organisms to adapt to environmental change and is therefore essential to population persistence (Frankham,

2002). A drastic decrease in population size results in depletion of the genetic variations found within a population, which has been shown to increase the likelihood for the expression of deleterious recessive alleles, and reduce fitness and time to extinction

(Hedrick, 2001; Spielman et al., 2004).

With the exception of N. tropicalis, whose extinction predates the application of genetic techniques for population monitoring, monk seal genetic diversity has been quantified by measuring haplotype, nucleotide, and allelic diversity, proportions of polymorphic loci, and heterozygosity (for a detailed review, see Schultz, 2011).

Assessments have included comparisons of nucleotide markers observed in non-coding regions of mitochondrial DNA (mtDNA) and in nuclear DNA (microsatellites), which are assumed to be relatively neutral to natural selection. Markers from a major histocompatibility complex (MHC) coding region have also been used in a preliminary study of adaptive genetic variation (Aldridge et al., 2006).

The first indications of low genetic diversity were reported by Kretzman et al.

(1997), who found only three unique haplotypes in mtDNA samples from 50 individuals from the N. schauinslandi population. Using microsatellites, Gemell et al. (1997) also found low levels of variation (three polymorphic loci out of 20 examined) in N. schauinslandi as compared to M. monachus (17 out of 20 loci examined were

7

polymorphic). This difference is noteworthy considering the contrasting population size and structure of each species. In subsequent studies Pastor et al. (2004, 2007) found similar levels of polymorphism by testing a larger number of markers in genetic samples collected from individuals among remaining reproductively isolated M. monachus

(Western Sahara and eastern Mediterranean). Both populations were found to share 52% of alleles, and to have low mean expected heterozygosity (0.38 and 0.32 for Western

Saharan and eastern Mediterranean, respectively). This confirmed that despite the smaller sizes of these populations, they both have greater genetic variability than N. schauinslandi.

Most recently, Schultz et al. (2009, 2011) determined that N. schauinslandi comprises a single, panmictic population that experienced a severe population bottleneck in recent evolutionary history and is now severely depauperate of genetic variation.

Research Focus

Genus Neomonachus (N. tropicalis and N. schauinslandi) offers a valuable opportunity to piece together an understanding of the historical and contemporary genetic influences on population viability, and the implications for conservation and management. To this end, the goal of this dissertation is to develop new tools for improved molecular resolution when using evolutionary and population genetic approaches, and to use them to resolve outstanding uncertainties regarding evolutionary patterns and processes associated with species origination and extinction and population gene flow for Genus Neomonachus.

In Chapter 2, the mitochondrial genome sequence for N. schauinslandi is characterized and used to expand historical perspectives on relationships within tribe

Monachini, and among tribes Monachini, Miroungini, and Lobodontini. It also lends

8

perspective on long-standing debates surrounding more distant relatives in the subfamily

Phocina.

Chapter 3 reports on the estimation, using new nuclear microsatellite tools, of the genetic diversity of N. schauinslandi among 24 newly-identified polymorphic loci of

1,192 individuals, thereby doubling the amount of known genetic regions upon which to base conservation genetic investigations.

In Chapter 4, the conservation genetic status of N. schauinslandi is assessed by integrating genotypes for the newly-described 24 loci with 18 previously described loci, with the findings interpreted with regard to the use of translocation as a management tool to mitigate mortality.

Chapter 5 concludes by synthesizing research outcomes and further considering their relevance to conservation and management. Related and emerging research opportunities are also discussed.

9

References

Aldridge, B., Bowen, L., Smith, B., Antonelis, G., Gulland, F., Stott, J., 2006. Paucity of class I MHC gene heterogeneity between individuals in the endangered Hawaiian monk seal population. Immunogenetics 58, 203–215. doi:10.1007/s00251-005-0069-y

Arnason, U., Gullberg, A., Janke, A., Kullberg, M., Lehman, N., Petrov, E.A., Väinölä, R., 2006. Pinniped phylogeny and a new hypothesis for their origin and dispersal. Molecular Phylogenetics and Evolution 41, 345–354. doi:16/j.ympev.2006.05.022

Baker, J.D., Johanos, T.C., 2004. Abundance of the Hawaiian monk seal in the main Hawaiian Islands. Biological Conservation 116, 103–110. doi:16/S0006-3207(03)00181- 2

Baker, J.D., Thompson, P.M., 2007. Temporal and spatial variation in age-specific survival rates of a long-lived , the Hawaiian monk seal. Proceedings of the Royal Society B: Biological Sciences 274, 407–415. doi:10.1098/rspb.2006.3737

Bonnell, M.L., Selander, R.K., 1974. Elephant seals: genetic variation and near extinction. Science 184, 908–909.

Cardillo, M., Mace, G.M., Gittleman, J.L., Jones, K.E., Bielby, J., Purvis, A., 2008. The predictability of extinction: biological and external correlates of decline in mammals. Proc. R. Soc. B 275, 1441–1448. doi:10.1098/rspb.2008.0179

Courchamp, F., Angulo, E., Rivalan, P., Hall, R.J., Signoret, L., Bull, L., Meinard, Y., 2006. Rarity value and species extinction: the anthropogenic Allee effect. PLoS Biol 4, e415. doi:10.1371/journal.pbio.0040415

Davis, C.S., Delisle, I., Stirling, I., Siniff, D.B., Strobeck, C., 2004. A phylogeny of the extant Phocidae inferred from complete mitochondrial DNA coding regions. Molecular Phylogenetics and Evolution 33, 363–377. doi:10.1016/j.ympev.2004.06.006

Dulvy, N.K., Sadovy, Y., Reynolds, J.D., 2003. Extinction vulnerability in marine populations. and Fisheries 4, 25–64. doi:10.1046/j.1467-2979.2003.00105.x

Forcada, J., Hammond, P.S., Aguilar, A., 1999. Status of the Mediterranean monk seal Monachus monachus in the western Sahara and the implications of a mass mortality event. Mar Ecol Prog Ser 188, 249–261. doi:10.3354/meps188249

Frankham, R., 2002. Introduction to Conservation Genetics. Cambridge University Press, Cambridge, UK.

Fyler, C.A., Reeder, T.W., Berta, A., Antonelis, G., Aguilar, A., Androukaki, E., 2005. Historical biogeography and phylogeny of monachine seals (Pinnipedia: Phocidae) based

10

on mitochondrial and nuclear DNA data. Journal of Biogeography 32, 1267–1279. doi:10.1111/j.1365-2699.2005.01281.x

Gemmell, N.J., Goodman, S.J., Allen, P.J., Reed, J.Z., 1997. Interspecific microsatellite markers for the study of pinniped populations. Molecular Ecology 6, 661.

Gucu, A.C., Gucu, G., Orek, H., 2004. Habitat use and preliminary demographic evaluation of the critically endangered Mediterranean monk seal (Monachus monachus) in the Cilician Basin (Eastern Mediterranean). Biological Conservation 116, 417–431. doi:16/S0006-3207(03)00234-9

Hedrick, P.W., 2001. Conservation genetics: where are we now? Trends in Ecology & Evolution 16, 629–636. doi:16/S0169-5347(01)02282-0

Higdon, J., Bininda-Emonds, O., Beck, R., Ferguson, S., 2007. Phylogeny and divergence of the pinnipeds (: Mammalia) assessed using a multigene dataset. BMC Evolutionary Biology 7, 216. doi:10.1186/1471-2148-7-216

Hoelzel, A.R., 1999. Impact of population bottlenecks on genetic variation and the importance of life-history; a case study of the . Biological Journal of the Linnean Society 68, 23–39. doi:06/bijl.1999.0328

International Union for the Conservation of Nature, 2016. IUCN Red List of Threatened Species. Version 2016.1. [WWW Document]. URL http://www.iucnredlist.org (accessed 10.15.10).

Johnson, W.M., Lavigne, D.M., 1999. Monk seals in antiquity. The Mediterranean monk seal, pp.1-101.

Keller, L.F., Waller, D.M., 2002. Inbreeding effects in wild populations. Trends in Ecology & Evolution 17, 230–241. doi:10.1016/S0169-5347(02)02489-8

Kretzmann, M.B., Gilmartin, W.G., Meyer, A., Zegers, G.P., Fain, S.R., Taylor, B.F., Costa, D.P., 1997. Low genetic variability in the Hawaiian monk seal. 11, 482–490. doi:10.1046/j.1523-1739.1997.96031.x

Liberg, O., Andrén, H., Pedersen, H.-C., Sand, H., Sejberg, D., Wabakken, P., Åkesson, M., Bensch, S., 2005. Severe inbreeding depression in a wild lupus population. Biology Letters 1, 17–20. doi:10.1098/rsbl.2004.0266

Mace, G.M., Purvis, A., 2008. Evolutionary biology and practical conservation: bridging a widening gap. Molecular Ecology 17, 9–19. doi:10.1111/j.1365-294X.2007.03455.x

McClenachan, L., Cooper, A.B., 2008. Extinction rate, historical population structure and ecological role of the Caribbean monk seal. Proc Biol Sci 275, 1351–1358. doi:10.1098/rspb.2007.1757

11

McKinney, M.L., 2006. Correlated non-native species richness of birds, mammals, herptiles and plants: scale effects of area, human population and native plants. Biological Invasions 8, 415–425. doi:10.1007/s10530-005-6418-9

Myers, N., 1996. The biodiversity crisis and the future of evolution. The Environmentalist 16, 37–47. doi:10.1007/BF01325613

Pastor, T., Garza, J.C., Aguilar, A., Tounta, E., Androukaki, E., 2007. Genetic diversity and differentiation between the two remaining populations of the critically endangered Mediterranean monk seal. Conservation 10, 461–469. doi:10.1111/j.1469- 1795.2007.00137.x

Pastor, T., Garza, J.C., Allen, P., Amos, W., Aguilar, A., 2004. Low genetic variability in the highly endangered Mediterranean monk seal. Journal of Heredity 95, 291–300. doi:10.1093/jhered/esh055

Pimm, S.L., Russell, G.J., Gittleman, J.L., Brooks, T.M., 1995. The future of biodiversity. Science 269, 347–350. doi:10.1126/science.269.5222.347

Purvis, A., Agapow, P.-M., Gittleman, J.L., Mace, G.M., 2000a. Nonrandom extinction and the loss of evolutionary history. Science 288, 328–330. doi:10.1126/science.288.5464.328

Purvis, A., Gittleman, J.L., Cowlishaw, G., Mace, G.M., 2000b. Predicting extinction risk in declining species. Proceedings of the Royal Society of London. Series B: Biological Sciences 267, 1947–1952. doi:10.1098/rspb.2000.1234

Raga, A.J., 2002. Molecular pedigree analysis in natural populations of : approaches, applications, and practical considerations. Canadian Journal of Fisheries & Aquatic Sciences 59, 1696.

Reeves, R.R., Reijnders, P.J.H., 2002. Conservation and management, in: Biology: An Evolutionary Approach, A. Rus Hoelzel, Ed. Blackwell Science, Oxford.

Rosser, A.M., Mainka, S.A., 2002. Overexploitation and species . Conservation Biology 16, 584–586. doi:10.1046/j.1523-1739.2002.01635.x

Saccheri, I., Kuussaari, M., Kankare, M., Vikman, P., Fortelius, W., Hanski, I., 1998. Inbreeding and extinction in a butterfly metapopulation. Nature 392, 491–494. doi:10.1038/33136

Scheel, D.-M., Slater, G., Kolokotronis, S.-O., Potter, C., Rotstein, D., Tsangaras, K., Greenwood, A., Helgen, K.M., 2014. Biogeography and of extinct and

12

endangered monk seals illuminated by ancient DNA and skull morphology. ZooKeys 409, 1–33. doi:10.3897/zookeys.409.6244

Schultz, J.K., 2011. Population genetics of the monk seals (Genus Monachus): a review. Aquatic Mammals 37, 227–235.

Schultz, J.K., Baker, J.D., Toonen, R.J., Bowen, B.W., 2009. Extremely low genetic diversity in the endangered Hawaiian monk seal (Monachus schauinslandi). Journal of Heredity 100, 25–33. doi:10.1093/jhered/esn077

Schultz, J.K., Baker, J.D., Toonen, R.J., Harting, A.L., Bowen, B.W., 2011. Range-wide genetic connectivity of the Hawaiian monk seal and implications for translocation. Conservation Biology 25, 124–132. doi:10.1111/j.1523-1739.2010.01615.x

Schultz, J.K., Marshall, A.J., Pfunder, M., 2010. Genome-wide loss of diversity in the critically endangered Hawaiian monk seal. Diversity 2, 863–880. doi:10.3390/d2060863

Soulé, M.E., 1987. Viable populations for conservation. Cambridge University Press.

Soulé, M.E., Mills, L.S., 1998. No need to isolate genetics. Science 282, 1658–1659. doi:10.1126/science.282.5394.1658

Spielman, D., Brook, B.W., Frankham, R., 2004. Most species are not driven to extinction before genetic factors impact them. Proceedings of the National Academy of Sciences of the United States of America 101, 15261–15264. doi:10.1073/pnas.0403809101

Stockwell, C.A., Hendry, A.P., Kinnison, M.T., 2003. Contemporary evolution meets conservation biology. Trends in Ecology & Evolution 18, 94–101. doi:10.1016/S0169- 5347(02)00044-7

US Endangered Species Act of 1973, 1973. , 87 Stat. 884, 16 U.S.C. 1531-1544.

Wilson, E.O., 2001. The diversity of life, New ed. ed. , London.

Wyss, A.R., 1988. Evidence from structure for a single origin of pinnipeds. Nature 334, 427–428. doi:10.1038/334427a0

13

CHAPTER 2. A RECONSTRUCTION OF THE MITOCHONDRIAL GENOME

OF THE CARRIBEAN MONK SEAL

Formatted for submission to Gene

Authors A. Nicole Mihnovetsa, b, Stephen J. Gaughranc, David W. Mohrd, Beth Morosyd, Peng Zhangd, George Amatoa, Alan F. Scottd

Author affiliations a Sackler Institute for Comparative Genomics, American Museum of Natural History, New York, New York 10024 b Department of Ecology, Evolution, and Environmental Biology, Columbia University, New York, New York 10027 c Department of Ecology and Evolutionary Biology, Yale University, New Haven, CT 06520 d Johns Hopkins Genomics, McKusick-Nathans Institute of Genetic Medicine, Johns Hopkins University, Baltimore MD 21287

Corresponding author A. Nicole Mihnovets Email: [email protected] Phone: 212-854-8253 Fax: 212-854-8188

14

Abstract

Pinnipeds are marine carnivores comprised of the families Phocidae, Otariidae, and . The Phocidae include northern “true seals”, which are grouped in the subfamily , and southern “true” seals, grouped in the subfamily .

Despite extensive interest in the evolutionary relationships and biogeography of pinnipeds, the phylogenetic placement of the Monachinae has been contentious, as paleontological, neozoological, and incomplete molecular data have given rise to several hypotheses of origin and divergence. Molecular studies of pinnipeds have lacked an especially important puzzle piece in the Monachinae lineage: the mitochondrial genome sequence of the Caribbean monk seal (Neomonachus tropicalis), which was driven to extinction by unsustainable human consumption.

To fill this gap, ancient DNA (aDNA) was extracted from a museum specimen and used to sequence a complete mitochondrial genome (16,726 bp) for Neomonachus tropicalis using the Illumina HiSeq 2500 platform. Typical metazoan genome features were observed: 13 protein-coding genes, 22 tRNA, and two rRNA sequences, as well as

12 intergenic spaces including the control region. The nucleotide composition reflected an A+T bias (60.4%), which was within the range observed in other pinniped species.

The resulting information was used in a concatenated sequence alignment consisting of 21 species to resolve phylogenetic uncertainties for pinnipeds, particularly seals in the subfamily Monachinae. Neighbor-joining, maximum likelihood, and

Bayesian inference methods yielded identical topologies. Within the Monachini, a basal placement of Monachus monachus relative to Neomonachus schauinslandi and

15

Neomonachus tropicalis was confirmed. A branching order (Monachini, (Miroungini,

Lobodontini)) was also strongly supported.

By generating the first complete mitochondrial genome for an extinct pinniped, this study clarifies long-standing uncertainties in pinniped systematics. Molecular data will likely be valuable for further consideration of biogeographic factors associated with the split between cold- and warm-water species among the Monachinae, as well as for investigation into the non-random extinction of closely related species that have been subjected to anthropogenic selection pressures.

Keywords: carnivore, pinniped, Neomonachus tropicalis, extinct, ancient DNA, mitochondrial genome, phylogenetic relationships

16

Introduction

The Caribbean monk seal (Neomonachus tropicalis, Figure 1) was once widely distributed along the coasts of the Caribbean Sea, reaching as far east as Guyana and the

Lesser Antilles (Adam, 2004; Adam and Garcia, 2003; McClenachan and Cooper, 2008;

Timm et al., 1997). Population estimates based on a synthesis of historical accounts indicate that between 233,000 and 338,000 seals were once distributed among at least 13 breeding colonies throughout the West Indies (McClenachan and Cooper, 2008). Fossil and historical sighting records suggest that their range was bound by the southeastern

United States to the north, and Venezuela to the south (Adam and Garcia, 2003).

Early European explorers including and Juan Ponce de

Leon recorded the first reports of seal harvests (McClenachan and Cooper, 2008). Settlers of the region harvested seals for food and oil in the following centuries, until the mid-

1880s, when there were no longer sufficient numbers to support industry demands

(McClenachan and Cooper, 2008). In the early to mid-twentieth century, individual seals from remaining colonies were collected for museums and small-scale consumption. The last known wild individual Caribbean monk seal was observed in 1952, and, in 2008, the species was officially declared extinct.

Today, the two extant sister species, the Hawaiian (Neomonachus schauinslandi) and Mediterranean (Monachus monachus) monk seals, are threatened with extinction due to anthropogenic impacts (International Union for the Conservation of Nature, 2016).

These three monk seal species comprise the tribe Monachini, which, along with tribes

Miroungini (elephant seals) and Lobodontini (Antarctic seals), constitute the subfamily

Monachinae (southern “true” seals). This subfamily has been of particular research

17

interest, as the Monachini are a unique lineage of anti-frigid phocids, while the

Miroungini and Lobodontini are cold-water species. The Monachini are adapted to the relatively warm sub-tropical and tropical regions of the Pacific Ocean as well as the eastern and western Atlantic Ocean and the Mediterranean and Caribbean seas; the

Miroungini are distributed in the cooler Pacific waters along the coasts of California and the Mexican peninsula of Baja California; while the Lobodontini are adapted to the extreme conditions of the frigid Antarctic waters (Figure 2).

Despite extensive consideration of pinniped systematics in prior phylogenetic and biogeographic studies, the phylogenetic characterization of the Monachinae has remained elusive, as several lines of evidence, including paleontological, neozoological, and molecular data, have given rise to equivocal hypotheses of Monachinae origin and divergence (Arnason et al., 2006, 1995; Bininda-Emonds and Russell, 1996; Davis et al.,

2004; Fulton and Strobeck, 2010; Fyler et al., 2005; Higdon et al., 2007; Repenning and

Ray, 1977; Scheel et al., 2014; Wyss, 1988a, 1988b). The relevant datasets are not all complete. Those based on morphological characters have been at an advantage due to the broad taxonomic coverage of pinniped fossils, skeletons, and pelts housed in natural history and research institutions. Those based on molecular characters are incomplete due to the extinction of Neomonachus tropicalis and the associated challenge and expense of assembling a mitochondrial genome regions from historical specimens.

Among the various hypotheses posed over the years, the one most commonly suggested by studies based on morphological and molecular characteristics of extant species is that Monachus monachus is sister to Neomonachus schauinslandi and

Neomonachus tropicalis, and that the tribe Monachini is sister to the tribes Miroungini

18

and Lobodontini (Arnason et al., 2006; Bininda-Emonds and Russell, 1996; Fulton and

Strobeck, 2010; Fyler et al., 2005). These relationships were revisited most recently in a study that combined skull morphology with a single mitochondrial gene (cytb) from a museum specimen of Neomonachus tropicalis (Scheel et al., 2014).

Yet, seeking confirmation with a greater number of mitochondrial genes is still warranted, due to two important factors. First, a high intraspecific variability in cranial anatomy is a common characteristic of pinnipeds (Miller et al., 2007); therefore, conclusions based on skull characteristics should be considered with caution (Bininda-

Emonds et al., 1999; Davies, 1958). Second, despite observations showing that phylogenetic analysis of even a single genetic region can be informative (Huelsenbeck et al., 1996), the importance of using more than one molecular marker to ensure phylogenic resolution and precision has been well-documented (Cummings et al., 1995; Duchêne et al., 2011; Zardoya and Meyer, 1996). Indeed, further clarification of the phylogenetic placement of the warm-water Monachini relative to the cold-water Lobodontini and

Miroungini was not possible in analyses that incorporated a single mitochondrial gene from Neomonachus tropicalis (Scheel et al., 2014).

To resolve these uncertainties and expand molecular resolution for future applications, the aims of this study were (1) to sequence and annotate a complete mitochondrial genome for Neomonachus tropicalis, in order (2) to confirm the placement of the Caribbean monk seal relative to the Hawaiian (Neomonachus schauinslandi) and

Mediterranean (Monachus monachus) monk seals, and (3) to resolve the branching order of tribes Monachini, Miroungini, and Lobodontini.

19

Materials and Methods

Tissue sampling and aDNA extraction

Tissue samples were obtained from a Neomonachus tropicalis specimen (ID

14441, Figure 3) collected in 1897 and housed in the collections of the American

Museum of Natural History (AMNH). Sterile blades and forceps were used to collect perioral, periorbital, and nostril tissue samples, which were processed in the AMNH ancient DNA (aDNA) laboratory using strict protocols to prevent contamination

(Willerslev and Cooper, 2005). Total genomic DNA was isolated from the tissue samples using Qiagen DNEasy tissue single-column extraction protocols (Qiagen,

Valencia, CA), with modifications to the manufacturer’s instructions (see supplementary information). The purified genomic aDNA was quantified with a NanoDrop® ND-1000 spectrometer (Thermo Scientific, Wilmington, DE, USA) and sent to the Johns Hopkins

Genetic Resources Core Facility for library preparation and sequencing on the Illumina

HiSeq2500 platform (Illumina, San Diego, CA).

Mitochondrial genome sequencing and assembly

A PCR-free genomic library was created using 50 ng of genomic aDNA. An initial size profiling revealed that the samples were fragmented at 100-300 bp in size; thus, no shearing or initial bead-based cleanup was performed. DNA degradation was mitigated by using a KAPA Hyper Prep Kit (Kapa Biosystems, Woburn, MA), combining the end-repair and A-tailing methods, and performing indexed adapter ligation using standard Illumina indexed adapters and a post-ligation, bead-based cleanup, and size selection to produce a 300 bp insert library. The libraries were paired-end sequenced

20

with a read length of up to 100 bp, using TruSeq Rapid V2 SBS chemistry (Illumina, San

Diego, CA).

To reconstruct the mitogenome of Neomonachus tropicalis, the raw paired-end reads were assembled with MITObim v1.6 (with the optional “-quick” parameter; Hahn et al., 2013), using the mitochondrial sequence of Neomonachus schauinslandi (GenBank

Accession # AM181022.1) as the seed reference. The large size of the initial dataset required using reads that were down-sampled to 5% of the total, from which sequences converged after approximately 10 iterations.

Mitochondrial genome annotation and sequence analysis

The mitochondrial genome of Neomonachus tropicalis was characterized using

GENEIOUS version 8.1.8 (Biomatters, Ltd., Auckland, NZ) (Kearse et al., 2012)), using the mitochondrial genome of Neomonachus schauinslandi as a reference. The putative genome arrangement was further validated using automated de novo annotation analyses available through the MITOS web server (Bernt et al., 2013). Putative tRNAs were validated a third time using the automated tRNAscan-SE server (Lowe and Eddy, 1997), which was also used to generate maps of secondary tRNA structure. A circular map of the full mitochondrial genome was generated by using GENEIOUS version 8.1.8.

Phylogenetic analysis

Fourteen concatenated mitochondrial regions (COI-III, Cytb, NAD1-5, NAD4l,

ATPase6 and ATPase8, 12S, and the control region) of 21 species were used to infer the phylogenetic position of Neomonachus tropicalis. The in-group included 18 species with complete mitogenomes, one species (Monachus monachus) with partial mitochondrial sequences, and Neomonachus tropicalis, with the novel mitogenome sequence reported

21

here (Tables 1-2). The mitogenome of the polar ( maritimus) was constrained as an outgroup.

Mitochondrial sequences were downloaded from the NCBI GenBank database and aligned using MAFFT version 7.017 (Katoh et al., 2002) under default settings for fast Fourier transformation with iterative refinement (FFT-NSi x 1,000) methods.

Several adjustments were made by sight using GENEIOUS version 8.1.8. First, the light- strand-encoded NAD6 gene was removed from all sequences due to the potentially confounding effects of its distinctive nucleotide and amino acid composition (Arnason et al., 2006, 2002). Second, to eliminate gaps, and ensure correspondence with the maximum coverage available in the partial sequence of Monachus monachus, the aligned sequences were further trimmed to 13,073 bp. The aligned sequence data were then analyzed using four models (Akaike information criterion (AIC), AIC corrected for finite sample size (AICc), Bayesian information criterion (BIC), and decision-theoretic performance-based (DT)), using JMODELTEST version 2.1.7 (Darriba et al., 2012;

Posada, 2008) to determine the best-suited evolutionary model for phylogenetic analyses.

Three consensus-tree methods were used to infer phylogenetic relationships.

Distance-based neighbor-joining (NJ) analyses were performed as implemented in

GENEIOUS version 8.1.8 using the HKY genetic distance model (Hasegawa et al., 1985) and a bootstrap resampling method with 100,000 replicates. Character-based analyses were performed using maximum likelihood (ML) and Bayesian inference (BI) methods according to the GTR +Γ+ I nucleotide substitution model (Chai and Housworth, 2011;

Rogers, 2001) identified during JMODELTEST analysis. ML analyses were performed using RAXML (Stamatakis, 2006), with four substitution rate categories and 1,000

22

bootstraps. BI analyses were performed using MRBAYES (Huelsenbeck and Ronquist,

2001; Ronquist and Huelsenbeck, 2003), which calculated 1,100,000 iterations after a

100,000 burn-in period, and subsampled a tree every 200 iterations.

Results and Discussion

Mitochondrial genome sequencing and assembly

A complete mitochondrial genome of Neomonachus tropicalis was assembled from 291M total paired-end reads. A total of 145M read pairs were obtained, 150,000 of which aligned to the Neomonachus schauinslandi mitogenome (Arnason et al., 2006) for a mean depth of coverage of 796x. The mitochondrial sequences of the Neomonachus tropicalis and Neomonachus schauinslandi are 92.2% identical, similar to other interspecific comparisons in phocids (e.g., 94.5% for Leptonychotes weddellii and

Hydrurga leptonyx; 92.4% for Histriophoca fasciata and Pagophilus groenlandicus, per this study).

Genome organization and nucleotide composition

The complete mitochondrial genome of Neomonachus tropicalis (Genbank accession, TBD) is a closed-circular molecule comprised of 16,726 bp (Figure 4). This size is comparable to those of at least 18 other pinniped mitogenome sequences, which range from 16,551 to 16,970 bp (for Pagophilus groenlandicus and Mirounga leonia, respectively). Consensus among the three annotation methods used was achieved in all cases except that of a single tRNA gene that escaped detection by MITOS. As is commonly observed in other metazoan mitogenomes, the mitogenome of Neomonachus tropicalis contains 37 genes, including 13 protein-coding regions, two ribosomal RNA subunits, and 22 transfer RNA (tRNA) genes, as well as a non-coding control region

(Table 2). The heavy strand (H-strand) includes 12 protein-coding genes (COI-III, Cyt-b,

23

NAD1-5, NAD4l, ATPase6, and ATPase8), two rRNA, and 14 tRNA genes. The light strand (L-strand) includes one protein coding (NAD6) and eight tRNA genes.

The overall A+T content was 60.4% (A = 33.8%, T = 26.6%). The bias toward adenine and thymine was within the range observed in other pinniped species (from

58.1% in Phoca largha and sibirica to 60.8% in Mirounga leonina). The overall

A+T content was 58.3% for all rRNA genes and 61.1% for all protein-coding genes, was lowest in the control region (55.6%), and was highest for tRNA genes (63.2%).

Additional summary statistics for the nucleotide composition of each individual region are provided in Table 3.

Protein-coding genes

A total of 13 protein-coding genes (COI-III, Cyt-b, NAD1-6, NAD4l, ATPase6, and ATPase8) were detected in the mitogenome of Neomonachus tropicalis (Tables 2 and

3). The longest reading frame overlaps occurred between genes ATPase6 and ATPase8

(43 bp), and between genes NAD5 and NAD6 (17 bp). Codon usage was estimated at

5,575 codons; lysine (AAA) occurs most frequently (72.0%), and serine (TCG) occurs least frequently (8.2%). The overall nucleotide composition of all protein-coding genes was 33.7% for A, 27.3% for C, 11.6% for G, and 27.5% for T. tRNA and rRNA genes

A total of 22 transfer RNA genes (two encoding leucine, two encoding serine, and one for each of the 18 other amino acids) were identified in the mitogenome of

Neomonachus tropicalis (Table 4). They were found to range in size from 47 bp to 75 bp and total 1,493 bp. The aminoacyl stem of all putative tRNA secondary structures was 7 bp long. Anticodon loops were comprised of 7 bp in all tRNAs except tRNA-Leu(UUR)

24

and tRNA-Gln, which possessed loops that were 9 bp long. Most tRNAs exhibited a typical cloverleaf folding structure, with four exceptions (Figure 5). First, in tRNA

(Ser(AGY)), there were five unpaired nucleotides, whereas a DHU arm might otherwise be expected in other tRNAs. This variation is commonly observed in other animal mitochondria (Watanabe et al., 2014; Wolstenholme, 1992). Second, despite the tRNA(Lys) having the typical four-paired nucleotide pattern in their DHU arm, the arm lacked a nucleotide loop at the end. Reduction in the DHU arm is a common variation in placentals (Janke et al., 1994). Finally, in both tRNA(Met) and tRNA(Thr), the TYC arm was comprised of four paired nucleotides in a loop configuration rather than the expected ladder configuration. These variations may be non-canonical, perhaps due to RNA editing (Börner et al., 1997), or may be errors due to sequencing artifact. The overall nucleotide composition of tRNA genes was 35.4% for A, 27.3% for C, 15.4% for G, and

27.8% for T.

Separated by a tRNA-Val gene, the 12S and 16S rRNA genes were 962 bp and

1,575 bp long, respectively. Their overall nucleotide composition was 31.2% for A,

26.8% for C, 14.9% for G, and 27.2% for T.

Non-coding regions

The compact organization of the Neomonachus tropicalis mitogenome includes a total of 12 intergenic non-coding regions. The largest non-coding region, totaling 1,287 bp, occurred between the tRNA(Pro) and tRNA(Phe) genes, and was identified as the control region / D-loop. The hypervariable “Region II” consists of repeated elements of approximately 450 bp that may be prone to artifacts, due to the difficulty of using short read sequencing methods and ancient DNA. The nucleotide composition of the control

25

region was 28.7% for A, 26.2% for C, 18.3% for G, and 26.9% for T. The other 11 intergenic spaces, ranging in size between one and 36 basepairs, and totaling 84 basepairs, were distributed throughout the mitogenome. The largest intergenic space (36 bp) occurred between NAD4 and tRNA(His).

Phylogenetic analysis

The concatenated nucleotide sequences for 20 ingroup species representing

Phocidae, Otariidae, and Odobenidae, and a single outgroup species representing Ursidae were analyzed using neighbor-joining (NJ), maximum likelihood (ML), and Bayesian inference (BI) methods. Of 88 models tested using JMODELTEST, the GTR +Γ+ I nucleotide substitution model was the most highly ranked by all four (AIC, AICc, BIC,

DT) ranking parameters (-lnL = 83888.33; gamma distribution shape = 1.4360; transition/transversion (titv) ratio = 12.52), and was thus implemented in the ML and BI analyses.

The NJ, ML, and BI phylogenetic methods yielded identical topologies (Figure 6) that corroborate previous hypotheses for the monk seals (Arnason et al., 2006; Davies,

1958; Fulton and Strobeck, 2010; Fyler et al., 2005; Scheel et al., 2014). The split between Neomonachus tropicalis and Neomonachus schauinslandi was supported by

100% bootstrap consensus and 100% posterior probability. The position of Monachus monachus as the earliest branching lineage within tribe Monachini was also well supported (86.55% and 99% bootstrap support, 100% posterior probability).

Furthermore, the basal divergence of tribe Monachini from tribes Miroungini and

Lobodontini received maximum support (100% bootstrap support and posterior probability), and the split between Miroungini and Lobodontini was strongly supported

26

(99.92% and 100% bootstrap support, 100% posterior probability). This reinforces the findings of Fulton and Strobeck (2010), and clarifies the uncertainty that could not be resolved by Scheel et al. (2014), by using a single mitochondrial marker.

Finally, it is worth noting that within the subfamily Phocina, a paraphyletic

(Phoca (Haliochoerus, Pusa)) arrangement corroborates the previous findings of Davis et al. (2004) and Delisle and Strobeck (2005), and of an alternate scenario reported by

Fulton and Strobeck (2010). However, strong support for the relationship between

Haliochoerus and Pusa is still lacking (71.21% and 71% bootstrap support, 100% posterior probability). As discussed in previous studies (see Fulton and Stroebeck, 2010;

Higdon et al., 2007), this finding again calls into question the definitions of genera within the subfamily Phocina, and so warrants additional attention.

Conclusion

This study is the first to use ancient DNA to characterize a complete mitochondrial genome sequence for the extinct Caribbean monk seal (Neomonachus tropicalis). The mitogenome’s size and organization are highly similar to other pinniped species, and are typical for vertebrates in general. This study is also the first to incorporate an extinct species into a phylogenetic assessment for pinnipeds at the mitogenome scale, which resulted in strongly reinforcing previously hypothesized relationships within tribe Monachini.

The additional resolution generated by the Neomonachus tropicalis genome resulted in improved capacity for phylogenetic inference and generated strong support for the basal placement of Monachini relative to Miroungini and Lobodontini. The need for further resolution of relationships within subfamily Phocina was also highlighted. These results are likely to be valuable for future total-evidence-based studies seeking to

27

integrate fossil, molecular, and contemporary-morphological data for further understanding of origin and divergence among pinniped species.

Acknowledgements

We thank the AMNH Mammalogy Department for access to Caribbean monk seal specimens, Jonathan Foox for valuable discussions regarding genome annotation,

Rebecca Hersch for assistance with assessing aDNA quality, and Dr. Anthony Caragiulo for general aDNA lab support. Laboratory work was made possible in part by generous supporters of the McKusick-Nathans Institute of Genetic Medicine at Johns Hopkins

University, and the Sackler Institute for Comparative Genomics at the American Museum of Natural History.

28

References

Adam, P.J., 2004. Monachus tropicalis. Mamm. Species 1–9. doi:10.1644/747

Adam, P.J., Garcia, G.G., 2003. New information on the natural history, distribution, and skull size of the extinct West Indian monk seal, Monachus tropicalis. Mar. Mammal Sci. 19, 297–317. doi:10.1111/j.1748-7692.2003.tb01110.x

Allen, J.A., 1887. The West Indian seal. Science. 9, 59–59. doi:10.1126/science.ns- 9.207.59

Arnason, U., Adegoke, J.A., Bodin, K., Born, E.W., Esa, Y.B., Gullberg, A., Nilsson, M., Short, R.V., Xu, X., Janke, A., 2002. Mammalian mitogenomic relationships and the root of the eutherian tree. Proc. Natl. Acad. Sci. 99, 8151–8156. doi:10.1073/pnas.102164299

Arnason, U., Bodin, K., Gullberg, A., Ledje, C., Suzette, M., 1995. A molecular view of pinniped relationships with particular emphasis on the true seals. J. Mol. Evol. 40, 78–85. doi:10.1007/BF00166598

Arnason, U., Gullberg, A., Janke, A., Kullberg, M., Lehman, N., Petrov, E.A., Väinölä, R., 2006. Pinniped phylogeny and a new hypothesis for their origin and dispersal. Mol. Phylogenet. Evol. 41, 345–354. doi:16/j.ympev.2006.05.022

Bernt, M., Donath, A., Jühling, F., Externbrink, F., Florentz, C., Fritzsch, G., Pütz, J., Middendorf, M., Stadler, P.F., 2013. MITOS: Improved de novo metazoan mitochondrial genome annotation. Mol. Phylogenet. Evol., Mitogenomics and Metazoan Evolution 69, 313–319. doi:10.1016/j.ympev.2012.08.023

Bininda-Emonds, O.R.P., Gittleman, J.L., Purvis, A., 1999. Building large trees by combining phylogenetic information: a complete phylogeny of the extant Carnivora (Mammalia). Biol. Rev. 74, 143–175. doi:10.1111/j.1469-185X.1999.tb00184.x

Bininda-Emonds, O.R.P., Russell, A.P., 1996. A morphological perspective on the phylogenetic relationships of the extant phocid seals (Mammalia: Carnivora: Phocidae). Bonn. Zool. Monogr. 41, 52–56.

Börner, G.V., Yokobori, S., Mörl, M., Dörner, M., Pääbo, S., 1997. RNA editing in metazoan mitochondria: staying fit without sex. FEBS Lett. 409, 320–324. doi:10.1016/S0014-5793(97)00357-8

Cadotte, M.W., Davies, J.T., 2010. Rarest of the rare: advances in combining evolutionary distinctiveness and scarcity to inform conservation at biogeographical scales. Divers. Distrib. 16, 376–385. doi:10.1111/j.1472-4642.2010.00650.x

29

Chai, J., Housworth, E.A., 2011. On Rogers’ proof of identifiability for the GTR + Γ + I model. Syst. Biol. syr023. doi:10.1093/sysbio/syr023

Cummings, M.P., Otto, S.P., Wakeley, J., 1995. Sampling properties of DNA sequence data in phylogenetic analysis. Mol. Biol. Evol. 12, 814–822.

Darriba, D., Taboada, G.L., Doallo, R., Posada, D., 2012. jModelTest 2: more models, new heuristics and parallel computing. Nat. Methods 9, 772–772. doi:10.1038/nmeth.2109

Davies, J.L., 1958. geography and the distribution of northern pinnipeds. Ecology 39, 97–113. doi:10.2307/1929971

Davis, C.S., Delisle, I., Stirling, I., Siniff, D.B., Strobeck, C., 2004. A phylogeny of the extant Phocidae inferred from complete mitochondrial DNA coding regions. Mol. Phylogenet. Evol. 33, 363–377. doi:10.1016/j.ympev.2004.06.006

Delisle I, Strobeck C., 2005. A phylogeny of the (order Carnivora) based on 12 complete protein-coding mitochondrial genes. Mol Phylogenet Evol. 37: 192-201. 10.1016/j.ympev.2005.04.025

Duchêne, S., Archer, F.I., Vilstrup, J., Caballero, S., Morin, P.A., 2011. Mitogenome phylogenetics: the impact of using single regions and partitioning schemes on topology, substitution rate and divergence time estimation. PLOS ONE 6, e27138. doi:10.1371/journal.pone.0027138

Fulton, T.L., Strobeck, C., 2010. Multiple fossil calibrations, nuclear loci and mitochondrial genomes provide new insight into biogeography and divergence timing for true seals (Phocidae, Pinnipedia). J. Biogeogr. 37, 814–829. doi:10.1111/j.1365- 2699.2010.02271.x

Fyler, C.A., Reeder, T.W., Berta, A., Antonelis, G., Aguilar, A., Androukaki, E., 2005. Historical biogeography and phylogeny of monachine seals (Pinnipedia: Phocidae) based on mitochondrial and nuclear DNA data. J. Biogeogr. 32, 1267–1279. doi:10.1111/j.1365-2699.2005.01281.x

Hasegawa, M., Kishino, H., Yano, T., 1985. Dating of the human-ape splitting by a molecular clock of mitochondrial DNA. J. Mol. Evol. 22, 160–174.

Higdon, J., Bininda-Emonds, O., Beck, R., Ferguson, S., 2007. Phylogeny and divergence of the pinnipeds (Carnivora: Mammalia) assessed using a multigene dataset. BMC Evol. Biol. 7, 216. doi:10.1186/1471-2148-7-216

Huelsenbeck, J.P., Bull, J.J., Cunningham, C.W., 1996. Combining data in phylogenetic analysis. Trends Ecol. Evol. 11, 152–158. doi:10.1016/0169-5347(96)10006-9

30

Huelsenbeck, J.P., Ronquist, F., 2001. MRBAYES: Bayesian inference of phylogenetic trees. Bioinformatics 17, 754–755. doi:10.1093/bioinformatics/17.8.754

International Union for the Conservation of Nature, 2016. IUCN Red List of Threatened Species. Version 2016.1. [WWW Document]. URL http://www.iucnredlist.org (accessed 07.15.16).

Janke, A., Feldmaier-Fuchs, G., Thomas, W.K., Haeseler, A. von, Pääbo, S., 1994. The marsupial mitochondrial genome and the evolution of placental mammals. Genetics 137, 243–256.

Katoh, K., Misawa, K., Kuma, K., Miyata, T., 2002. MAFFT: a novel method for rapid multiple sequence alignment based on fast Fourier transform. Nucleic Acids Res. 30, 3059–3066. doi:10.1093/nar/gkf436

Kearse, M., Moir, R., Wilson, A., Stones-Havas, S., Cheung, M., Sturrock, S., Buxton, S., Cooper, A., Markowitz, S., Duran, C., Thierer, T., Ashton, B., Meintjes, P., Drummond, A., 2012. Geneious Basic: An integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics 28, 1647–1649. doi:10.1093/bioinformatics/bts199

Lowe, T.M., Eddy, S.R., 1997. tRNAscan-SE: a program for improved detection of transfer RNA genes in genomic sequence. Nucleic Acids Res. 25, 955–964.

McClenachan, L., Cooper, A.B., 2008. Extinction rate, historical population structure and ecological role of the Caribbean monk seal. Proc. R. Soc. B Biol. Sci. 275, 1351–1358. doi:10.1098/rspb.2007.1757

Miller, E.H., Sung, H.-C., Moulton, V.D., Miller, G.W., Finley, J.K., Stenson, G.B., 2007. Variation and integration of the simple mandibular postcanine dentition in two species of phocid seal. J. Mammal. 88, 1325–1334. doi:10.1644/06-MAMM-A-243R.1

Posada, D., 2008. jModeltest: phylogenetic Model averaging. Mol. Biol. Evol. 25, 1253– 1256. doi:10.1093/molbev/msn083

Repenning, C., Ray, C., 1977. The origin of the Hawaiian monk seal. Proc. Biol. Soc. Wash. 89, 667–688.

Rogers, J.S., 2001. Maximum likelihood estimation of phylogenetic trees is consistent when substitution rates vary according to the invariable sites plus gamma distribution. Syst. Biol. 50, 713–722. doi:10.1080/106351501753328839

Ronquist, F., Huelsenbeck, J.P., 2003. MrBayes 3: Bayesian phylogenetic inference under mixed models. Bioinformatics 19, 1572–1574. doi:10.1093/bioinformatics/btg180

31

Scheel, D.-M., Slater, G., Kolokotronis, S.-O., Potter, C., Rotstein, D., Tsangaras, K., Greenwood, A., Helgen, K.M., 2014. Biogeography and taxonomy of extinct and endangered monk seals illuminated by ancient DNA and skull morphology. ZooKeys 409, 1–33. doi:10.3897/zookeys.409.6244

Stamatakis, A., 2006. RAxML-VI-HPC: maximum likelihood-based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics 22, 2688–2690. doi:10.1093/bioinformatics/btl446

Timm, R.M., Salazar, R.M., Peterson, A.T., 1997. Historical distribution of the extinct tropical seal, Monachus tropicalis (Carnivora: Phocidae). Conserv. Biol. 11, 549–551. doi:10.1046/j.1523-1739.1997.96035.x

Watanabe, Y., Suematsu, T., Ohtsuki, T., 2014. Losing the stem-loop structure from metazoan mitochondrial tRNAs and co-evolution of interacting factors. Front. . 5. doi:10.3389/fgene.2014.00109

Willerslev, E., Cooper, A., 2005. Ancient DNA. Proc. R. Soc. B Biol. Sci. 272, 3–16. doi:10.1098/rspb.2004.2813

Wolstenholme, D.R., 1992. Genetic novelties in mitochondrial genomes of multicellular . Curr. Opin. Genet. Dev. 2, 918–925. doi:10.1016/S0959-437X(05)80116-9

Wyss, A.R., 1988a. On “retrogression” in the evolution of the Phocinae and phylogenetic affinities of the monk seals. Am. Mus. Novit. 1–38.

Wyss, A.R., 1988b. Evidence from flipper structure for a single origin of pinnipeds. Nature 334, 427–428. doi:10.1038/334427a0

Zardoya, R., Meyer, A., 1996. Phylogenetic performance of mitochondrial protein-coding genes in resolving relationships among vertebrates. Mol. Biol. Evol. 13, 933–942.

32

Figure 2-1. N. tropicalis illustration

Illustration of N. tropicalis male, female, and pup published in the Bulletin of the American Museum of Natural History (Allen, 1887).

33

Figure 2-2. Geographical distribution of tribes Monachini, Miroungini, and Lobodontini.

Monachini ranges are highlighted in red from left to right: N. schauinslandi, N. tropicalis, and M. monachus. Map adapted from Fyler et al., 2005.

Monachini

Miroungini

Lobodontini

34

Figure 2-3. N. tropicalis pelt

N. tropicalis pelt housed in the Mammalogy collection of the American Museum of Natural History.

35

Figure 2-4. Mitochondrial genome sequence map of N. tropicalis.

Protein-coding regions are indicated in purple, rRNA regions in red, tRNA regions in pink, and the non-coding D-loop (control region) in orange.

36

Figure 2-5. Inferred secondary cloverleaf structures for the 22 tRNA genes of N. tropicalis

Below each tRNA gene structure is the abbreviation for its corresponding amino acid. Starting at the top, and following in a clockwise manner around the centrally-located variable loop, each tRNA is characterized by its amino acid acceptor arm, TYC arm, anticodon arm, and dihydrouridine (DHU) arm.

37

38

Figure 2-6. Cladogram based on results from neighbor-joining (NJ), maximum likelihood (ML), and Bayesian inference (BI) phylogenetic analyses of 20 pinniped species (ingroup) and one ursid species (outgroup).

Analyses were based on concatenated sequence data (13,073 bp from 12 protein-coding genes, the 12S rRNA gene, one tRNA gene, and the control region). The three methods generated identical topologies. Support values are shown as: NJ bootstrap values / ML bootstrap values / BI posterior probabilities. Monachus monachus is strongly supported as sister to Neomonachus spp. The (Monachini (Miroungini, Lobodontini)) branching order is also strongly supported.

39

Table 2-1. Mitochondrial genome GenBank accession information for species examined in phylogenetic analyses. Family names are underlined. ** Indicates incomplete mitogenome sequence for Monachus monachus assembled from 14 individual gene sequences available on GenBank. § Indicates outgroup used to root phylogenetic trees.

GenBank Species Common Name Source Accession # Phocidae Neomonachus tropicalis Caribbean monk seal TBD This study Monachus monachus Mediterranean monk ** ** Neomonachus schauinslandi Hawaiian monk seal AM181022 Arnason et al. (2006) Mirounga leonine AM181023 Arnason et al. (2006) Lobodon carcinophagus Crab-eater seal AM181024 Arnason et al. (2006) Hydrurga leptonyx seal AM181026 Arnason et al. (2006) Leptonychotes weddellii AM181025 Arnason et al. (2006) Halichoerus grypus NC001602 Arnason et al. (1993) Phoca sibirica NC008432 Arnason et al. (2006) Phoca hispida AM181036 Arnason et al. (2006) Phoca largha KT818831 Bahn et al. (2009) Phoca vitulina Harbor seal AM181032 Arnason et al. (2006) Histriophoca fasciata AM181029 Arnason et al. (2006) Pagophilus groelandicus NC008429 Arnason et al. (2006) Cystophora cristata AM181028 Arnason et al. (2006) Erignathus barbatus AM181027 Arnason et al. (2006) Eumetopias jubatus Steller AJ428578 Arnason et al. (2006) Odobenidae Odobenus rosmarus NC004029 Arnason et al. (2002) Otariiadae Callorhinus ursinus Northern AM181016 Arnason et al. (2006) Zalophus californianus Californian sea lion AM181017 Arnason et al. (2006) Ursidae§ Ursus maritimus AJ428577 Arnason et al. (2002)

40

** Mitochondrial sequences used to assemble a partial mitogenome for M. monachus. GenBank Accession # Gene Source KT935309 Control region Karamandlidis et al. (2016) GU174602 12S rRNA Fulton et al. (2010) AY377142 COX1 Davis et al. (2004) AY377165 COX2 Davis et al. (2004) AY377257 COX3 Davis et al. (2004) AY377327 Cytb Davis et al. (2004) AY377355 ND1 Davis et al. (2004) AY377275 ND2 Davis et al. (2004) AY377211 ND3 Davis et al. (2004) AY377337 ND4 Davis et al. (2004) AY377234 ND4L Davis et al. (2004) AY377373 ND5 Davis et al. (2004) AY377188 ATP8 Davis et al. (2004) AY377304 ATP6 Davis et al. (2004)

41

Table 2-2. Characteristics of genes and other elements in the mitochondrial genome of N. tropicalis. H/L indicate encoding by the heavy or light strand.

Genes/ Elements Strand Position Size (bp) (H/L) From To tRNA-Phe H 1 71 71 12S rRNA H 72 1,033 962 tRNA-Val H 1,034 1,100 67 16S rRNA H 1,101 2,675 1,575 tRNA-Leu(UUR) H 2,676 2,750 75 NADH1 H 2,753 3,709 957 tRNA-Ile H 3,710 3,777 68 tRNA-Gln L 3,775 3,847 73 tRNA-Met H 3,849 3,917 69 NADH2 H 3,918 4,961 1,044 tRNA-Trp H 4,960 5,026 67 tRNA-Ala L 5,037 5,105 69 tRNA-Asn L 5,107 5,179 73 Origin of rep - 5,180 5,215 36 tRNA-Cys L 5,212 5,279 68 tRNA-Tyr L 5,280 5,347 68 COI H 5,349 6,893 1,545 tRNA-Ser(UCN) L 6,891 6,959 69 tRNA-Asp H 6,986 7,032 47 COII H 7,033 7,716 684 tRNA-Lys H 7,718 7,786 69 ATPase 8 H 7,788 7,991 204 ATPase 6 H 7,949 8,629 681 COIII H 8,629 9411 783 tRNA-Gly H 9,413 9,480 68 NADH3 H 9,481 9827 347

42

tRNA-Arg H 9,828 9,896 69 NADH4L H 9,897 10,193 297 NADH4 H 10,187 11528 1342 tRNA-His H 11,565 11,633 69 tRNA-Ser(AGY) H 11,634 11,692 59 tRNA-Leu(CUN) H 11,693 11,762 70 NADH5 H 11,763 13,580 1,818 NADH6 L 13,564 14,091 528 tRNA-Glu L 14,092 14,160 69 cytb H 14,165 15,304 1,140 tRNA-Thr H 15,305 15,374 70 tRNA-Pro L 15,374 15,439 66 Control Region H 15,440 16,726 1,287

43

Table 2-3. Base composition of mitochondrial genome regions for N. tropicalis.

Base Composition (%) G + C A + T Region content content A C G T (%) (%) Protein-coding genes NADH1 32.3 29.2 11.4 27.2 40.5 59.5 NADH2 37.5 28.4 9.0 25.1 37.5 62.5 COI 28.2 24.1 17.0 30.7 41.1 58.9 COII 34.4 24.6 13.9 27.2 38.5 61.5 ATPase 8 40.2 24.0 7.4 28.4 31.4 68.6 ATPase 6 31.0 27.5 10.3 31.3 37.7 62.3 COIII 28.7 28.6 14.6 28.1 43.2 56.8 NADH3 30.8 29.1 12.1 28.0 41.2 58.8 NADH4L 31.0 25.3 12.1 31.6 37.4 62.6 NADH4 33.7 28.3 10.7 27.3 39.0 61.0 NADH5 34.3 27.8 10.1 27.7 38.0 62.0 NADH6 43.2 29.0 9.3 18.6 38.3 61.7 cytb 32.3 29.6 12.4 25.7 42.0 58.0 tRNA genes tRNA-Phe 43.7 18.3 16.9 21.1 35.2 64.8 tRNA-Val 35.8 23.9 16.4 23.9 40.3 59.7 tRNA-Leu(UUR) 29.3 25.3 18.7 26.7 44.0 56.0 tRNA-Ile 41.2 10.3 14.7 33.8 25.0 75.0 tRNA-Gln 41.1 24.7 9.6 24.7 34.2 65.8 tRNA-Met 29.0 27.5 17.4 26.1 44.9 55.1 tRNA-Trp 37.3 17.9 17.9 26.9 35.8 64.2 tRNA-Ala 36.2 24.6 11.6 27.5 36.2 63.8 tRNA-Asn 30.1 28.8 15.1 26.0 43.8 56.2 tRNA-Cys 25.0 22.1 23.5 29.4 45.6 54.4 tRNA-Tyr 29.4 20.6 16.2 33.8 36.8 63.2

44

tRNA-Ser(UCN) 33.3 24.6 15.9 26.1 40.6 59.4 tRNA-Asp 34.0 21.3 12.8 31.9 34.0 66.0 tRNA-Lys 39.1 14.5 13.0 33.3 27.5 72.5 tRNA-Gly 27.9 25.0 19.1 27.9 44.1 55.9 tRNA-Arg 39.1 15.9 10.1 34.8 26.1 73.9 tRNA-His 42.0 13.0 11.6 33.3 24.6 75.4 tRNA-Ser(AGY) 37.3 20.3 16.9 25.4 37.3 62.7 tRNA-Leu(CUN) 42.9 14.3 17.1 25.7 31.4 68.6 tRNA-Glu 36.2 23.2 15.9 24.6 39.1 60.9 tRNA-Thr 31.4 22.9 17.1 28.6 40.0 60.0 tRNA-Pro 37.9 30.3 12.1 19.7 42.4 57.6 rRNA genes 12S rRNA 37.4 23.0 16.9 22.7 39.9 60.1 16S rRNA 39.4 21.8 16.6 22.2 38.5 61.5 Non-coding Control region 28.7 26.2 18.3 26.9 44.4 55.6 Average of protein-coding genes 33.7 27.3 11.6 27.5 38.9 61.1 Average of tRNA genes 35.4 21.3 15.4 27.8 36.8 63.2 Average of rRNA genes 31.2 26.8 14.9 27.2 41.7 58.3 Average of all regions 34.8 23.6 14.3 27.4 37.8 62.2

45

Table 2-4. Codon composition based on an average of 5,575 codons in 13 protein-coding genes of N. tropicalis.

Amino Codon Number Freq (%) Amino Codon Number Freq (%) Acid Acid Ala GCA 49 31.2 Leu CTA 152 26.9 GCC 44 28.0 CTC 95 16.8 GCG 18 11.5 CTG 51 9.0 GCT 46 29.3 CTT 86 15.2 Arg CGA 40 29.6 TTA 134 23.7 CGC 24 17.8 TTG 48 8.5 CGG 25 18.5 Met ATA 141 65.9 CGT 46 34.1 ATG 73 34.1 Asn AAC 157 47.0 Phe TTC 82 48.8 AAT 177 53.0 TTT 86 51.2 Asp GAC 51 43.2 Pro CCA 109 26.4 GAT 67 56.8 CCC 120 29.1 Cys TGC 52 63.4 CCG 47 11.4 TGT 30 36.6 CCT 137 33.2 Gln CAA 175 68.4 Ser AGC 114 19.9 CAG 81 31.6 AGT 50 8.7 Glu GAA 82 58.2 TCA 118 20.6 GAG 59 41.8 TCC 98 17.1 Gly GGA 42 37.5 TCG 47 8.2 GGC 27 24.1 TCT 147 25.6 GGG 22 19.6 Thr ACA 127 26.5 GGT 21 18.8 ACC 130 27.1 His CAC 135 43.4 ACG 60 12.5 176 56.6 ACT 162 33.8 Ile ATC 120 50.6 Trp TGA 60 60.0

46

ATT 117 49.4 TGG 40 40.0 Lys AAA 195 72.0 Tyr TAC 131 42.7 AAG 76 28.0 TAT 176 57.3 Stop AGA 72 16.1 Val GTA 68 44.4 AGG 69 15.4 GTC 31 20.3 TAA 201 45.0 GTG 23 15.0 TAG 105 23.5 GTT 31 20.3

47

Supplementary Information

Ancient DNA extraction methods

Total genomic DNA was isolated from the tissue samples using Qiagen DNEasy tissue single-column extraction protocols (Qiagen, Valencia, CA), and with the following modifications to the manufacturer’s instructions: Prior to extraction, 20 mg of tissue was soaked in 400 uL of phosphate buffer solution for seven days, then transferred to a 1.5 mL microcentrifuge tube. After adding 20 uL of Proteinase K, 1 uL 1molar DTT (dithiothreitol), and 180 uL of Buffer ATL, the tube was gently mixed by repeatedly inverting it for 30 seconds. The mixture was then incubated overnight at 56 °C. Incubation was repeated the following day after adding an additional 10 uL of Proteinase K to ensure complete lysing of the tissue. Following incubation, samples were again gently mixed for 30 seconds before and after adding 200 uL of Buffer AL. This mixture was placed for 15 minutes on a heat block set to 65°C, then gently inverted for 30 seconds before 200 uL of pre-chilled ethanol (100%) was lightly mixed in using a pipette plunger. The mixture was refrigerated for one hour at -4 °C. The entire volume (620 uL) was pipetted into a spin column, placed in a collection tube, and processed with Buffers AW1, AW2, and AE, using standard Qiagen protocols. However, the Buffer AE was pre-warmed on a heat block prior to being added, and the mixture was incubated at room temperature for 40 minutes before being centrifuged. The spin column was then reused for a second addition of Buffer AE. The eluted solution was mixed using the pipette plunger and divided between two new tubes, each containing 75 uL of purified genomic DNA, which was stored at -20 °C in the aDNA lab of the American Museum of Natural History.

48

CHAPTER 3. A NOVEL MICROSATELLITE MULTIPLEX ASSAY FOR

THE ENDANGERED HAWAIIAN MONK SEAL

Published in Conservation Genetics Resources

Authors A. Nicole Mihnovets1,2, Jennifer K. Schultz3, Claudia Wultsch2, Charles L. Littnan4, George Amato2

Author Affiliations 1 Department of Ecology, Evolution, and Environmental Biology, Columbia University in the City of New York, 1200 Amsterdam Avenue, 10th Floor Schermerhorn Extension, New York, New York 10027

2 Sackler Institute for Comparative Genomics, American Museum of Natural History, 79th Street at Central Park West, New York, New York 10024

3 Endangered Species Division, NOAA Fisheries, 1315 East West Highway, Silver Spring, MD 20910

4 Hawaiian Monk Seal Research Program, NOAA Fisheries, 1845 WASP Blvd., Building 176, Honolulu, Hawai’i 96818

Corresponding Author A. Nicole Mihnovets Email: [email protected] Phone: 212-854-8253 Fax: 212-854-8188

49

Abstract

The Hawaiian monk seal (Neomonachus schauinslandi) is endemic to the

Hawaiian archipelago and is among the world’s most endangered marine mammal species. Prior studies used mitochondrial and microsatellite markers to investigate broad scale conservation questions about genetic diversity and population differentiation. As

Hawaiian monk seal population numbers continue to decline, outstanding fine-scale genetic questions require enhanced genotyping capacity of a rich archive of specimens.

Here we integrate 24 novel polymorphic microsatellite markers into three multiplex polymerase chain reactions (PCRs) and provide summary statistics. The mean number of alleles was 2.625; mean expected and observed heterozygosities were 0.359 and 0.349, respectively; and mean polymorphic information content was 0.027. This multiplex assay contributes additional analytical power to further investigate genetic questions related to population recovery of an iconic endangered species.

Key Words

Endangered species, Hawaiian monk seals, marine mammals, population decline, genetic diversity

50

Introduction

The endemic Hawaiian monk seal (HMS) is classified as endangered on the IUCN

Red List (IUCN 2015) and under the U.S. Endangered Species Act (United States Code

1976), and overall population numbers continue to decline (Littnan et al. 2015) despite myriad concerted conservation efforts. Previous HMS genetic studies used genomic markers (Aldridge et al. 2006; Gemmell et al. 1997; Kretzmann et al. 1997, 2001; Schultz et al. 2009, 2010) to characterize a sobering lack of genetic diversity in the remaining population, which reinforced concerns over the impacts of small population size and potential inbreeding depression on population recovery. The ability to further address these concerns with fine scale genetic analyses requires greater genotypic resolution than is currently available, thus necessitating additional microsatellite multiplex assay development, which, for the time being, remains the most cost-effective genetic tool for this study system.

For over three decades, the United States National Marine Fisheries Service’s

Hawaiian Monk Seal Research Program (HMSRP) has systematically and opportunistically collected HMS tissues during seasonal field research expeditions and during treatment or necropsy of sick or stranded animals. We randomly selected 15 samples (from distinct individuals) from the HMSRP tissue collection to design 30 new polymorphic primer pairs (see protocols in Supplementary Materials) for microsatellite genotyping in three multiplexes.

Though they performed well in singleplex methods, six of the 30 primer pairs were problematic (e.g., inconsistent amplification success rates) when included in various multiplex combinations during numerous optimization test runs and were thus excluded

51

from analyses. Nonetheless, we list their accession numbers and basic descriptive information as they can still be of utility. We combined the remaining 24 novel primers in three multiplex mixtures consisting of seven to ten primer sets to perform PCR amplifications.

Methods and Results

We optimized PCR volumes and concentrations (Table 1) of individual primer sets (containing fluorescently labeled forward primers) within primer pre-mixes, to be combined in a 6L reaction comprised of 3L master mix (Qiagen Multiplex PCR Kit),

1.4 L of deionized water, 0.60 L of primer mix, and 1 L of genomic DNA (10x concentration).

The PCR program for multiplexes one and two consisted of initial denaturation for 15 min at 95C, followed by 33 cycles of denaturation for 30 s at 94C, annealing for

3 min at 56C, extension for 30 s at 72C, and final elongation for 30 min at 60C.

Multiplex three required 38 cycles. We ran amplified products on an ABI 3730 Genetic

Analyzer (Applied Biosystems) with the GeneScan 500 LIZ (Applied Biosystems) internal size standard (20L Liz500: 980 L formamide) at the Sackler Institute for

Comparative Genomics and at Molecular Cloning Laboratories. Using GENEMAPPER

5.0 software (Applied Biosystems), we scored alleles for complete genotypes (i.e., at 24 loci) of 1192 samples.

We used MICROCHECKER 2.2.3 (Van Oosterhout et al. 2004) to detect genotypic errors and allelic dropout, and EXCEL MICROSATELLITE TOOKIT 3.1

(Park 2001) to estimate the allelic richness, observed heterozygosity (Ho), expected

52

heterozygosity (He), and polymorphism information content (PIC) of each locus (Table

1).

Using GIMLET 1.3.3 (Valière 2002), we determined that a minimum of five loci are needed to estimate the unbiased probability that a genotype is shared by two individuals in a population (PID), and a minimum of nine loci are needed to estimate the probability that a genotype is shared by two siblings in a population (PID(sib), Fig. 1).

We used GenAlEx 6.5 (Peakall and Smouse 2006, 2012) to test for Hardy-

Weinberg equilibrium (HWE) using a chi-square test statistic and found evidence of heterozygote deficiency for six loci (Table 1). We grouped genotypes according to their geographic origins, and confirmed that the genetic structure among individual loci was not contributing to these significant departures from HWE.

We used FSTAT 2.9.3.2 (Goudet 1995) to estimate the linkage disequilibrium

(LD) between all pairs of loci and found mild evidence of LD (p<0.01) between four combinations of primer sets (Neosch4553 and Neosch10475; Neosch7472 and

Neosch15522; Neosch9043 and Neosch15057; Neosch10475 and Neosch15057). We also used FSTAT to estimate FIS (Weir and Cockerham 1984, Table 1).

We used ML-NULLFREQ (Kalinowski and Taper 2006), to estimate null allele frequencies based on maximum likelihood methods. Null alleles occur at low frequency

(0.002– 0.025) for 13 loci (Table 1) and at medium frequency for one locus

(Neosch15522 = 0.206).

Our multiplex assay consisting of 24 novel polymorphic loci will augment the current capacity to batch process large tissue sample collections, and will enhance

53

analytical power to conduct fine-scale individual and population level genetic analyses for an iconic endangered species.

Acknowledgements

This work was funded in part by the Papahānaumokuākea Marine National

Monument and UNESCO World Heritage Site, the National Marine Fisheries Service’s

Hawaiian Monk Seal Research Program, and the American Museum of Natural History’s

Sackler Institute for Comparative Genomics (SICG). We thank Rebecca Hersch, Stephen

Gaughran, Ramon Perez, Emily Lipstein, and Karoline Lake at SICG, Drs. Johannes

Haugstetter and Andres Buser at Ecogenics GmbH, and Lin Wang and Paul Lee at

Molecular Cloning Laboratories for their invaluable laboratory contributions. We thank

Angela Kaufman and Thea Johanos for their assistance with specimen and data archives.

All HMS samples were collected under NMFS MMPA/ESA Permit Nos. 848-1695,

10137 and 932-1905/MA-009526.

Author’s Contributions

ANM extracted genomic DNA, developed and optimized multiplex PCR recipes, coordinated multiplex assay batch processing, genotyped all samples, conducted analyses, and wrote the manuscript. JKS coordinated primer design and multiplex assay batch processing, extracted genomic DNA, and contributed to manuscript development.

CW contributed to manuscript development. CLL provided biological specimens and contributed to manuscript development. GA provided laboratory supplies and equipment and contributed to manuscript development. All authors read and approved the final manuscript.

54

Conflict of Interest

The authors declare that they have no conflicts of interest.

55

References

Aldridge B, Bowen L, Smith B, et al (2006) Paucity of class I MHC gene heterogeneity between individuals in the endangered Hawaiian monk seal population. Immunogenetics 58:203–215. doi: 10.1007/s00251-005-0069-y

Gemmell NJ, Goodman SJ, Allen PJ, Reed JZ (1997) Interspecific microsatellite markers for the study of pinniped populations. Mol Ecol 6:661.

Goudet J (1995) FSTAT (Version 1.2): A Computer Program to Calculate F-Statistics. J Hered 86:485–486.

IUCN (2015) The IUCN Red List of Threatened Species. Version 2015-3. http://www.iucnredlist.org/details/13654/0.

Kalinowski ST, Taper ML (2006) Maximum likelihood estimation of the frequency of null alleles at microsatellite loci. Conserv Genet 7:991–995. doi: 10.1007/s10592-006- 9134-9

Kretzmann MB, Gemmell NJ, Meyer A (2001) Microsatellite Analysis of Population Structure in the Endangered Hawaiian Monk Seal. Conserv Biol 15:457–466. doi: 10.1046/j.1523-1739.2001.015002457.x

Kretzmann MB, Gilmartin WG, Meyer A, et al (1997) Low Genetic Variability in the Hawaiian Monk Seal. Conserv Biol 11:482–490. doi: 10.1046/j.1523-1739.1997.96031.x

Littnan C, Harting A, Baker J (2015) Neomonachus schauinslandi: The IUCN Red List of Threatened Species 2015: e.T13654A45227978.

Park S (2001) Trypanotolerance in West African cattle and the population genetic effects of selection. Dissertation, University of Dublin.

Peakall R, Smouse PE (2006) GENALEX 6: genetic analysis in Excel. Population genetic software for teaching and research. Mol Ecol Notes 6:288–295. doi: 10.1111/j.1471- 8286.2005.01155.x

Peakall R, Smouse PE (2012) GenAlEx 6.5: genetic analysis in Excel. Population genetic software for teaching and research—an update. Bioinformatics 28:2537–2539. doi: 10.1093/bioinformatics/bts460

Schuelke M (2000) An economic method for the fluorescent labeling of PCR fragments. Nat Biotechnol 18:233–234. doi: 10.1038/72708

56

Schultz JK, Baker JD, Toonen RJ, Bowen BW (2009) Extremely low genetic diversity in the endangered Hawaiian monk seal (Monachus schauinslandi). J Hered 100:25 –33. doi: 10.1093/jhered/esn077 Schultz JK, Marshall AJ, Pfunder M (2010) Genome-wide loss of diversity in the critically endangered Hawaiian monk seal. Diversity 2:863–880. doi: 10.3390/d2060863

United States Code (1976) Endangered Species Act of 1973. 16 U.S.C. 1531 et seq. Valière N (2002) GIMLET: a computer program for analysing genetic individual identification data. Mol Ecol Notes 2:377–379. doi: 10.1046/j.1471-8286.2002.00228.x- i2

Van Oosterhout C, Hutchinson WF, Wills DPM, Shipley P (2004) Micro‐checker: software for identifying and correcting genotyping errors in microsatellite data. Mol Ecol Notes 4:535–538. doi: 10.1111/j.1471-8286.2004.00684.x

Weir BS, Cockerham CC (1984) Estimating F-Statistics for the Analysis of Population Structure. Evolution 38:1358–1370. doi: 10.2307/2408641

57

Figure 3-1. Comparison of the cumulative probability of identity for unrelated individuals (P(ID)) and siblings (P(ID)sib), P(ID) < 0.01

58

Table 3-1. Summary information for microsatellite multiplex assay comprised of 24 polymorphic loci. Annealing temperature (Ta), forward primer fluorescent dye (Dye), reaction concentration (RC) in μM, number of samples genotyped (N), allelic richness (Ar), expected heterozygosity (He), observed heterozygosity (Ho), polymorphic information content (PIC), inbreeding coefficient (FIS) with statistically significant deviations from Hardy-Weinberg equilibrium denoted with * (P<0.050) and ** (P<0.001), and null allele frequency (FrNull). Six polymorphic loci were excluded from final analyses. GenBank accession numbers appear in parentheses under each locus ID.

Repeat Size Ta RC Primer Pair Primer Sequences 5'-3' Dye N Ar He Ho PIC FIS FrNull Motif (bp) (C) (M) Multiplex 1 Neosch00871 F: TCACTCACACCAGGAGAAGG (GT)14 140-144 56 FAM 0.04 1192 3 0.178 0.174 0.162 0.024 0.008 (KU220201) R: CCACGCTAGACATGGTCTTTG 0.04 Neosch01807 F: CCACTTGGAAGCAAAAGTAAAGC (AC)17 95-99 56 FAM 0.03 1192 3 0.497 0.718 0.376 -0.445** 0.000 (KU220202) R: TAGGAAAGCCTTCTCTGGGC 0.03 Neosch03204 F: ACCAGGAGTAGTCCAAACAGG (AC)13 187-189 56 NED 0.03 1192 2 0.389 0.372 0.313 0.046 0.015 (KU220204) R: CACCCGTGTTACAGAGGAAG 0.03 Neosch08948 F: GTATGTTGTTGCAAACGGCAG (TG)15 123-127 56 NED 0.02 1192 3 0.102 0.094 0.097 0.077 0.021 (KU220213) R: AGTGCCTGTTGTGATGGATG 0.02 Neosch10588 F: GATACCAGTGGCCCAAACTC (CA)16 105-111 56 VIC 0.03 1192 3 0.261 0.258 0.227 0.011 0.004

59 (KU220216) R: TCAAGCATTCAACCAAGTGC 0.03

Neosch12980 F: CTGACACCTTCTCTACCGGC (AC)20 218-224 56 VIC 0.03 1192 4 0.422 0.419 0.348 0.006** 0.025 (KU220219) R: CAGTGCCCACGCATGTTTC 0.03 Neosch16869 F: GGGACAATTTCTCTCTCTCCC (CA)20 170 - 172 56 PET 0.03 1192 2 0.388 0.395 0.313 -0.019 0.000 (KU220227) R: ATGTTCTTTAGTCTCACTTCACG 0.03 Multiplex 2 Neosch03980 F: GCTTTGGCTCTCATTCTCAGC (GT)15 94-100 56 VIC 0.03 1192 3 0.177 0.170 0.161 0.035 0.011 (KU220205) R: AACATCACTGCCCACTCTCC 0.03 Neosch04022 F: AGATGTGCATGATGTTACCCAAG (CCAT)11 141-145 56 FAM 0.04 1192 2 0.376 0.353 0.305 0.060* 0.019 (KU220206) R: GTGCCACCAGTTAAAGAGGG 0.04 Neosch04553 F: TGGGTTGACTACCATCGCTC (TG)21 170-172 56 PET 0.04 1192 2 0.470 0.467 0.360 0.007 0.002 (KU220207) R: AGCAAGGGACGTAGTGACAG 0.04 Neosch10909 F: GTGCGTGCTCTCTATTTTCCC (CA)16 113-117 56 NED 0.08 1192 2 0.284 0.289 0.243 -0.017 0.000 (KU220217) R: TGTGCATTCCTGTAGCAGAG 0.08 Neosch11181 F: ACTGGCTTTCCTCACTCCTG (AC)15 222-228 56 FAM 0.04 1192 2 0.378 0.387 0.307 -0.023 0.000 (KU220218) R: CGTGAGAATCAACCCTCTGC 0.04 Neosch14403 F: AGGAGGGATTCCAGCAAGAG (CA)12 297-301 56 FAM 0.04 1192 3 0.165 0.166 0.152 -0.010** 0.013 (KU220220) R: ACAGCAATGGTGATGGGATG 0.04 Neosch18265 F: AGCCACATGCAGAAGGTTTG (TG)16 104-106 56 PET 0.04 1192 2 0.420 0.422 0.332 -0.004 0.000 (KU220229) R: TCTCTAATCATGTGAGCCAATTCC 0.04 Multiplex 3 Neosch04584 F: CGCATTCACGCACTTTTCAC (CA)18 219-223 56 FAM 0.03 1192 3 0.208 0.208 0.187 0.000 0.000

59 blah

Repeat Size Ta RC Primer Pair Primer Sequences 5'-3' Dye N Ar He Ho PIC FIS FrNull Motif (bp) (C) (M) (KU220208) R: CGATGGTAGAGGAGGTTGGG 0.03 Neosch06554 F: TCATTGGTATTCATGTTCCCCC (AC)14 176-178 56 PET 0.03 1192 2 0.496 0.474 0.373 0.044 0.015 (KU220209) R: AGCCCAGGCTTGTTGAAATG 0.03 Neosch07472 F: AGAATGAAAGGCAGACATGCG (CA)17 110-116 56 FAM 0.03 1192 3 0.500 0.477 0.376 0.046 0.015 (KU220211) R: GTCCCTCGAACAGACTGGAG 0.03 Neosch08549 F: AGCCTTGTGATGTCAATCTGC (AC)13 173-179 56 VIC 0.02 1192 3 0.441 0.445 0.344 -0.011 0.000 (KU220212) R: TGCGGACAATTCATCCAAATATTAAC 0.02 Neosch09043 F: GCCCTTCATCACCTGCTTTG (CA)13 157-159 56 FAM 0.03 1192 2 0.499 0.471 0.375 0.058* 0.019 (KU220214) R: GAGCTGTGGCGAAAGACATC 0.03 Neosch10475 F: GAGGATGGCTGAACCAGGAG (GT)18 82-84 56 FAM 0.03 1192 2 0.154 0.160 0.142 -0.037 0.000 (KU220215) R: TGATGCCACAGTCTCTCCAC 0.03 Neosch15057 F: AAATCTTGTGGATTCTTCAGCC (AC)16 79-81 56 VIC 0.03 1192 2 0.305 0.306 0.258 -0.005 0.000 (KU220223) R: AACCAATGGCTGGGTCTGTG 0.03 Neosch15522 F: TGAATAGCACTTTCTCTCCACTTC (AC)22 112-118 56 NED 0.03 1192 4 0.657 0.316 0.583 0.519** 0.206 (KU220224) R: GACCATGAGGGAGGAGCTG 0.03 Neosch15716 F: TTCTCCGACCTTGCCTTACC (AC)13 120-132 56 PET 0.03 1192 4 0.353 0.362 0.291 -0.024 0.000 (KU220225) R: ACACTTTCACTGCCCTTTGC 0.03 Neosch17310 F: TCAAATGGTGGAGGTGAGGG (GT)14 124-126 56 VIC 0.03 1192 2 0.485 0.470 0.367 0.032 0.011 (KU220228) R: ACTTTTGGTATGTGCTCTGTTCTC 0.03

60 Mean 3 0.359 0.349 0.291 0.027 0.016

Singleplexes Neosch02578 F: CTTTGATTCTGGTCGGGTGC (AC)16 229 - 231 56 FAM NA 15 2 NA NA NA NA NA (KU220203) R: AAGTCACATTGGACGTTGGG Neosch07171 F: CTACTCAGGCACAAATGGGG (GT)12 249 - 251 56 VIC NA 15 2 NA NA NA NA NA (KU220210) R: TCTCATGATCCCCTTCCACG Neosch14583 F: TTCCAGCTTCTGGTAGTGGG (TG)20 166 - 170 56 VIC NA 15 3 NA NA NA NA NA (KU220221) R: AGTACAGCCTGACACAGACG Neosch14857 F: TGAAGCCTCTTGACTGGTGG (AC)16 158 - 160 56 VIC NA 15 2 NA NA NA NA NA (KU220222) R: TGCTGTAAGAACATCCAGTGTG Neosch15901 F: TCCATGCACACCCATTTCTC (AC)19 98 - 100 56 PET NA 15 2 NA NA NA NA NA (KU220226) R: GAACGTCCCCACGTGTTTG Neosch18320 F: TGTCCACTTTCACCTGGAGC (GT)13 252 - 256 56 VIC NA 15 2 NA NA NA NA NA (KU220230) R: TCAGATCCACGTCTCCAACC

blah 60

Supplementary Material

To design novel primers, we extracted total genomic DNA from 25 mg subsamples of tissue specimens archived in DMSO (20%) or ethanol (90-100%), using standard single column and 96-well plate Qiagen DNEasy tissue extraction protocols

(Qiagen, Valencia, CA USA), and we contracted with Ecogenics GmbH to isolate microsatellite loci. Size-selected fragments were enriched from genomic DNA for simple sequence repeat (SSR) content by using magnetic streptavidin-coated beads and biotin- labeled GT and CT repeat oligonucleotides. The SSR-enriched library was analyzed on a

Roche 454 platform (454 Life Sciences, a Roche Company, Branford, CT, USA) using

GS FLX titanium reagents (Microsynth AG, Balgach, Switzerland). Following the protocols of M. Schuelke (Schuelke 2000), an 18-bp long M13-tail was added to the 5’ end of each forward primer to determine polymorphism. Allele calls were based on patterns detected in individual DNA samples using an ABI3730 DNA analyzer (Applied

Biosystems, Foster City, CA, USA).

PCR amplifications were set up in a 10L reaction volume of approximately 10ng of DNA template, 1 X Qiagen buffer stock (Qiagen) containing 15 mM MgCl2, 200M of dNTPs, 2 uM of each primer, 0.16 M of fluorescent labeled M13 primer and 0.5 units of

Hot Start taq DNA polymerase. The PCR program consisted of 15 min at 95C, followed by 30 cycles of 30 s at 95C, 45 s at the annealing temperature (Table 1) and 45 s at

72C, followed by 8 cycles of 30 s at 95C, 45 s at 53C, and 45 s at 72C, ending with a final elongation stage of 30 min at 72C.

A total of 18, 551 reads consisted of an average of 219 base pairs. Of these, 2,680 reads contained a microsatellite insert with a dinucleotide of at least ten repeat units or a

61

tri- or a tetra-nucleotide of at least six repeat units. However, only 1,017 reads were suitable for primer design. Out of the 190 primer pairs that were tested, 55 were excluded due to amplification failure. Of the remaining 135 primer pairs, only 30 were polymorphic.

62

CHAPTER 4. REVISITING THE GENETIC DIVERSITY AND CONNECTIVITY

OF THE ENDANGERED HAWAIIAN MONK SEAL

Formatted for BMC Genetics

Authors A. Nicole Mihnovets1,2, Claudia Wultsch2, Jennifer K. Schultz3, Thea C. Johanos4, Stacie J. Robinson4, Charles L. Littnan4, George Amato2

Author Affiliations 1 Department of Ecology, Evolution, and Environmental Biology, Columbia University in the City of New York, 1200 Amsterdam Avenue, 10th Floor Schermerhorn Extension, New York, New York 10027

2 Sackler Institute for Comparative Genomics, American Museum of Natural History, 79th Street at Central Park West, New York, New York 10024

3 Endangered Species Division, NOAA Fisheries, 1315 East West Highway, Silver Spring, MD 20910

4 Hawaiian Monk Seal Research Program, NOAA Fisheries, 1845 WASP Blvd., Building 176, Honolulu, Hawai’i 96818

Corresponding Author A. Nicole Mihnovets Email: [email protected] Phone: 212-854-8253 Fax: 212-854-8188

63

Abstract

Molecular tools enable the assessment of genetic diversity and the inference of population connectivity in imperiled wild populations, thereby contributing valuable insights into the development of conservation strategies. Endangered throughout its range, the Hawaiian monk seal (Neomonachus schauinslandi, HMS) experienced an overall population decline of more than 70% throughout the Northwestern Hawaiian

Islands during the past 60 years. The existence of up to 32% of the current population has been attributed to intensive recovery efforts, and the influence of genetic processes on the capacity for HMS population recovery is an ongoing management concern.

Regardless of sample size or extent of molecular resolution, HMS population genetic studies over the past two decades reported extremely low levels of genetic diversity, along with evidence of high levels of gene flow within the population—albeit with ephemeral cases of mild differentiation. These findings suggest that the use of translocation to promote population recovery would not have a detrimental genetic impact. This study builds upon previous HMS genetic studies by using 42 polymorphic microsatellite loci in 785 seals first encountered on eight islands in 1972 and between

1981 and 2007, to evaluate genetic diversity, population differentiation, and inter-island movement.

Overall levels of genetic diversity were low (AR = 1.93, HE = 0.38). Spatial and temporal comparisons between island populations revealed mild, but insignificant, differences between diversity estimates. Over the time periods examined, increasingly uniform estimates of heterozygosity were observed among island populations.

64

Genetic differentiation was evaluated using population-based and individual- based assignment methods. Population-based assessments based on Fst and four analogous estimators revealed some cases of extremely mild and temporary differentiation between the Northwestern Hawaiian Islands and the main Hawaiian

Islands when the data were partitioned into four timeframes, but not when all of the data were combined.

Overall patterns arising from individual-based analyses of contemporary gene flow were generally suggestive of admixture using STRUCTURE (K = 1) and principal component analyses. Isolation-by-distance and autocorrelation analysis, carried out separately for each sex, revealed nonrandom spatial structuring at close geographic distances (< 50km). Genotypic clustering among males and females was not significantly different, suggesting that dispersal in HMS was not male-biased, the appearance of which may be an artifact of translocation.

In analyses of inter-island movement using GENECLASS, version 2, forty individuals were identified as first generation migrants. Estimates of inter-island movement using the diveRsity R-package generally increased over time, although rates between Laysan and French Frigate shoals were consistently high (ranging from 0.90 to

1.00). Significant evidence of asymmetric (directional) migration was not found in pairwise comparisons of populations.

Results from this study generally align with findings of earlier HMS genetic studies. Across consecutive, complementary studies, concerted effort to maximize the utility of microsatellite markers has culminated in a relatively consistent understanding of the genetic status of the Hawaiian monk seal. There is no evidence to suggest that

65

continuing management under a single population paradigm, or that using translocation to mitigate mortality, would threaten the genetic status of the Hawaiian monk seal.

66

Introduction

Genetic diversity is the raw material required for adaptation in wild populations, and is essential to population persistence [1]. Thus, the factors that influence the genetic status of an endangered species are of interest to managers charged with crafting and implementing effective conservation strategies [2–4]. Genetic diversity is maintained under spatial and environmental constraints when rare alleles are preserved in isolated populations, or is lost because gene flow between populations is prevented [5]. Managers are often concerned with inbreeding depression and the loss of heterozygosity. Lower heterozygosity may reflect a greater likelihood of expressing deleterious recessive alleles, fitness reductions, and reduced time to extinction [6,7]. Higher heterozygosity is correlated with higher evolutionary potential, which is assumed to be important for a species’ capacity to adapt to environmental change or recover from anthropogenic impacts such as over-exploitation [5–7].

Endangered throughout its range, the endemic Hawaiian monk seal

(Neomonachus schauinslandi, HMS) narrowly survived the most recent period of human over-exploitation, which lasted for an estimated eight HMS generations [8]. Despite the initial increase in population size that followed during the first half of the twentieth century, a chronic decline after the 1950s was estimated to portend extinction by the end of the twenty-first century [9]. However, recent estimates have suggested that the rate of population decline may be slowing (Baker et al., unpublished data), and the survival of as much as 32% of the current population has been attributed to successful management interventions [10].

67

Managers with the U.S. National Marine Fisheries Service’s (NMFS) Hawaiian

Monk Seal Research Program (HMSRP) have used translocation to improve survival rates since 1984 [11-12], well before conservation genetics became a commonly-used management tool. With the increasing accessibility and feasibility of molecular genetics tools, studies in subsequent decades have provided valuable perspectives on the genetic status of HMS, including potential implications of translocation and associated gene flow between populations [8, 13–17]. Across these studies, estimates of genetic diversity were consistently low, but revealed no evidence of contemporary inbreeding, and demonstrated little to no significant genetic differentiation throughout the population’s range. A primary challenge in these studies was the low resolution of the few genetic markers that were available for genetic analyses.

Using mitochondrial and nuclear genetic markers, two studies by Kretzmann et al.

[14, 15] tested the hypothesis that low estimates of within-island genetic variability and between-island genetic differentiation would be found in the Hawaiian monk seal, due to its small population size and strong site fidelity. The first study found that while mitochondrial haplotype distributions suggested panmixia, nuclear multilocus fingerprints were significantly differentiated between two pairs of islands located in the

Northwestern Hawaiian Islands. The second study found inconclusive evidence of population differentiation between one pair of islands based on three microsatellite loci, which were new conservation genetics markers at that time.

Schultz et al. estimated that a minimum effective population size of possibly as small as 23 individuals survived a bottleneck in the late 1800s [8], and found a genome- wide paucity of genetic variation [16]. In a subsequent study, Schultz et al. used 18

68

microsatellite loci [8, 16, 18] to evaluate genetic status based on an estimated 85% of the

HMS population (~1,897 individuals). Again, results suggested that populations across the archipelago had low levels of genetic diversity, and were panmictic, albeit with ephemeral genetic differentiation in some years [17]. In addition to establishing an important understanding of the genetic status of the HMS, this study determined that the practice of translocating individuals to increase survival rates was not likely to have detrimental genetic effects such as compromising the capacity for local adaptation or causing outbreeding depression [17,19, 20].

The feasibility of conducting genetic analyses has grown rapidly since the first translocations, and even since the most recent genetic assessment of HMS. Here, newly available polymorphic microsatellite markers [21] offer an opportunity to revisit previous analyses and conduct additional assessments using unprecedented genetic resolution.

The aims of this HMS study were to: (1) evaluate genetic diversity and population genetic differentiation with expanded genotypes using 42 polymorphic microsatellite markers,

(2) assess inter-island movements to increase understanding of dispersal behavior, and

(3) consider the effect of translocation on the population genetics of HMS in light of expanded genetic resolution.

Methods and Materials

Study system

HMS inhabit subtropical waters and sandy beaches along roughly 2,500 km of

Hawaiian Archipelago coastlines. Approximately 85% of all HMS occur as eight subpopulations on the Northwestern Hawaiian Islands (NWHI), and 15% occur in the main Hawaiian Islands (MHI, Figure 1), where they are classified as a single

69

subpopulation [22, 23]. Subpopulation sizes range from 600 to 250 seals. Overall, the population has gone through dramatic and continued declines since the 1950s. However, the MHI subpopulation survival and growth rates are relatively high and consistent with a growing population, and although numbers have been declining in the six most studied

NWHI for many years, recent census numbers in certain locations suggest that trends in this region may be stabilizing. To support population recovery, managers use translocation to mitigate skewed sex ratios, male aggression towards females, negative interactions due to habituation, and low juvenile survival [11].

Sample collection, DNA extraction, and microsatellite genotyping

The HMSRP collected all tissue specimens (under NMFS MMPA/ESA Permit

Numbers 848-1695, 10137 and 932-1905/MA-009526), and provided subsamples for this study. Samples were collected from seals on ten islands and atolls of the Hawaiian

Archipelago (, Kure Atoll, Pearl and Hermes Reef, French Frigate Shoals,

Lisianksi, Laysan, and Nihoa (within the NWHI); and Hawaii, Ohau, , and Kuai

(within the MHI, Figure 1). With the exception of one sample that was collected in 1972, all samples were collected between 1981 and 2007.

Tissue samples were obtained while attaching individual identifier tags to the rear flippers of live captured animals, or during the necropsy of dead animals, and subsequently archived at the NMFS Pacific Islands Research Center (see [14, 16, 24] for detailed methods). The DNA extraction and microsatellite genotyping protocols used were previously reported [17, 21], and the resulting genotypes were integrated into a single data set for this study. In brief, total genomic DNA was extracted using Qiagen single-well and 96-well DNeasy Blood and Tissue Kits (Qiagen, Valencia, CA, USA),

70

following standard protocols. Negative controls were used for both extractions and polymerase chain reactions (PCR) to track potential contamination. Microsatellite loci were scored with GENEMAPPER (Applied Biosystems™, Carlsbad, CA, USA), version

5.0 [21]. Genotypes consisting of 24 loci described by Mihnovets et al. [21] were integrated with genotypes consisting of 18 loci examined by Schultz et al. [17], for a total of 2409 individuals. A total of 785 individuals (391 females, 393 males, and 2 of unknown sex; Table 1) were completely genotyped at all 42 possible loci (Table 2), thus enabling the highest possible resolution for genetic analyses. A master consensus genotype file was stored in a Microsoft Access, version 7.0 (Microsoft Corporation,

Redmond, WA, USA) relational database containing additional demographic data (see

Supplementary Information for further details).

Study design

Five genotype datasets (Table 1) were generated from the master genotype table for analyses. The first dataset (henceforth referred to as “All years”), represented the entire spatial and temporal scope of this study, and consisted of genotypes for 785 individuals. Data for all years were divided into four partitions based on four to five-year time increments, as the age at first reproduction is thought to be approximately 5 years.

The first subset contained genotypes from individuals sampled prior to 1991 (N=77).

The second, third, and fourth subsets contained genotypes from individuals that were sampled between 1991 and 1996 (N=144), 1997 and 2002 (N=296), and 2003 and 2007

(N=268), respectively. All five data sets were further partitioned according to sex to examine potential differences between female- and male -mediated gene flow in individual-based population assignment analyses. For all data sets, individuals observed

71

on the same island were assigned uniform latitude and longitude coordinates recorded from the center of each island.

Several key factors influenced the rationale behind data partitioning. First, to distinguish potential influences of translocation on population genetic patterns, genetic data were parsed out for seals born prior to 1984, which was the first year translocation was used as a management tool to reduce mortality threats. However, it was not possible to explicitly test for genetic differentiation in that initial timeframe because sample sizes were spatially and temporally uneven across island populations due to variable sampling regimes over several decades (Table 1). Therefore, data from the first five years following translocation (1985-1990) were added to increase spatial and temporal coverage. The addition of these years was not expected to skew analyses since the fifteen seals moved between 1984 and 1990 were one-year-old juvenile females, with the exception of a single male pup last seen at age one, and the age at earliest known reproduction is four and five years for females in the MHI and NWHI, respectively.

Data analysis

Quality assurance was previously reported for the 18 loci characterized by Schultz et al. [17]. For the 24 loci described by Mihnovets et al. [21], the allele scoring error was manually estimated by quantifying discrepancies among two to five replicate genotypes

(533 in total) of 252 known individuals. MICROCHECKER, version 2.2.3 [25] was used to scan microsatellite genotypes for null alleles and scoring errors caused by stuttering and large allele drop-out. Furthermore, GIMLET, version 1.3.3 [26] was used to assess the differentiation power of all 42 loci by calculating the probability of identity (P(ID)) and the probability of identity between siblings (P(ID)sibs).

72

All genetic diversity statistics were analyzed using the R analytical software environment [27, 28]. The diveRsity R-package [29] was used to calculate the following conservation genetic parameters at each single locus and across all loci: private alleles

(AP), allelic richness (AR) using rarefaction based on the smallest sample in the dataset (N

= 2); observed (Ho) and expected (He) heterozygosity; unbiased expected heterozygosity

(UHe); global Hardy-Weinberg equilibrium (HWE[glb]) based on log-likelihood tests for goodness of fit using 9,999 Markov chain Monte Carlo (MCMC) steps; one-tailed tests for homozygosity/heterozygosity excess (HWE(hom) and HWE(het)); and inbreeding

(Fis) coefficients using 999 bootstrap replicates across individuals to generate 95% bias- corrected lower and upper confidence limits (Fis (low), Fis (up)). Total allele counts, polymorphic information content (PIC), and tests for null alleles were previously reported by Mihnovets et al. [21] and Schultz et al. [17].

Population- and individual-based genetic differentiation analyses were implemented to quantify genetic structure within the total population and between all island populations for all five datasets. All population-level analyses were conducted using the R analytical software environment [27, 28]. First, the PopGenReport R package [30] was used with Weir and Cockerham’s Fst [31] and its analogues (Nei and

Chesser’s gst [32], Hedrick’s Gst [33], Meirmans and Hedrick’s GGst [34], and Jost’s D

[35]). In addition, Nei’s genetic distance (Ds) between island subpopulations was visualized using principal components analysis (PCoA) as per Jombart et al. [36]. The diveRsity R package [29] was used to perform global and pairwise Fisher’s exact tests to assess sample independence (significant at p < 0.05) from genotype counts between populations using 2,000 MCMC replications to test for departures from panmixia.

73

To assess individual-level contemporary gene flow, the Bayesian clustering approach in STRUCTURE, version 2.3.4 [37] was used to determine the number of genetic clusters (K) for HMS. Ten independent runs per K value (ranging from 1 to 10) were conducted assuming correlated allele frequencies in an admixture model with an a priori designation of collection location based on island of origin (LOCPRIOR) [38]. All runs consisted of a ‘burn-in’ period of 200,000 iterations, followed by 2,000,000 MCMC iterations. The optimal number of K was determined by calculating the rate of change of a likelihood function with respect to K (delta K) described by Evanno et al. [39] and implemented in STRUCTURE HARVESTER, version 0.6.94 [40]. Using the Greedy algorithm and 1,000 permutations in CLUMPP, version 1.1 [41] the ten replicates of the

Bayesian clustering results generated by STRUCTURE were aligned at the most likely value of K.

Furthermore, inter-island movement (represented by the estimated number of

‘first-generation migrants’ at each island) was analyzed using GENECLASS, version 2

[42] by applying likelihood-based assignment tests with the Bayesian criterion of

Rannala and Mountain [43] in combination with the MCMC resampling method of

Paetkau et al. [44]. Using an alpha level of 0.01 and simulations for 10,000 individuals, the probabilities of individuals being residents of each reference population were estimated without assuming a uniform prior distribution of allele frequencies [44]. The maximum likelihood algorithm was sampled according to the ratio of L_home (the highest likelihood of the individual genotype within the population from which the individual was sampled) to the highest likelihood value of all available population samples, to include the population from which the individual was sampled (L_max) [42].

74

The diveRsity R-package [29] was then used to estimate the relative migration rates between all pairs of population samples, and to generate a weighted network plot to test for significant directional differentiation of gene flow, which may help to reveal signals of barriers to gene flow [29, 45].

Isolation-by-distance (IBD) and spatial autocorrelation patterns were also characterized for HMS using GenAlEx, version 6.5 [46, 47] to detect spatial patterns in allele frequencies and evaluate potential ecological and evolutionary factors driving genetic differentiation. IBD and spatial autocorrelation patterns were assessed for males and females separately to identify potential sex-specific differences in movement. First, simple Mantel tests were conducted by correlating codominant genotypic distances with

Euclidean (straight-line) geographic distances. Next, spatial autocorrelation analysis was conducted by estimating pairwise squared genetic and Euclidean distance matrices, which were used to generate an autocorrelation coefficient (r) for several distance classes and were visualized as correlograms. The distance classes were chosen based on the distribution of geographic distances among island populations.

Results

DNA sample summary

Discounting controls, contamination, and replicates, PCR amplification yielded results for 2404 individuals. Of these, complete coverage for all 24 loci was achieved for

1192 individuals (see [21]). In total, 429 allelic discrepancies were identified among the

17,902 alleles examined among replicate genotypes, resulting in a genotyping error rate of 2.4%. Genotypes for these 24 novel loci were integrated with 18 polymorphic loci previously characterized by Schultz et al. [17] to generate genotypes for all 2,409 samples. Between the two sets of markers, 100% coverage across all 42 loci was

75

achieved for 785 unique individuals, which were used in subsequent analyses. No evidence was found of large allele dropout for any of the 42 loci examined, and significant evidence (p < 0.001) for the presence of a null allele was found only at locus

Neo15522, where homozygote excess could be attributed to scoring errors associated with stuttering. The probability of identity estimates (p<0.01) suggested that a minimum of three of the most informative loci out of the 42 available loci are necessary to distinguish between unrelated individuals (P(ID)). The conservative upper bound,

(P(ID)sib), suggested that a minimum of seven out of the available 42 loci are required to distinguish a sibling relationship using genetic data alone (Figure 2).

Sample sizes for the “Before 91” and “91 to 96” data partitions were skewed due to uneven spatial and temporal sampling. Therefore, results for all genetic diversity, population differentiation, and inter-island movement analyses based on these data alone should be interpreted with caution due to extremely small sample size. Nevertheless, they lend some limited perspective. For example, sample sizes for French Frigate Shoals were most consistent of all islands across the four chronological data partitions, enabling the consideration of the population genetic trend at that location across all of the years of the study.

Genetic diversity

Across all data partitions, a total of 21 private alleles were identified (Table 3), with the lowest number (N = 1) occurring at Kure Atoll, and the highest (N= 11) occurring at French Frigate Shoals. Four private alleles were identified once in each of the data partitions based on time increments, and not in the “All years” partition.

Thirteen private alleles were detected in at least one partitioned time increment as well as

76

in the “All years” data partition. For example, for Kure Atoll, allele 114 at Neo7472 was identified in the “97 to 02” partition and again in the “All years” partition. Two private alleles were first identified at French Frigate Shoals in the “Before 91” partition and subsequently identified as private alleles at two new locations (Pearl and Hermes Reef and Laysan) in the “91 to 96” data partition, possibly suggesting gene flow from the southeast to the northwest (assuming no homoplasy or genotyping error). The “Before

91” and “91 to 96” data partitions may have been prone to false positive calls due to uneven sample sizes and/or uneven representation of true allelic diversity and frequencies.

Out of 42 loci, significant departures from Hardy-Weinberg equilibrium (HWE) occurred at locus Neo3204 for the Laysan population in all data partitions, at Pearl and

Hermes Reef and Lisianski in the “97 to 02” partition, and at in the “03 to 07” and

“All years” partitions. Departures from HWE also occurred at locus Neo8948 for the

Laysan population in all data partitions, at Lisianski in the “97 to 02” partition, and at

Kauai in the “03 to 07” and “All years” partitions. However, there were no uniform patterns of significant departure of HWE for any single locus across all islands (data not shown) to suggest the exclusion of any locus from analyses.

When data for all islands were examined as a single population in all time increments (Table 4), allelic richness ranged from 1.79 to 2.50. Observed heterozygosity ranged from 0.39 to 0.41, and expected heterozygosity ranged from 0.36 to 0.40.

Unbiased expected heterozygosity calculations increased slightly over time, ranging from

0.39 to 0.41. Observed heterozygosity was slightly higher than expected heterozygosity in all time partitions, but observed heterozygosity values were nearly equal to unbiased

77

expected heterozygosity values. There were no significant results in tests for departures from HWE, or for homozygous and heterozygous excess. All estimates (with a 95% confidence interval) of inbreeding were below zero, ranging from -0.10 to -0.02.

Heterozygosity, homozygosity, and allelic richness estimates were highest, and the Fis value was closest to zero, for the “97 to 02” data partition.

In comparison to overall population estimates, genetic diversity estimates for each island population across all data partitions were more variable (Tables 5A-E, Figures 3-

4). Allelic richness remained below 1.99 for all islands across all data partitions except for the “97 to 02” partition, for which estimates ranged between 2.41 and 2.50. Observed heterozygosity estimates generally became more uniform among islands over time.

Estimates of observed and expected heterozygosity, and Fis were most consistent at

Lisianski across all data partitions, and were more variable for populations with smaller sample sizes (Figure 5). Expected heterozygosity was greater than observed heterozygosity at Laysan for all data partitions except “Before 91.” At Laysan, Fis values were positive for all partitions except for “Before 91” (Figure 6). Weakly positive Fis values (< 0.07) were also found for Pearl and Hermes Reef and Laysan Island (“91 to

96”), Kure Atoll (“97 to 02”), and French Frigate Shoals (“03 to 07”). All other Fis estimates were approximately zero or weakly negative (> -0.07 for islands with sufficient sample size) across all time increments.

Among data partitions, deviation from HWE was found for the following islands and time increments: French Frigate Shoals (“Before 91”), Pearl and Hermes Reef (“91 to

96”), for all island populations tested except Midway atoll (“97 to 02”), and for Kure

78

Atoll, Midway Atoll, and Lisianksi (“03 to 07”). However, no significant evidence for directional deviation (homozygous or heterozygous excess) was found.

Genetic differentiation

Among calculations of Fst and its four analogous estimators (D, gst, Gst, GGst) at the overall population level, values and confidence intervals for Fst and Jost’s D were the most complementary and least variable across all partitions (Figure 7). The greatest degree of concordance among all indices per data partition occurred in the “97 to 02” partition. Confidence intervals (95%) and p-values (< 0.05) of Kruskal-Wallis tests (not shown) indicated significant differences for Fst, D, gst, Gst, and GGst estimates between the “91 to 96” and “97 to 02” data partitions, and a significant difference in Jost’s D estimate between the “91 to 96” and “03 to 07” data partitions. However, differences are an artifact of the uneven sampling effort in the “91 to 96” data set, rather than an indication of the influence of biological processes.

Estimates of Fst and its analogues across all data partitions (Tables 6A-E) and between island populations across all data partitions (Appendix, Tables 1 A-D) suggested mildly variable degrees of differentiation in some cases. Higher values for three out of five indices (Fst, Gst, and GGst) based on data for “All Years” indicated greater pairwise genetic differences between Molokai and Pearl and Hermes Reef, and Molokai and

Lisianski. Again, highly skewed sample sizes in the “Before 91” and “91 to 96” data sets limited the informativeness of pairwise comparisons. Perhaps most notably, estimated values for the “97 to 02” data partition were the least variable of all data partitions, showed the greatest concordance among all indices, and suggested little to no differentiation between island populations. In the “03 to 07” data partitions, estimates for

79

Fst, Gst, and GGst were most concordant for three pairwise comparisons (Molokai/Kure

Atoll, Molokai/Pearl and Hermes Reef, and Molokai/Lisianski) and thus suggestive of possible mild population differentiation. However, sample size at Molokoi was too small to allow conclusive interpretation.

Results for Fisher’s Exact tests are reported for all data partitions in Tables 7A-E.

Among the eight islands (six NWHI and two MHI) examined in the “All Years” data partition, significant differences were found among all populations in the Northwestern

Hawaiian Islands, but not among those in the main Hawaiian Islands, nor between populations in the main and Northwestern Hawaiian Islands, respectively. Again, small sample sizes in the main Hawaiian Islands limit the interpretation of findings. No significant difference was found for any pairwise comparison among the four islands examined in the “Before 1991” data set. A fifth island was excluded from this analysis due to insufficient sample size (N=1). Significant differences were found for three pairwise comparisons among the five islands examined in the “1991 to 1996” data set.

Significant differences were found among the pairwise comparisons for all six NWHI examined in the “1997 to 2002” data set. Significant differences were found for nine pairwise comparisons among the eight islands (six NWHI and two MHI) examined in the

“2003 to 2007” data set.

Results of individual assignment tests based on allele frequency distribution and applying Bayesian clustering analysis with STRUCTURE were generally suggestive of a single population, despite some variation when data were subdivided according to certain time periods. Delta K calculations for all tested data partitions ranged from K = 2 to K =

8 (Figure 8). However, in most scenarios, associated bar plots reflected nearly equal

80

distributions of alleles across different genetic clusters, suggesting low genetic structure

(K = 1). For example, regardless of K value, individuals in the “Before 91” and “91 to

96” data partitions were equally likely to be assigned to any of the proposed populations.

A slight signal of differentiation at Pearl and Hermes Reef was reflected by slightly different proportions in the “97 to 02” data partitions, but bar plots of the overall distribution of most allele frequencies across islands were suggestive of a single population. In the “03 to 07” partition, mild differences in proportions were evident for

Kure Atoll, Midway Atoll, and Pearl and Hermes Reef, while greater uniformity was evident in the remaining islands to the southeast except at Laysan.

Overall, PCoA analyses did not distinctly separate island populations for any of the five data partitions, nor for finer scale comparisons between sexes for all partitions.

Although genetic clusters for all populations overlapped in PC space, in some cases populations, ranging from Kure Atoll to Pearl and Hermes Reef to the west, tended to cluster more closely together, particularly for all individuals in the “All years” data partition (Figure 9), for females in the “91 to 96” (Figure 10B) and “03 to 07” data partitions (Figure 11B), and for males in the “03 to 07” data partition (Figure 11B).

Males from the Laysan and French Frigate Shoals populations also tended to cluster more closely over time when compared across the “91 to 96,” “97 to 02,” and “03 to 07” partitions (Figures 10B-11B). A common pattern of slight differentiation was revealed among males in the “Before 91,” “91 to 96,” and “97 to 02” data partitions, whereby genetic clusters for Pearl and Hermes Reef and French Frigate Shoals were oriented most distantly from each other in PC space relative to both PC axes.

81

Comparisons between sexes in the “03 to 07” data partition, which contained genetic data for individuals in the main Hawaiian Islands, suggested that genetic clusters of males originating from Kauai and Molokai were more proximate in PC space to Kure

Atoll and Midway Atoll, despite the greater relative geographic remoteness of the latter.

Yet, males originating from Nihoa and , which are in relatively close geographic proximity to one another, were more remotely situated in PC space. When examined relative to all other populations in the “All years” partition, individuals from Oahu and

Kauai were clustered most closely within the PC space. Results for MHI populations should be interpreted with caution due to the small sample size, which may have resulted in over-estimation of spatial patterns [48].

Assessment of IBD patterns in HMS revealed significant positive correlations between individual genetic and geographic distances (“All years,” r = 0.096, P = 0.010;

“Before 1991,” r = 0.257, P = 0.001; “91 to 96,” r = 0.116, P = 0.001; “97 to 02,” r =

0.094, P = 0.001; “03 to 07,” r = 0.106, P = 0.010; Figures 12A- E), suggesting that geographic distance was partially driving the genetic structure detected. Spatial autocorrelation analysis showed significantly positive correlations in the first distance class at 50 km (“91 to 96,” r = 0.047, P = 0.001; “97 to 02,” r = 0.069, P = 0.001; “03 to

07,” r = 0.071, P = 0.001; Figures 13 A-D), confirming the presence of nonrandom spatial structuring. Individuals detected within 50 km from each other share a higher proportion of shared alleles than individuals detected at further geographic distances. For female HMS, spatial autocorrelation was also detected for the 50 km distance class (“91 to 96,” r = 0.048, P = 0.001; “97 to 02,” r = 0.066, P = 0.001; “03 to 07,” r = 0.070,

P = 0.001; (“All years,” r = 0.063, P = 0.001; Figures 13 E-H)). For male HMS, a similar

82

spatial autocorrelation was detected for the 50 km distance class ((“91 to 96,” r = 0.054,

P = 0.001; “97 to 02,” r = 0.073, P = 0.001; “03 to 07,” r = 0.071, P = 0.001; (“All years”, r = 0.068, P = 0.001; Figure 13 I-L)).

Inter-island movement

Among the 40 individuals that were identified as first generation migrants

(probability threshold < 0.01), 16 were male and 24 were female (Table 8). The greatest number of potential migrants was assigned to Kure Atoll (N = 10), with at least one individual originating from every island except Lisianski, Kauai, and Oahu (however,

Kauai and Oahu were represented by only three samples). The smallest number of potential migrants was assigned to French Frigate Shoals (N = 4), and none were assigned to Nihoa or Molokai (the populations of these islands were represented by extremely small sample sizes). Two individuals were identified as having possibly arrived at Kure Atoll from Molokai, which are located at the greatest geographic extremes among the locations in this study. Neither of the two translocated seals included in the data were identified by GENECLASS as potential migrants.

Relative migration networks (Figure 14) show all of the relative migration rates for all data partitions, none of which were found to be significantly asymmetric (no bias in the direction of movement between islands) by pairwise comparisons within each partition (Tables 9A-E, 95% CI, 999 bootstrap iterations). Across all data partitions, relative migration rates generally increased over time, and rates between Laysan and

French Frigate shoals were consistently high, usually ranging between 0.90 and 1.00.

Notably high rates also occurred between Midway Atoll and Pearl and Hermes Reef in the “All years” partition, for Pearl and Hermes Reef and French Frigate Shoals in the “91

83

to 96” partition, for Lisianski and Laysan in the “97 to 02” partition, and for Kure and

Midway atolls in the “03 to 07” partition. Relative migration rates in the latter two partitions averaged around 0.50 for two islands located at geographic extremes (Kure

Atoll and French Frigate Shoals), which may reflect the legacy of translocating individuals from French Frigate Shoals to Kure. Results for the “Before 91” and “91 to

96” data partitions were limited by uneven sample size and are therefore uninformative.

Discussion

Genetic diversity

In this study, a dataset capturing unprecedented genetic coverage at 42 polymorphic microsatellite loci for 785 individuals was used to revisit questions regarding genetic diversity, population structure, and inter-island movement for the endangered Hawaiian monk seal. Complete genotypes were used to avoid potential biases in genotype frequency estimates caused by missing or incomplete data at any given locus, which improves accuracy when evaluating heterozygosity, inbreeding, and genetic differentiation [49,50].

This study also provided the first evaluation of private alleles for HMS. At least one private allele was observed at each of the six NWHI, and a total of 21 were identified across all data partitions. Though not directly comparable, this is similar to the closely related Mediterranean monk seal (Monachus monachus), which was characterized by two highly differentiated populations, one with 14 private alleles and the other with 18 private alleles, using a separate set of markers [51].

Despite more than twice as many markers being used than ever before, the results were generally suggestive of low genetic diversity with very mild spatial differences along an east-west gradient, particularly during years of more consistent sampling, and

84

high levels of gene flow, thereby generally reinforcing findings by previous genetic studies [8, 13–17].

In a meta-analysis of genetic variations in mammalian populations, DiBattista et al. [52] estimated that healthy, undisturbed populations have an average of 8.84 ± 0.57 alleles, while those that have been subjected to hunting or harvest have 6.89 ± 0.46.

Here, HMS estimates fell well below both of these values. Low estimates of allelic richness ranging from 1.79 to 2.50, and peaking in the “97 to 02” data partition, were even slightly lower than in previous studies (mean allelic diversity = 3.5 ± 2.62) [16]).

This is likely due to limited variation (two to three alleles) among a majority of the 24 new loci used [21], which accords with previous findings of limited variation among loci throughout the HMS genome [8]. The peak in allelic richness observed in the “97 to 02” data partition may be due to detecting the greatest number of private alleles (N = 6) in contrast to other partitions, since calculations of allelic richness using rarefaction can be highly sensitive to private alleles [53]. Alternatively, differences between data partitions may be due to variable and uneven sampling efforts at several islands throughout the study period. Nevertheless, overall estimates are suggestive of historical and long-term disturbance [3].

Neither observed nor unbiased expected heterozygosities exceeded 0.41 in any given data partition; and overall values across all years were nearly equal at 0.39 and

0.40, respectively. In contrast, the highest observed and expected heterozygosities previously described by Kretzmann [15] and Schultz [17] were 0.80 and 0.53, respectively. Heterozygosity estimates likely declined with increased sampling at more loci due to limited allelic diversity (two to three alleles) for the majority of the 24 new

85

loci [21]. In further comparisons between data partitions, heterozygosity estimates became slightly more uniform among island populations over time, which may be suggestive of enhanced gene flow facilitated by translocation.

Diversity estimators between islands within each data partition revealed significant departure from HWE in three populations in the NHWI between 1997 and

2002, but not between 2003 and 2007. Departures from HWE indicate that the population may not meet analytical assumptions of genetic equilibrium [8], which is not unusual for small populations that have experienced a bottleneck. Thus, differences in allele frequencies between populations could be due to processes other than reproductive isolation, such as non-random mating or variable relative reproductive success among sexually mature males. Further clarification requires additional research to determine paternity and male reproductive success, since HMS mating occurs underwater and cannot be directly observed.

For the four chronological data partitions examined, direct estimates of heterozygosity or homozygosity excess were not found to be significant for any island population. However, when data for all years were grouped together, all island populations except those on Midway Atoll were estimated to have either form of directional excess, and three islands were estimated to have both. This may be attributed to analytical artifacts or to underlying biological processes. Two separate algorithms were used to test hypotheses of excess based on observed and expected heterozygosities, and on the three islands where significance was found, estimates of observed and expected heterozygosities were essentially identical. Excess heterozygosity can be associated with population admixture or with loss of alleles at selectively neutral loci due

86

to population bottlenecks [54], both of which previously described for this system [8].

Non-random mating can also lead to an excess of heterozygotes and a deficiency of homozygotes.

Detection of heterozygosity excess may at first seem counter-intuitive, given the low level of allelic diversity observed in the majority of microsatellite loci in the

Hawaiian monk seal. However, cross-homozygote mating (for example, the AA of a female crossed with the BB of a male) will result in exclusively heterozygous (AB) progeny. Therefore, it is possible that indicators of heterozygosity excess arise because cross-homozygote mating occurs more frequently in HMS than would be expected at random.

Negative values of Fis can also be indicative of heterozygosity excess. Across all data partitions, spatial and temporal variation among inbreeding coefficients on each island were often negative, vacillated extremely close to zero, and never exceeded 0.06, suggesting that despite a small population size and an historical bottleneck, mating between closely related individuals is not a common occurrence in HMS. This also aligns with previous findings [16]; however, direct analyses of individual relatedness are still needed for further validation.

Genetic differentiation

Deviation from panmixia was found for all NWHI using Fisher’s exact test for the data partitions encompassing the years 1997 to 2002, 2003 to 2007, and all of the years of the study. There were notably more significant results in the “97 to 02” partition than in the “03 to 07” partition, suggesting the possibility of ephemeral genetic differentiation due to random fluctuations in gene frequencies, as has previously been reported for HMS

87

[17]. Similar patterns were also observed using allele-based and relative fixation indices of differentiation.

Weir and Cockerham’s Fst, as well as four additional analogous estimators, were examined due to their individual strengths and weaknesses relative to data characteristics.

While distanced-based Fst and Gst estimates are based on relative fixation patterns between populations as indicators for differentiation, they are prone to underestimating genetic differentiation for polymorphic loci. By scaling indices of estimated genetic diversity, gst, Gst, and GGst are adjusted for small sample size or high polymorphism, whereas Jost’s D is an allele-based indicator that discounts genetic distance between alleles when estimating differences among populations. However, Jost’s D weighs rare alleles less heavily than common alleles, making it difficult to discern the relative influences of both on differentiation [35, 55], and, like Gst, it can be highly sensitive to mutation rate [1, 56]. Relative concordance among at least three estimators indicated several significant cases of genetic differentiation within and between data partitions, particularly for pairwise comparisons between the NWHI. Concordance between these estimators elicits greater certainty when interpreting results, while greater variability between them can indicate that further consideration of underlying data conditions or analytical assumptions is needed [1, 55].

The management of HMS is based on the guidelines that “a little genetic variation is suggestive of a single stock/population” if 0.0 < Fst ≤ 0.05, of “moderate genetic variation among stocks/populations” if 0.05 < Fst ≤ 0.15, and “great genetic variation” if

FstT > 0.15 [8]. Thus, Fst estimates for the “03 to 07” and “All year” data sets suggest moderate genetic variation among NWHI populations and between the NWHI and MHI

88

populations, however the MHI results should be omitted due to extremely small sample sizes. Cases of weak but significant population-level genetic differentiation may be due to the presence of closely related individuals (e.g., siblings) in the datasets used.

The findings of this study contrast with previous studies in several ways. Early analyses by Kretzmann et al. [14-15] found evidence of differentiation according to within-island DNA fingerprint similarity indices ranging from 0.49 to 0.73 and Fst =

0.038, and also found significant evidence of population subdivision for pairwise comparisons of Lynch’s F’st [57] between Laysan and Lisianski (F’st = 0.20), and Pearl and Hermes and Kure Atoll (F’st = 0.13). And a subsequent study [15] of one microsatellite locus in 108 individuals and two microsatellite loci in 46 individuals found inconclusive evidence of population differentiation between Kure Atoll and the French

Frigate Shoals (Rst = 0.127) [58, 59]. Although they found evidence of mild and ephemeral differentiation in estimates for individual cohort years, Schultz et al. [17] did not find significant evidence of spatial or temporal differentiation based on Fisher’s exact tests, Fst < 0.05, nor on estimates of Weir and Cockram’s θ = -0.03 [31] based on 18 polymorphic microsatellite loci in 1,897 individuals, and thus determined that HMS should be managed as a single population.

Bayesian clustering analysis for the data partitions encompassing all of the years in this study generally agreed with the finding of Schultz et al. [17] of a single population. A few mild differences in allele proportions arose among results for data partitions representing different time frames. Most notably, in assignment tests for data collected between 2003 and 2007, individuals originating from Kure Atoll, Midway

Atoll, and Pearl and Hermes Reef were clustered most closely together in PC space,

89

while plots of individuals from Laysan Island spanned the greatest range in PC space.

Overall, there does not appear to be a strong case for sex-biased influences on proportions of allele frequencies, which may be facilitated in part by management interventions to offset skewed sex ratios using translocation [60].

PCoA findings complement a subset of results from a study by Schmelzer et al.

[61], which used cluster analysis to identify three distinct biogeographical regions (Figure

15) that Hawaiian monk seals inhabit within the Hawaiian Archipelago. Determinations were based on a non-uniform spatial distribution of environmental variables (sea surface temperature (SST), chlorophyll-a, mixed layer depth, and primary productivity).

Differentiation of the first three variables was distinct among Kure, Midway, Pearl and

Hermes Reef (collectively designated as Region 3), Lisianski and Laysan (Region 1),

French Frigate Shoals, Nihoa, and Necker (Region 2).

The parallels between environmental data and genetic data for islands in Region 3 suggest that biophysical conditions may influence movement and habitat selection, which in turn may be an important factor in the genetic connectivity of Hawaiian monk seals.

However, there was more variability among remaining islands. For example, Schmelzer et al.’s environmental clustering of Lisianski with Laysan agreed with genetic clustering only for males in the “All years” partition of this study, but genetic assignments for all individuals in the “All years” partition indicated more similar allele frequency distributions between Laysan and French Frigate Shoals than between Laysan and

Lisianski.

Assessment of isolation-by-distance and spatial autocorrelation analysis revealed that geographic distance was partially driving the genetic structure detected. Results

90

indicated non-random spatial structuring at close geographic distances, suggesting that males and females were more closely related at distances of up to 50 km than were individuals taken at random from the population. In contrast to other marine mammals, for which dispersal is male-biased [62-63], no differences in the degree of sex-biased dispersal were detected. A proximate cause for this finding may be the translocation of approximately 110 seals (the majority of which were female) during the study period, which increased inter-island gene flow. Dispersal behavior between the sexes can also vary depending on various other factors, including habitat characteristics and population demographics [64-66]. Further clarification of relationships between biophysical habitat characteristics and genetic connectivity will require directly integrating environmental and genetic data in a seascape genetics-based analytical framework [67–71].

Inter-island movement

Translocation is an important factor in HMS dispersal and gene flow, and is at least partially reflected in the results of first-generation migrant analyses. Seventy-five weaned or immature female seals were translocated from French Frigate Shoals to Kure

Atoll between 1984 and 1995, and they later gave birth to 68 pups ([72], NMFS unpublished data). Population survey data have since indicated that at least 24 females observed at Kure Atoll may have lineage origins from French Frigate Shoals ([72],

NMFS unpublished data). Indeed, several individuals with this ancestral profile were identified as immigrants in this study. Yet the only translocated individual directly accounted for in this study (due to the availability of complete genotypes) was not detected.

91

Factors that could be influencing detection in this analysis are: (1) level of genetic diversity, as low genetic diversity in the system limits the power to detect migrants; (2) high levels of gene flow, as a relatively even distribution of alleles across all islands makes it difficult to track individuals back to source populations; (3) geographic coverage of the study system, as a failure to include appropriate populations in the analysis could result in a false positive detection that an individual is descended from immigrants, or is an immigrant; and (4) random chance (e.g., when α values are too high) of false detection due to a large sample size requiring a larger number of pairwise comparisons [43].

Further analysis that includes full genotypes for all translocated individuals and even sampling across all populations of interest would allow for more rigorous assessment of the utility and power of first generation migrant analysis for an endangered population with low genetic diversity, little to no population differentiation, and systematic assisted migration via translocation.

Results from directional migration analyses yielded two relatively consistent patterns across data sets regardless of sample size. First, the highest levels of gene flow and the greatest genetic similarity were found for populations from French Frigate Shoals and Laysan. Second, greater genetic similarity was found for populations from Lisianski,

Pearl and Hermes Reef, and Midway Atoll, albeit with variable, and relatively lower, rates of gene flow. The algorithm used in this analysis is still considered to be in the early stages of development since its performance has not been fully assessed across a comprehensive suite of evolutionary scenarios that may hold true for different study systems [29]. Furthermore, rare alleles are interpreted as evidence of no migration; therefore, incomplete detection of rare alleles in a population could lead to an

92

overestimation of migration. Nevertheless, preliminary results indicate similar patterns of genetic differentiation between populations to those of assignment and clustering tests, suggesting that the diveRsity R-package [29] could well be a reliable and useful tool for further investigation of HMS migration and dispersal.

Conclusion

Using the largest set of genetic loci ever examined for the Hawaiian monk seal, the results of this genetic assessment generally corroborate previous findings of low genetic diversity, and weak population structure as well as panmixia, depending on the metric and analysis of choice and spatial and temporal coverage of data. This study also contributes a new, preliminary perspective on the use of genotype data to infer migration patterns for a highly endangered species that has undergone at least one demographic bottleneck.

The integration of genetic investigations into conservation effort enables consideration of long-term species persistence, and complements ecologically-based approaches targeting short-term extinction factors of concern. Genetic variation provides the raw material for adaptive capacity, and plays an important role in individual and population fitness. Movement (e.g., the migration or assisted migration) of individuals throughout the Hawaiian Archipelago promotes gene flow, and is therefore one important mechanism for maintaining genetic variation. Any additional management built around the current understanding of genetic diversity and differentiation for the Hawaiian monk seal is unlikely to be a ‘wrong approach’.

Acknowledgements

93

We thank all who have been involved with the Hawaiian monk seal population surveys and sample collections, the Hawaiian monk seal team at the Honolulu Laboratory of the

Pacific Islands Science Center, and Jason Baker, Albert Harting, and Angie Kaufman for their thoughtful comments and generosity with data queries. We are grateful to Ramon

Perez, Emily Lipstein, Karoline Lake, Becky Hersch, and Stephen Gaughran for their invaluable contributions in the laboratory at the Sackler Institute for Comparative

Genomics.

94

References

1. Frankham R. Genetics and conservation biology. Comptes Rendus Biologies. 2003; 326, Supplement 1:22–9.

2. Shaffer ML. Minimum population sizes for species conservation. BioScience. 1981; 31: 131–4.

3. Lande R. Genetics and demography in biological conservation. Science. 1988 ; 241: 1455–60.

4. Allendorf FW, Leary RF. Heterozygosity and fitness in natural populations of animals. Conservation Biology: the Science of Scarcity and Diversity. Sunderland, Massachusetts: Sinauer Associates; 1986. p. 57–76. Available from: http://ci.nii.ac.jp/naid/10020129642/

5. Chen Y-S, Su Y-C, Pan W. Effect of spatial constraints on Hardy-Weinberg equilibrium. Scientific Reports. 2016; 6: 19297.

6. Spielman D, Brook BW, Frankham R. Most species are not driven to extinction before genetic factors impact them. Proceedings of the National Academy of Sciences of the United States of America. 2004; 101: 15261–4.

7. Hedrick PW. Conservation genetics: where are we now? Trends in Ecology & Evolution. 2001; 16: 629–36.

8. Schultz JK, Marshall AJ, Pfunder M. Genome-wide loss of diversity in the critically endangered Hawaiian monk seal. Diversity. 2010; 2: 863–80.

9. IUCN. The IUCN Red List of Threatened Species. Version 2015-3. http://www.iucnredlist.org/details/13654/0 [Internet]. 2015 [cited 2015 Nov 2]. Available from: http://www.iucnredlist.org

10. Harting AL, Johanos TC, Littnan CL. Benefits derived from opportunistic survival- enhancing interventions for the Hawaiian monk seal: the silver BB paradigm. Endang Species Res. 2014; 25: 89–96.

11. Baker JD, Becker BL, Wurth TA, Johanos TC, Littnan CL, Henderson JR. Translocation as a tool for conservation of the Hawaiian monk seal. Biological Conservation. 2011; 144: 2692–701.

12. Griffith B, Scott JM, Carpenter JW, Reed C. Translocation as a species conservation tool: status and strategy. Science. 1989; 245: 477–80.

13. Gemmell NJ, Goodman SJ, Allen PJ, Reed JZ. Interspecific microsatellite markers for the study of pinniped populations. Molecular Ecology. 1997; 6: 661.

95

14. Kretzmann MB, Gilmartin WG, Meyer A, Zegers GP, Fain SR, Taylor BF, et al. Low genetic variability in the Hawaiian monk seal. Conservation Biology. 1997; 11: 482–90.

15. Kretzmann MB, Gemmell NJ, Meyer A. Microsatellite analysis of population structure in the endangered Hawaiian monk seal. Conservation Biology. 2001; 15: 457– 66.

16. Schultz JK, Baker JD, Toonen RJ, Bowen BW. Extremely low genetic diversity in the endangered Hawaiian monk seal (Monachus schauinslandi). Journal of Heredity. 2009; 100: 25–33.

17. Schultz JK, Baker JD, Toonen RJ, Harting AL, Bowen BW. Range-wide genetic connectivity of the Hawaiian monk seal and implications for translocation. Conservation Biology. 2011; 25: 124–32.

18. Gemmell NJ, Goodman SJ, Allen PJ, Reed JZ. Interspecific microsatellite markers for the study of pinniped populations. Molecular Ecology. 1997; 6: 661.

19. Goossens B, Funk SM, Vidal C, Latour S, Jamart A, Ancrenaz M, et al. Measuring genetic diversity in translocation programmes: principles and application to a chimpanzee release project. Animal Conservation. 2002; 5: 225–36.

20. Storfer A. Gene flow and endangered species translocations: a topic revisited. Biological Conservation. 1999; 87: 173–80.

21. Mihnovets AN, Schultz JK, Wultsch C, Littnan CL, Amato G. A novel microsatellite multiplex assay for the endangered Hawaiian monk seal (Neomonachus schauinslandi). Conservation Genet Resour. 2016; 1–5.

22. National Marine Fisheries Service. Recovery plan for the Hawaiian monk seal (Monachus schauinslandi). Second revision. National Marine Fisheries Services, Silver Spring, MD. 165 pp.; 2007.

23. Baker JD, Harting AL, Wurth TA, Johanos TC. Dramatic shifts in Hawaiian monk seal distribution predicted from divergent regional trends. Marine Mammal Science. 2011; 27: 78–93.

24. Henderson JR, Johanos TC. Effects of tagging on weaned Hawaiian monk seal pups. Wildlife Society Bulletin. 1988; 16: 312–317.

25. Van Oosterhout C, Hutchinson WF, Wills DPM, Shipley P. Micro‐checker: software for identifying and correcting genotyping errors in microsatellite data. Molecular Ecology Notes. 2004; 4: 535–8.

96

26. Valière N. GIMLET: a computer program for analysing genetic individual identification data. Molecular Ecology Notes. 2002; 2: 377–9.

27. Ihaka R, Gentleman R. R: a language for data analysis and graphics. Journal of Computational and Graphical Statistics. 1996; 5: 299–314.

28. Venables B, Smith D, Gentleman R, Ihaka R. Notes on R: a programming environment for data analysis and graphics [Internet]. Auckland: University of Auckland; 1997 [cited 2016 Apr 7]. Available from: https://ftp.nice.ch/pub/next/science/mathematics/_R.0.49.I.b/R-0.49/_R.pdf

29. Keenan K, McGinnity P, Cross TF, Crozier WW, Prodöhl PA. diveRsity: An R package for the estimation and exploration of population genetics parameters and their associated errors. Methods Ecol Evol. 2013 ;4: 782–8.

30. Adamack AT, Gruber B. PopGenReport: simplifying basic population genetic analyses in R. Methods in Ecology and Evolution. 2014; n/a–n/a.

31. Weir BS, Cockerham CC. Estimating F-statistics for the analysis of population structure. Evolution. 1984; 38: 1358–70.

32. Nei M, Chesser RK. Estimation of fixation indices and gene diversities. Annals of Human Genetics. 1983; 47: 253–9.

33. Hedrick PW. A standardized genetic differentiation measure. Evolution. 2005; 59: 1633–8.

34. Meirmans PG, Hedrick PW. Assessing population structure: FST and related measures. Molecular Ecology Resources. 2011; 11: 5–18.

35. Jost L. GST and its relatives do not measure differentiation. Molecular Ecology. 2008; 17: 4015–26.

36. Jombart T, Devillard S, Balloux F. Discriminant analysis of principal components: a new method for the analysis of genetically structured populations. BMC Genetics. 2010 ;11: 94.

37. Pritchard JK, Stephens M, Donnelly P. Inference of population structure using multilocus genotype data. Genetics. 2000; 155: 945–59.

38. Hubisz MJ, Falush D, Stephens M, Pritchard JK. Inferring weak population structure with the assistance of sample group information. Molecular Ecology Resources. 2009; 9: 1322–32.

39. Evanno G, Regnaut S, Goudet J. Detecting the number of clusters of individuals using the software STRUCTURE: a simulation study. Molecular Ecology. 2005; 14: 2611–20.

97

40. Earl DA, vonHoldt BM. STRUCTURE HARVESTER: a website and program for visualizing STRUCTURE output and implementing the Evanno method. Conservation Genet Resour. 2011; 4: 359–61.

41. Jakobsson M, Rosenberg NA. CLUMPP: a cluster matching and permutation program for dealing with label switching and multimodality in analysis of population structure. Bioinformatics. 2007; 23: 1801–6.

42. Piry S, Alapetite A, Cornuet J-M, Paetkau D, Baudouin L, Estoup A. GENECLASS2: a software for genetic assignment and first-generation migrant detection. J Hered. 2004; 95: 536–9.

43. Rannala B, Mountain JL. Detecting immigration by using multilocus genotypes. PNAS. 1997; 94: 9197–201.

44. Paetkau D, Slade R, Burden M, Estoup A. Genetic assignment methods for the direct, real-time estimation of migration rate: a simulation-based exploration of accuracy and power. Molecular Ecology. 2004; 13: 55–65.

45. Sundqvist L, Zackrisson M, Kleinhans D. Directional genetic differentiation and asymmetric migration. arXiv:1304.0118 [q-bio] [Internet]. 2013 [cited 2016 Jan 3]; Available from: http://arxiv.org/abs/1304.0118

46. Peakall R, Smouse PE. GENALEX 6: genetic analysis in Excel. Population genetic software for teaching and research. Molecular Ecology Notes. 2006; 6: 288–95.

47. Peakall R, Smouse PE. GENALEX 6.5: genetic analysis in Excel. Population genetic software for teaching and research—an update. Bioinformatics. 2012; 28: 2537–9.

48. Sinclair EA, Hobbs RJ. Sample size effects on estimates of population genetic structure: implications for ecological eestoration. Restoration Ecology. 2009; 17: 837–44.

49. Pemberton JM, Slate J, Bancroft DR, Barrett JA. Nonamplifying alleles at microsatellite loci: a caution for parentage and population studies. Molecular Ecology. 1995; 4: 249–52.

50. Bonin A, Bellemain E, Bronken Eidesen P, Pompanon F, Brochmann C, Taberlet P. How to track and assess genotyping errors in population genetics studies. Molecular Ecology. 2004; 13: 3261–73.

51. Pastor T, Garza JC, Aguilar A, Tounta E, Androukaki E. Genetic diversity and differentiation between the two remaining populations of the critically endangered Mediterranean monk seal. Animal Conservation. 2007; 10: 461–9.

98

52. DiBattista JD. Patterns of genetic variation in anthropogenically impacted populations. Conserv Genet. 2007; 9: 141–56.

53. Kalinowski ST. Counting alleles with rarefaction: private alleles and hierarchical sampling designs. Conservation Genetics. 2004; 5: 539–43.

54. Roderick GK, Navajas M. Genes in new environments: genetics and evolution in biological control. Nat Rev Genet. 2003; 4: 889–99.

55. Bird CE, Karl SA, Smouse PE, Toonen RJ. Detecting and measuring genetic differentiation. Phylogeography and population genetics in crustacea. CRC Press; 2011. p. 31–55.

56. Leng L, Zhang D-X. Measuring population differentiation using GST or D? A simulation study with microsatellite DNA markers under a finite island model and nonequilibrium conditions. Molecular Ecology. 2011; 20: 2494–509.

57. Lynch M. Analysis of population genetic structure by DNA fingerprinting. EXS. 1991; 58: 113–26.

58. Slatkin M. A measure of population subdivision based on microsatellite allele frequencies. Genetics. 1995; 139: 457–62.

59. Goodman SJ. RST Calc: a collection of computer programs for calculating estimates of genetic differentiation from microsatellite data and determining their significance. Molecular Ecology. 1997; 6: 881–5.

60. Johanos TC, Becker BL, Baker JD, Ragen TJ, Gilmartin WG, Gerrodette T. Impacts of sex ratio reduction on male aggression in the Critically Endangered Hawaiian monk seal Monachus schauinslandi. Endang Species Res. 2010; 11: 123–32.

61. Schmelzer I. Seals and seascapes: covariation in Hawaiian monk seal subpopulations and the oceanic landscape of the Hawaiian Archipelago. Journal of Biogeography. 2000; 27: 901–14.

62. Slade RW, Moritz C, Hoelzel AR, Burton HR. Molecular population genetics of the southern elephant seal Mirounga leonina. Genetics. 1998; 149: 1945–57.

63. Oosthuizen WC, Bester MN, Tosh CA, Guinet C, Besson D, Nico de Bruyn P j. Dispersal and dispersion of southern elephant seals in the Kerguelen province, Southern Ocean. Antarctic Science. 2011; 23: 567–577.

64. Greenwood PJ. Mating systems, philopatry and dispersal in birds and mammals. Animal Behaviour. 1980; 28: 1140–62.

99

65. Sugg DW, Chesser RK, Stephen Dobson F, Hoogland JL. Population genetics meets behavioral ecology. Trends in Ecology & Evolution. 1996; 11: 338–42.

66. Matthysen E. Density-dependent dispersal in birds and mammals. Ecography. 2005; 28: 403–16.

67. Selkoe KA, Henzler CM, Gaines SD. Seascape genetics and the spatial ecology of marine populations. Fish and Fisheries. 2008; 9: 363–77.

68. Galindo HM, Olson DB, Palumbi SR. Seascape genetics: a coupled oceanographic- genetic model predicts population structure of Caribbean corals. Current Biology. 2006; 16: 1622–6.

69. Treml EA, Halpin PN, Urban DL, Pratson LF. Modeling population connectivity by ocean currents, a graph-theoretic approach for . Landscape Ecol. 2007; 23: 19–36.

70. Selkoe KA, Watson JR, White C, Horin TB, Iacchei M, Mitarai S, et al. Taking the chaos out of genetic patchiness: seascape genetics reveals ecological and oceanographic drivers of genetic patterns in three temperate reef species. Molecular Ecology. 2010; 19: 3708–3726.

71. Kershaw F, Rosenbaum HC. Ten years lost at sea: response to Manel and Holderegger. Trends in Ecology & Evolution. 2014; 29: 69–70.

72. Schultz JK. Conservation genetics of the Hawaiian monk seal: A dissertation submitted to the graduate division of the University of Hawai’i in partial fulfilllment of the requirements for the degree of Doctor of Philosophy [Internet]. University of Hawai’i; 2009 [cited 2011 Feb 20]. Available from: http://proquest.umi.com.ezproxy.cul.columbia.edu/pqdlink?Ver=1&Exp=02-19- 2016&FMT=7&DID=1904236691&RQT=309&attempt=1&cfc=1

100

Figure 4-1. Map of the Northwestern Hawaiian Islands (NWHI) and main Hawaiian Islands (MHI). Image credit: www.kumukahi.org.

101

Figure 4-2. Comparison of the probability, with threshold P(ID) < 0.01, of distinguishing between unrelated individuals, (P(ID)), and between siblings (P(ID)sib) in a given population.

The conservative upper bound, (P(ID)sib), suggests that a minimum of seven out of the available 42 loci are required to distinguish a sibling relationship using genetic data alone.

102

Figure 4-3. Calculations of mean expected heterozygosity plotted against mean observed heterozygosity for data partitions. (Dotted line represents equal observed and expected heterozygosity values.)

Figure 4-4. P-values of HWE exact tests for homozygosity and heterozygosity excess for data partitions. (Red dashed lines indicate 0.05 significance thresholds.)

103

Figure 4-5. Inbreeding coefficient (Fis) plotted relative to sample sizes of data partitions examined

Figure 4-6. Observed (Ho) and expected (He) heterozygosity and inbreeding coefficient (Fis) plotted for each time increment and for combined time increments (all years).

104

Figure 4-7. Unbiased global estimates for Fst and four analogues, with associated 95% confidence intervals calculated for each data partition using the diveRsity R package [29]. Estimators are as follows: Nei & Chesser’s gst [32], Hedrick’s Gst [33], Meirmans and Hedrick’s GGst [34], Jost’s D [35], and Weir and Cockerham’s Fst [31]. Comparisons between respective confidence intervals for each estimator suggest that there is a significant difference between gst, Gst and GGst values between the “91 to 96” and “07 to 02” data partitions. Negative values can be a result of small sample size. The greatest degree of concordance among all indices occurs in the “97 to 02” data partition.

105

Figure 4-8. Results of genetic individual assignment tests for five data partitions.

“All years” represents the entire dataset comprised of 785 individuals. Line plots in the left column indicate DeltaK (ΔK) as the likelihood (y-axis) of individual genotypes clustering into K groups (x-axis), indicating the rate of change in lnP(X/K). Bar plots show the clustering of individuals by STRUCTURE [37], and the width of plots for each island population is relative to respective sample sizes. Plots are arranged relative to geographic location, where the origin of the graph represents the most northwestern Hawaiian island and the far right of the graph represents the most southeastern Hawaiian island. Due to small sample sizes, Nih, Kau, Oah, and Mol are labeled as “MHI” (main Hawaiian Islands), but were independently designated as LOCPRIORS for analysis.

106

Figure 4-9. Principal component analysis (PCoA) of the “All years” data partition based on the genetic diversity of the 42 microsatellite loci used in this study. Analyses were based on data for all individuals, females only, and males only using the PopGenReport R-package [30]. Color-coded dots with an associated numeric code indicate the geographic origin of samples.

All Years

All

Female

Male

107

Figure 4-10. PCoA plots for data partitions of samples collected (A) before 1991 and (B) between 1991 and 1996. Plots include combined data for both males and females (all), as well as sub-divided data for comparisons between sexes (female and male).

(A) (B)

108

Figure 4-11. PCoA plots for data partitions of samples collected between (A) 1997 and 2002 and (B) between 2003 and 2007. Plots include combined data for both males and females (all), as well as sub-divided data for comparisons between sexes (female and male).

(A) (B)

109

Figure 4-12. Isolation-by-distance (IBD) in Hawaiian monk seals (HMS). IBD patterns were assessed for HMS samples collected between (A) 1972-2007 (all sampling years), (B) before 1991, (C) 1991-1996, (D) 1997-2002, and (E) 2003-2007.

(A)

(B)

(C)

110

(D)

(E)

111

112

Figure 4-13. Spatial autocorrelation analysis in Hawaiian monk seals (HMS). Spatial autocorrelograms (A-D: all seals; E-H: females; I-L: males) showed genetic correlation coefficients (r) as a function of geographic distance across several spatial distance classes. Dashed red lines represent upper (U) and lower (L) bounds of null distribution based on 10,000 random permutations. Error bars represent 95% confidence intervals for r based on 1,000 bootstraps.

(A)

(B)

113

(C)

(D)

(E)

114

(F)

(G)

(H)

115

(I)

(J)

(K)

116

(L)

117

Figure 4-14. Relative migration networks generated based on estimates of Nm for all data partitions, assuming an infinite allele model, equilibrium of gene diversities, and an island migration model, using the diveRsity R package [29]. Line darkness indicates gene flow level (higher levels are represented by darker lines), and relative proximity of populations indicates the genetic similarity of populations. Populations that are in close proximity to each other and are connected by lighter lines are geographically, but not genetically, close. Significant differences between relative migration rates suggestive of directionality were not found in any of the data partitions.

118

Figure 4-15. Classification by Schmelzer et al. [61] of Hawaiian monk seal breeding areas into biogeographical regions according to differential environmental variables. In geographical order, Kure, Midway, and Pearl and Hermes Reef comprise Region 3, Lisianski and Laysan comprise Region 1, and French Frigate Shoals, Nihoa, and Necker comprise Region 2.

Region 3

Region 1

Region 2

119

Table 4-1. Overall sample size (All) and partitioned according to sex ((U)nknown, (F)emale, (M)ale), location where individuals where first observed, and year when first identified. Data were partitioned according to years prior to translocation plus subsequent years in which translocated individuals were not expected to have contributed to the gene pool (“Before 91”), and into four- or five-year time increments thereafter.

“Before s91” “91 to 96” “97 to 02” “03 to 07” “All years”

Island All F M All F M All F M All U F M All U F M

Kur 3 1 2 22 8 14 48 20 28 48 21 27 121 50 71 Mid 2 2 25 14 11 29 19 10 56 35 21 PHR 4 1 3 40 19 21 51 27 24 27 1 14 12 122 1 61 60

120 Lis 1 1 13 8 5 37 19 18 61 30 31 112 57 55

Lay 8 1 7 7 5 2 78 40 38 56 36 20 149 82 67 FFS 61 37 24 60 25 35 57 26 31 39 15 24 217 103 114 Nih 1 1 1 1 Kau 3 1 2 3 1 2 Oah 1 1 1 1 Mol 3 2 1 3 2 1 All 77 40 37 144 67 77 296 146 150 268 1 138 129 785 1 391 393

120 blah

Table 4-2. Genotypes integrated in this study were developed using 42 polymorphic microsatellite loci with the following Genbank accession numbers, repeat motifs, forward and reverse primer sequences, allele sizes, observed (Ho) and expected (He) heterozygosities, and number of alleles (A). Genbank Repeat Locus Reference Primers Size Ho He A Accession # motif F: TCACTCACACCAGGAGAAGG Neosch00871 Mihnovets et al. 2016 KU220201 (GT)14 140-144 0.17 0.18 3 R: CCACGCTAGACATGGTCTTTG F: CCACTTGGAAGCAAAAGTAAAGC Neosch01807 Mihnovets et al. 2016 KU220202 (AC)17 95-99 0.72 0.50 3 R: TAGGAAAGCCTTCTCTGGGC F: ACCAGGAGTAGTCCAAACAGG Neosch03204 Mihnovets et al. 2016 KU220204 (AC)13 187-189 0.37 0.39 2 R: CACCCGTGTTACAGAGGAAG F: GCTTTGGCTCTCATTCTCAGC Neosch03980 Mihnovets et al. 2016 KU220205 (GT)15 94-100 0.17 0.18 3 R: AACATCACTGCCCACTCTCC F: AGATGTGCATGATGTTACCCAAG Neosch04022 Mihnovets et al. 2016 KU220206 (CCAT)11 141-145 0.35 0.38 2 R: GTGCCACCAGTTAAAGAGGG F: TGGGTTGACTACCATCGCTC Neosch04553 Mihnovets et al. 2016 KU220207 (TG)21 170-172 0.47 0.47 2 121 R: AGCAAGGGACGTAGTGACAG F: CGCATTCACGCACTTTTCAC Neosch04584 Mihnovets et al. 2016 KU220208 (CA)18 219-223 0.21 0.21 3 R: CGATGGTAGAGGAGGTTGGG F: TCATTGGTATTCATGTTCCCCC Neosch06554 Mihnovets et al. 2016 KU220209 (AC)14 176-178 0.47 0.50 2 R: AGCCCAGGCTTGTTGAAATG F: AGAATGAAAGGCAGACATGCG Neosch07472 Mihnovets et al. 2016 KU220211 (CA)17 110-116 0.48 0.50 3 R: GTCCCTCGAACAGACTGGAG F: AGCCTTGTGATGTCAATCTGC Neosch08549 Mihnovets et al. 2016 KU220212 (AC)13 173-179 0.45 0.44 3 R: TGCGGACAATTCATCCAAATATTAAC F: GTATGTTGTTGCAAACGGCAG Neosch08948 Mihnovets et al. 2016 KU220213 (TG)15 123-127 0.09 0.10 3 R: AGTGCCTGTTGTGATGGATG F: GCCCTTCATCACCTGCTTTG Neosch09043 Mihnovets et al. 2016 KU220214 (CA)13 157-159 0.47 0.50 2 R: GAGCTGTGGCGAAAGACATC F: GAGGATGGCTGAACCAGGAG Neosch10475 Mihnovets et al. 2016 KU220215 (GT)18 82-84 0.16 0.15 2 R: TGATGCCACAGTCTCTCCAC F: AAATCTTGTGGATTCTTCAGCC Neosch15057 Mihnovets et al. 2016 KU220223 (AC)16 R: AACCAATGGCTGGGTCTGTG 79-81 0.31 0.31 2

blah121

Genbank Repeat Locus Reference Primers Size Ho He A Accession # motif F: GATACCAGTGGCCCAAACTC Neosch10588 Mihnovets et al. 2016 KU220216 (CA)16 105-111 0.26 0.26 3 R: TCAAGCATTCAACCAAGTGC F: GTGCGTGCTCTCTATTTTCCC Neosch10909 Mihnovets et al. 2016 KU220217 (CA)16 113-117 0.29 0.28 2 R: TGTGCATTCCTGTAGCAGAG F: ACTGGCTTTCCTCACTCCTG Neosch11181 Mihnovets et al. 2016 KU220218 (AC)15 222-228 0.39 0.38 2 R: CGTGAGAATCAACCCTCTGC F: CTGACACCTTCTCTACCGGC Neosch12980 Mihnovets et al. 2016 KU220219 (AC)20 218-224 0.42 0.42 4 R: CAGTGCCCACGCATGTTTC F: AGGAGGGATTCCAGCAAGAG Neosch14403 Mihnovets et al. 2016 KU220220 (CA)12 297-301 0.17 0.17 3 R: ACAGCAATGGTGATGGGATG F: TGAATAGCACTTTCTCTCCACTTC Neosch15522 Mihnovets et al. 2016 KU220224 (AC)22 112-118 0.32 0.66 4 R: GACCATGAGGGAGGAGCTG F: TTCTCCGACCTTGCCTTACC Neosch15716 Mihnovets et al. 2016 KU220225 (AC)13 120-132 0.36 0.35 4 R: ACACTTTCACTGCCCTTTGC Neosch16869 F: GGGACAATTTCTCTCTCTCCC 170 -

122 Mihnovets et al. 2016 KU220227 (CA)20 0.40 0.39 2 R: ATGTTCTTTAGTCTCACTTCACG 172

F: TCAAATGGTGGAGGTGAGGG Neosch17310 Mihnovets et al. 2016 KU220228 (GT)14 124-126 0.47 0.49 2 R: ACTTTTGGTATGTGCTCTGTTCTC F: AGCCACATGCAGAAGGTTTG Neosch18265 Mihnovets et al. 2016 KU220229 (TG)16 104-106 0.42 0.42 2 R: TCTCTAATCATGTGAGCCAATTCC F: ATTTTAATTATGGGTTACTTTGAACC Msc01 Schultz et al. 2010 GU206362 (AC)18 161-167 0.47 0.49 3 R: TCACCATTTAATGCATATGAGC F: TGGTCTTTCTTAAGGCCAAG Msc03 Schultz et al. 2010 GU206363 (TG)21 126-136 0.70 0.71 5 R: ATATGGAAGCAGCCCAAGTG F: CTTTAGTTTCCGGTGTTCAGTG Msc04 Schultz et al. 2010 GU206364 (TATC)14 159-167 0.52 0.52 4 R: CTCAGGGTTGTGAGTTCAAGC F: TGGTCTCAAGTTGGAGGATTG 198- Msc05 Schultz et al. 2010 GU206365 (GATA)11 0.27 0.27 3 R: AAGCCATTGAGTGTGATGGAC 206 F: GCCTGATTTGCCTCTTCTTC Msc09 Schultz et al. 2010 GU206366 (CTAT)13 R: GCGTCAGAAAGACACAGGAG 208-216 0.28 0.29 3

blah122

Genbank Repeat Locus Reference Primers Size Ho He A Accession # motif F: CCTCCATCGCTACCATCTTC Msc10 Schultz et al. 2010 GU206367 (TATC)13 141-149 0.47 0.48 3 R: TGAACGCAAGTGGATGAGTC F: CACCTTTGGCTTCCAGTGTC Msc13 Schultz et al. 2010 GU206368 (CA)20 194-202 0.40 0.40 3 R: ATTCGGTGGTGGCTTTTATG F: GCAAGGAGAAGCTACAGAAGG 115- Msc17 Schultz et al. 2010 GU206370 (GT)24 0.50 0.51 3 R: CCCTGCTAACTGGTTTCTGC 119 F: GGCTATTGGCCAACTGGTAG Msc19 Schultz et al. 2010 GU206371 (CTAT)11 118-138 0.24 0.24 4 R: TTGGCCTGCTCCAATAAGAC (GATA)2 (GAT) F: GCTTCTCTGTTTCTATCTCAAATAAAT Msc23 Schultz et al. 2010 GU206372 (GATA)2 160-168 0.48 0.50 3 R: CTCCTTCCTGGCTGCTTATG (GAT) (GATA)15 Ms9 F: CCAAAGCCTATTTCTTTCAATCC

123 Schultz et al. 2009 EU913766 (GAAA)18 297-317 0.67 0.68 6 R: AGCAGAGGCCCTAAGACAGG

(CCTT)6 F: CTGAATTCATGCTGTATCTTGG Ms15 Schultz et al. 2009 EU913767 CCCT 203-315 0.54 0.56 3 R: GTGCTTGGGACATGATGG (CCTT)6 (GAAA)9 F: CGCTTAGTGTGGAGTCACTTAGG Ms23 Schultz et al. 2009 EU913768 GGAA 340-370 0.76 0.78 9 R: GTGAGATGAATGCCCTTTGG (GAAA)8 F: GACTGGTAATTTACGCCCTACC Ms265 Schultz et al. 2009 EU913769 (GT)13 158, 162 0.50 0.49 2 R: AAGTGTTGGGTTGAAAATTGG F: ATCAGCTATCAGGGGTAGGG Ms504 Schultz et al. 2009 EU913763 (AAG)24 308, 326 0.28 0.29 2 R: GTCATTCCCTAGTGGTAAAGACTC F: GAACTCCAAACAGCCATTCC Ms647 Schultz et al. 2009 EU913765 (TG)14 115, 117 0.44 0.46 2 R: CCTGCTCCTTCTTTCTGATCC F: TCAACTTCTCAATTTAGGATTCACA Ms663 Schultz et al. 2009 EU913764 (TC)11 290, 294 0.31 0.31 2 R: GCAAAAAGGGATGAGCCATA F: CAGGGGACCTGAGTGCTTATG Hg6.3 Allen et al. 1995 G02092 (GT)18 227, 237 0.36 0.36 2 R: GACCCAGCATCAGAACTCAAG

blah123

Table 4-3. Distribution of private alleles identified at 17 loci in time partitions for any single island (grey), in one partition for any single island (red), and at one island in one partition, followed by another island in a subsequent time partition (teal). No data (--). Data Neo Neo Neo Neo Neo Neo Neo Neo Neo Msc Msc Msc Msc Msc Msc Msc Ms Set 871 1807 10588 12980 14403 4584 7472 15522 15716 17 10 19 01 04 05 03 23 Before

91 91-96 KUR 97-02 114 03-07 All 114 years Before ------91 91-96 MID 97-02 03-07 223 206 All 223 years Before

91 91-96 301 138 PHR 97-02 220 138 03-07 All 138 years Before ------91 91-96 LIS 97-02 169 03-07 118 All 118 169 years Before

91 91-96 119 Lay 97-02 132 03-07 132 All 132 years Before 140 99 301 119 130 136 340 91 91-96 111 172 FFS 97-02 224 03-07 149 All 140 111 224 149 130 172 340 years

124

Table 4-4. Summary of genetic diversity estimates for all loci of entire population (NWHI and MHI combined), according to partitioned time increments (“Before 91,” “91 to 96,” “97 to 02,” “03 to 07”) and over entire time of study (“All years”). Allelic richness with rarefaction (Ar), sample size (N), observed (Ho) and expected (He) heterozygosity, unbiased heterozygosity (UHe), inbreeding coefficient (Fis), Hardy- Weinberg equilibrium (HWE threshold P-value <0.05) likelihood ratio (glb), homozygosity excess (hom), and heterozygosity excess (het), and 95% confidence interval for Fis (Fis(low), Fis(up)) based on 1000 bootstrap iterations in the diveRsity R package [29].

Statistic Before 91 91 to 96 97 to 02 03 to 07 All years Ar 1.93 1.79 2.46 1.92 1.92 N 77 144 296 268 785

Ho 0.39 0.39 0.41 0.39 0.39

He 0.36 0.38 0.40 0.37 0.38

UHe 0.39 0.41 0.40 0.39 0.40

Fis -0.10 -0.04 -0.02 -0.05 -0.05 HWE(glb) 0.75 0.66 0.03 0.52 0.25 HWE(hom) 0.94 0.89 0.36 0.87 0.40 HWE(het) 0.93 0.79 0.54 0.82 0.34

Fis(low) -0.45 -0.24 -0.07 -0.29 -0.28

Fis(up) -0.07 -0.01 0.01 -0.04 -0.03

125

Tables 4-5 (A-E). Genetic diversity summary per island population for all data partitions. Listed from northwest to southeast, the populations in the Northwestern Hawaiian Islands include Kure Atoll (Kur), Midway Atoll (Mid), Pearl and Hermes Reef (PHR), Lisianski (Lis), Laysan (Lay), and French Frigate Shoals (FFS), and Nihoa (Nih). The main Hawaiian Islands include Kauai (Kau), Oahu (Oah), and Molokai (Mol). Allelic richness with rarefaction (Ar), sample size (N), observed (Ho) and expected (He) heterozygosity, unbiased heterozygosity (UHe), inbreeding coefficient (Fis), Hardy-Weinberg equilibrium (HWE threshold P-value < 0.05) likelihood ratio (glb), homozygosity excess (hom), heterozygosity excess (het), and 95% confidence interval for Fis (Fis(low), Fis(up)) based on 999 bootstrap iterations in the diveRsity R package [29]. Bold type indicates significance; underlining highlights the positive inbreeding coefficient (Fis).

(A) All years NWHI MHI Ni Oa Statistic Kur Mid PHR Lis Lay FFS Kau Mol h h Ar 1.92 1.99 1.97 1.97 1.93 1.95 NA 1.86 NA 1.83

N 121 56 122 112 149 217 1 3 1 3

Ho 0.39 0.42 0.41 0.40 0.37 0.40 NA 0.39 NA 0.36

He 0.38 0.41 0.41 0.41 0.39 0.40 NA 0.32 NA 0.31

UHe 0.39 0.42 0.41 0.41 0.39 0.40 NA 0.39 NA 0.37

Fis -0.01 -0.02 -0.02 -0.01 0.03 0.00 NA -0.20 NA -0.17

HWE 0.00 0.00 0.00 0.00 0.00 0.00 NA 1.00 NA 1.00 (glb)

HWE 0.02 0.09 0.00 0.31 0.79 0.02 NA 1.00 NA 1.00 (hom)

HWE 0.01 0.68 0.00 0.05 0.00 0.00 NA 1.00 NA 1.00 (het) - Fis(low) -0.05 -0.06 -0.05 -0.03 0.01 NA -1.00 NA -1.00 0.03

Fis(up) 0.01 0.01 0.01 0.01 0.05 0.02 NA -0.20 NA -0.17 1.92 1.99 1.97 1.97 1.93 1.95 NA 1.86 NA 1.83

126

(B) Before 91 NWHI MHI Ni Oa Statistic Kur Mid PHR Lis Lay FFS Kau Mol h h Ar 1.88 -- 1.95 NA 1.95 1.93 ------N 3 -- 4 1 8 61 ------

Ho 0.37 -- 0.39 NA 0.40 0.40 ------

He 0.31 -- 0.35 NA 0.38 0.39 ------

UHe 0.37 -- 0.40 NA 0.41 0.39 ------

Fis -0.19 -- -0.15 NA -0.04 -0.01 ------HWE 1.00 -- 1.00 NA 1.00 0.01 ------(glb) HWE 1.00 -- 1.00 NA 1.00 0.76 ------(hom) HWE 1.00 -- 1.00 NA 1.00 0.74 ------(het)

Fis(low) -1.00 -- -0.53 NA -0.20 -0.05 ------

Fis(up) -0.19 -- -0.11 NA -0.02 0.01 ------

127

(C)

91 to 96 NWHI MHI Ni Oa Statistic Kur Mid PHR Lis Lay FFS Kau Mol h h

Ar 1.75 1.91 1.78 1.80 1.73 1.77 ------

N 22 2 40 13 7 60 ------

Ho 0.39 0.42 0.40 0.40 0.33 0.40 ------

He 0.38 0.33 0.40 0.40 0.36 0.40 ------

UHe 0.39 0.44 0.41 0.42 0.38 0.41 ------

Fis -0.01 -0.27 0.00 -0.01 0.07 0.00 ------HWE 0.95 1.00 0.00 1.00 0.99 0.04 ------(glb)

HWE 1.00 1.00 0.40 1.00 1.00 0.95 ------(hom)

HWE 0.98 1.00 0.42 1.00 1.00 0.35 ------(het)

Fis(low) -0.11 -1.00 -0.05 -0.11 -0.13 -0.04 ------

Fis(up) 0.04 -0.27 0.03 0.02 0.09 0.03 ------

128

(D) 97 to 02 NWHI MHI Ni Oa Statistic Kur Mid PHR Lis Lay FFS Kau Mol h h Ar 2.44 2.50 2.47 2.41 2.48 2.48 ------N 48 25 51 37 78 57 ------

Ho 0.37 0.43 0.42 0.41 0.38 0.43 ------

He 0.38 0.41 0.40 0.40 0.39 0.40 ------

UHe 0.38 0.42 0.41 0.41 0.39 0.41 ------

Fis 0.00 -0.03 -0.04 -0.01 0.02 -0.06 ------HWE 0.00 0.19 0.00 0.00 0.00 0.00 ------(glb)

HWE 0.53 0.76 0.08 0.42 0.40 0.01 ------(hom)

HWE 0.09 0.98 0.95 0.28 0.05 0.89 ------(het)

Fis(low) -0.05 -0.10 -0.09 -0.07 -0.01 -0.10 ------

Fis(up) 0.04 -0.01 -0.01 0.02 0.04 -0.03 ------

129

(E) 03 to 07 NWHI MHI Ni Oa Statistic Kur Mid PHR Lis Lay FFS Kau Mol h h

Ar 1.91 1.99 1.96 1.96 1.91 1.93 NA 1.86 NA 1.83

N 48 29 27 61 56 39 1 3 1 3

Ho 0.40 0.42 0.40 0.40 0.36 0.37 NA 0.39 NA 0.36

He 0.38 0.41 0.40 0.40 0.38 0.39 NA 0.32 NA 0.31

UHe 0.39 0.41 0.40 0.40 0.38 0.39 NA 0.39 NA 0.37

Fis -0.05 -0.01 -0.03 -0.02 0.02 0.04 NA -0.20 NA -0.17 HWE 0.04 0.04 1.00 0.02 0.12 0.92 NA 1.00 NA 1.00 (glb)

HWE 0.17 0.96 1.00 0.86 1.00 1.00 NA 1.00 NA 1.00 (hom)

HWE 1.00 0.96 1.00 0.87 0.16 0.57 NA 1.00 NA 1.00 (het)

Fis(low) -0.10 -0.07 -0.09 -0.06 -0.02 -0.02 NA -1.00 NA -1.00

Fis(up) -0.02 0.02 -0.02 0.00 0.04 0.07 NA -0.20 NA -0.17

130

Tables 4-6 (A-E). Genetic diversity values for pairwise comparisons of Fst and analogous estimators between populations in the “All years” data set (N = 785). (Fst values greater than 0.05 are highlighted in bold. Estimators are as follows: Weir & Cockerham’s Fst [31], Nei and Chesser’s gst [32], Hedrick’s Gst [33], Meirmans and Hedrick’s GGst [34], and Jost’s D [35]. Results from pairwise comparisons across the four other data partitions are provided in Appendix I.) (A) Fst NWHI MHI Pop Kur Mid PHR Lis Lay FFS Kau Mol Kur -- Mid 0.01 -- PHR 0.01 0.01 -- Lis 0.02 0.01 0.02 -- Lay 0.02 0.01 0.02 0.01 -- FFS 0.02 0.01 0.02 0.01 0.01 -- Kau 0.01 -0.01 0.01 0.00 0.00 0.01 -- Mol 0.04 0.04 0.06 0.05 0.03 0.03 0.01 --

(B) D NWHI MHI Pop Kur Mid PHR Lis Lay FFS Kau Mol Kur -- Mid 0.00 -- PHR 0.00 0.00 -- Lis 0.00 0.00 0.00 -- Lay 0.00 0.00 0.01 0.00 -- FFS 0.00 0.00 0.00 0.00 0.00 -- Kau 0.00 0.00 0.00 0.00 0.00 -- -- Mol 0.00 0.00 0.00 0.00 0.00 0.00 0.00 --

(C) gst NWHI MHI Pop Kur Mid PHR Lis Lay FFS Kau Mol Kur -- Mid 0.01 -- PHR 0.01 0.00 -- Lis 0.01 0.01 0.01 -- Lay 0.01 0.01 0.01 0.01 -- FFS 0.01 0.01 0.01 0.01 0.00 -- Kau 0.00 0.00 0.01 0.00 0.00 0.01 -- Mol 0.02 0.02 0.03 0.03 0.02 0.02 0.01 --

131

(D) Gst NWHI MHI Pop Kur Mid PHR Lis Lay FFS Kau Mol Kur -- Mid 0.01 -- PHR 0.01 0.01 -- Lis 0.03 0.02 0.02 -- Lay 0.02 0.02 0.02 0.02 -- FFS 0.02 0.02 0.02 0.01 0.01 -- Kau 0.01 -0.01 0.01 0.01 0.00 0.02 -- Mol 0.04 0.05 0.07 0.07 0.04 0.04 0.02 --

(E) GGst NWHI MHI Pop Kur Mid PHR Lis Lay FFS Kau Mol Kur -- Mid 0.02 -- PHR 0.02 0.01 -- Lis 0.04 0.02 0.03 -- Lay 0.03 0.02 0.04 0.02 -- FFS 0.03 0.02 0.03 0.02 0.01 -- Kau 0.01 -0.01 0.02 0.01 0.00 0.02 -- Mol 0.06 0.07 0.10 0.09 0.05 0.06 0.03 --

132

Tables 4-7 (A-E). Matrices of P-values for pairwise Fisher’s Exact test comparisons of genotype distributions between island populations. (Color codes are used to highlight the comparisons according to time increments sampled. NA indicates that no data were available. P-values (< 0.05) in bold indicate a significant difference in the distribution of genotypes between island populations.) (A) Before 1991 Pop Kur Mid PHR Lis Lay FFS Kau Mol Kur -- NA 1.00 NA 1.00 0.67 NA NA Mid 1.00 -- NA NA NA NA NA NA PHR 0.01 0.99 -- 0.94 0.13 NA NA Lis 0.05 1.00 0.06 -- NA NA NA NA Lay 0.33 1.00 0.39 0.91 -- 0.32 NA NA FFS 0.01 1.00 0.00 0.06 0.23 -- NA NA Kau NA NA NA NA NA NA -- NA Mol NA NA NA NA NA NA NA -- (B) 1991 to 1996

(C) 1997 to 2002 Pop Kur Mid PHR Lis Lay FFS Kau Mol Kur -- 0.00 0.00 0.00 0.00 0.00 NA NA Mid 0.01 -- 0.05 0.01 0.00 0.00 NA NA PHR 0.10 0.15 -- 0.00 0.00 0.00 NA NA Lis 0.00 0.00 0.03 -- 0.00 0.00 NA NA Lay 0.00 0.00 0.00 0.00 -- 0.00 NA NA FFS 0.00 0.15 0.08 0.00 0.11 -- NA NA Kau 0.94 0.96 0.74 0.80 0.90 0.98 -- NA Mol 0.57 0.91 0.36 0.46 0.78 0.98 1.00 -- (D) 2003 to 2007

(E) All years Pop Kur Mid PHR Lis Lay FFS Kau Mol Kur -- 0.00 0.00 0.00 0.00 0.00 0.89 0.84 Mid -- 0.00 0.00 0.00 0.00 0.97 0.80 PHR -- 0.00 0.00 0.00 0.77 0.43 Lis -- 0.00 0.00 0.86 0.48 Lay -- 0.00 0.94 0.77 FFS -- 0.68 0.83 Kau -- 1.00 Mol --

133

Table 4-8. First generation migrant (F0) assignments across islands using GENECLASS, version 2 [42]. Islands of first observation (O) and islands identified for emigration (A) as indicated by |–log (L_home/L_max) values. Significant (P < 0.01) values for –log(L) calculations in bold.

-log(L) by Island Sample -log(L_home ID (F0) Sex O /L_max) A Kur Mid PHR Lis Lay FFS Nih Kau Oah Mol RI25 F Mol 5.64 Kur 16.36 16.80 17.05 17.48 17.02 16.65 21.67 18.27 25.80 22.00 R5AL F Mol 5.54 Kur 16.05 18.16 18.37 18.49 17.24 17.10 25.84 21.30 27.91 21.59 RV08 M Oah 8.57 Lay 21.50 20.94 21.04 21.03 20.57 21.62 27.23 26.15 0.00 27.51 RI19 M Kau 2.22 PHR 17.74 18.16 17.63 17.71 18.00 18.18 22.53 19.85 26.74 19.50 RO26 F Kau 3.08 Lay 20.08 19.58 20.11 19.50 18.60 19.93 24.88 21.69 27.20 24.32 RO12 M Kau 6.93 Mid 18.91 18.04 18.89 19.45 18.31 18.70 24.87 24.97 27.81 24.59 RB22 M Nih 5.72 Kur 19.16 20.16 19.93 21.16 22.30 22.13 0.00 21.25 28.03 24.00

134 Y323 F FFS 3.42 PHR 24.69 22.21 20.72 22.09 22.96 24.13 28.45 26.60 30.18 24.76

YC42 F FFS 2.24 Mid 18.59 17.95 18.73 19.62 19.63 20.18 24.56 22.04 26.01 23.57 YP27 M FFS 1.65 Kur 17.01 18.27 19.33 19.40 18.63 18.66 25.65 22.99 24.95 24.18 YE32 F FFS 2.18 Mol 16.46 17.76 17.62 17.24 16.35 16.38 24.07 19.33 25.19 14.21 YI13 F FFS 2.71 Kau 21.53 20.32 20.66 19.93 19.28 20.31 25.94 17.60 28.82 24.04 YB26 F FFS 1.75 Lis 20.03 19.49 18.46 18.14 18.33 19.89 27.20 24.38 25.23 24.81 TK25 M Lay 2.18 Lis 29.63 26.33 26.59 24.94 27.12 25.60 30.55 28.28 25.90 30.02 TY78 M Lay 2.94 Kau 18.40 18.73 19.00 20.30 18.97 19.24 23.91 16.03 26.67 22.50 TY80 F Lay 1.65 Kur 20.15 21.22 20.57 20.90 21.80 20.61 28.27 25.75 29.97 21.61 T809 M Lay 1.68 Kur 18.57 19.94 19.27 19.31 20.25 19.34 28.90 27.51 24.47 28.65 TO74 M Lay 2.21 Mid 21.85 17.46 18.45 21.07 19.67 20.51 23.35 20.01 26.83 25.32 TO24 F Lay 2.78 Kur 17.59 19.11 18.21 19.65 20.36 19.98 23.33 23.04 25.26 27.16 GC06 F Lis 2.01 Mid 20.95 19.77 20.41 21.78 20.75 20.80 27.80 24.51 27.17 25.56 GE26 F Lis 1.75 Lay 14.72 14.16 14.47 15.43 13.69 14.17 20.89 15.43 24.79 15.33 GR40 M Lis 1.93 Mol 22.69 21.13 22.44 21.58 20.68 20.45 27.20 22.51 27.99 19.64 PO28 F Lis 2.24 PHR 18.98 19.75 18.85 21.10 21.56 20.76 25.68 26.92 26.30 24.89

blah134

BC37 F PHR 2.11 FFS 21.01 21.35 21.97 20.67 20.57 19.85 25.60 26.55 25.41 23.14 B4AN F PHR 2.09 Kur 16.38 17.18 18.47 18.42 16.81 16.98 23.23 20.02 27.74 20.93 B5AM F PHR 1.84 Mid 23.21 22.22 24.07 24.62 22.29 24.71 29.77 28.16 29.43 27.25 BQ25 M PHR 2.48 FFS 24.15 22.47 23.29 22.90 21.88 20.80 26.09 23.11 26.26 22.04 BY00 F PHR 1.61 Kur 15.23 16.80 16.84 16.79 17.70 16.27 20.35 20.90 25.40 21.78 B6AL F PHR 2.01 Mid 23.70 21.71 23.72 22.69 23.45 23.26 23.97 25.61 26.01 22.63 RH12 F Mid 1.99 Kur 19.53 21.52 20.86 21.08 20.62 20.81 26.99 25.67 27.94 25.99 RH20 M Mid 2.47 PHR 19.91 20.99 18.52 20.90 21.70 21.14 23.45 22.61 27.70 24.20 RM12 F Mid 2.35 Lay 19.78 21.24 20.31 19.56 18.88 18.89 25.34 22.11 30.04 19.94 RR00 F Mid 1.85 FFS 17.10 18.43 18.19 17.59 17.78 16.58 23.99 22.52 26.93 21.78 RR02 F Mid 1.78 Lay 21.13 19.60 21.37 20.77 17.82 18.30 25.43 26.14 24.36 21.68 RR04 M Mid 2.52 FFS 21.56 21.59 20.86 20.12 19.40 19.07 28.48 27.93 26.67 23.22 PV00 F Mid 1.72 Lis 17.36 17.65 17.02 15.93 17.66 16.73 25.25 21.90 24.72 23.44 KI03 M Kur 2.01 Lis 20.26 19.80 19.92 18.25 20.70 20.24 24.61 21.62 26.32 25.85

135 KI39 M Kur 1.77 Lis 19.90 19.54 19.70 18.13 18.95 19.72 27.70 22.91 29.22 23.90

KO42 M Kur 1.68 Mid 20.39 18.72 19.80 20.31 19.20 19.61 24.20 20.81 24.78 23.84 KO30 F Kur 2.26 PHR 24.03 23.01 21.77 23.31 25.25 26.33 26.91 27.04 30.33 29.55 Total # migrants assigned 10 7 5 5 5 4 0 2 0 2

blah135

Table 4-9 (A-E). Results of bidirectional estimates of relative migration rates using the divMigrate function in the diveRsity R package [29]. (Underlined values highlight relative migration rates of 0.80 or greater. Islands listed in the far left column are interpreted as the island from which a migrant leaves. For example, the migration rate from Kure (row 1) to Midway (column 2) is 0.63, whereas the migration rate from Midway atoll (row 2) to Kure (column 1) is 0.50.)

(A) All NWHI MHI years

Kur Mid PHR Lis Lay FFS Kau Mol

Kur -- 0.63 0.65 0.39 0.42 0.49 0.05 0.03 Mid 0.50 -- 0.89 0.49 0.49 0.51 0.05 0.03 PHR 0.55 0.85 -- 0.43 0.38 0.52 0.04 0.03 Lis 0.37 0.49 0.44 -- 0.57 0.71 0.04 0.03 Lay 0.46 0.54 0.41 0.62 -- 1.00 0.04 0.04 FFS 0.46 0.57 0.52 0.68 0.93 -- 0.04 0.03 Kau 0.12 0.15 0.14 0.16 0.19 0.15 -- 0.03 Mol 0.11 0.11 0.10 0.10 0.10 0.13 0.02 --

136

(B) Before 91 NWHI MHI

Kur Mid PHR Lis Lay FFS Kau Mol

Kur -- -- 0.27 -- 0.28 0.64 -- -- Mid ------PHR 0.29 ------0.28 0.6 -- -- Lis ------Lay 0.16 -- 0.15 -- -- 0.49 -- -- FFS 0.18 -- 0.21 -- 0.99 ------Kau ------Mol ------

137

(C) 91 to 96 NWHI MHI

Kur Mid PHR Lis Lay FFS Kau Mol

Kur -- 0.14 0.62 0.32 0.23 0.73 -- -- Mid 0.26 -- 0.27 0.18 0.14 0.28 -- -- PHR 0.56 0.11 -- 0.42 0.25 0.95 -- -- Lis 0.34 0.09 0.45 -- 0.25 0.61 -- -- Lay 0.29 0.09 0.47 0.33 -- 0.49 -- -- FFS 0.64 0.12 1.00 0.54 0.28 ------Kau ------Mol ------

138

(D) 97 to 02 NWHI MHI

Kur Mid PHR Lis Lay FFS Kau Mol

Kur -- 0.53 0.58 0.51 0.64 0.57 -- -- Mid 0.43 -- 0.65 0.59 0.53 0.37 -- -- PHR 0.50 0.61 -- 0.55 0.50 0.53 -- -- Lis 0.48 0.61 0.57 -- 0.90 0.76 -- -- Lay 0.72 0.57 0.62 0.94 -- 1.00 -- -- FFS 0.48 0.40 0.58 0.67 0.85 ------Kau ------Mol ------

139

(E) 03 to 07 NWHI MHI

Kur Mid PHR Lis Lay FFS Kau Mol

Kur -- 0.80 0.78 0.57 0.44 0.54 0.08 0.05 Mid 0.64 -- 0.75 0.64 0.62 0.78 0.07 0.05 PHR 0.63 0.73 -- 0.56 0.44 0.61 0.05 0.03 Lis 0.53 0.71 0.54 -- 0.66 0.70 0.06 0.04 Lay 0.48 0.68 0.44 0.74 -- 0.91 0.06 0.06 FFS 0.55 0.84 0.59 0.79 1.00 -- 0.06 0.05 Kau 0.21 0.22 0.17 0.25 0.27 0.21 -- 0.05 Mol 0.17 0.18 0.14 0.16 0.16 0.19 0.03 --

140

APPENDIX

Tables A-1 (A-D). Pairwise comparisons of Fst and analogous genetic diversity estimators (E) with lower (-) and upper (+) bounds of 95% confidence intervals for all data partitions. (No data due to lack of population sampling effort is indicated by (--). The “03 to 07” is the only partition that includes populations from Kauai and Molokai. However, sample size is very small for these islands (N = 3).

(A) Fst D gst Gst GGst Before 91 Est - + Est - + Est - + Est - + Est - + FFS vs. 0.01 -0.01 0.03 0.00 0.00 0.00 0.00 0.00 0.01 0.01 -0.01 0.03 0.01 -0.02 0.04 Lay FFS ------vs. Lis FFS vs. 0.03 0.00 0.06 0.00 0.00 0.01 0.01 0.00 0.03 0.03 0.00 0.08 0.05 -0.02 0.09 PHR FFS vs. ------Mid FFS vs. 0.02 -0.01 0.04 0.00 0.00 0.01 0.01 -0.01 0.03 0.02 -0.02 0.06 0.03 -0.03 0.07 Kur Lay ------vs. Lis Lay vs. 0.03 -0.01 0.08 0.00 0.00 0.01 0.02 -0.01 0.04 0.04 -0.01 0.10 0.05 -0.04 0.12 PHR Lay vs. ------Mid Lay vs. 0.01 -0.04 0.06 0.00 0.00 0.00 0.01 -0.02 0.03 0.01 -0.04 0.08 0.02 -0.07 0.10 Kur Lis vs. ------PHR Lis vs. ------Mid Lis vs. ------Kur PHR vs. ------Mid PHR - vs. -0.09 -0.03 -0.02 -0.04 0.00 -0.03 -0.04 -0.02 -0.07 -0.10 -0.03 -0.10 -0.12 -0.02 0.06 Kur Mid vs. ------Kur

141

(C) Fst D gst Gst GGst 97 to 02 Est - + Est - + Est - + Est - + Est - +

FFS vs. 0.01 0.00 0.02 0.00 0.00 0.00 0.01 0.00 0.01 0.01 0.00 0.02 0.02 0.00 0.03 Lay FFS vs. 0.01 0.01 0.02 0.00 0.00 0.00 0.01 0.00 0.01 0.01 0.01 0.02 0.02 0.00 0.02 Lis FFS vs. 0.02 0.01 0.03 0.00 0.00 0.01 0.01 0.00 0.02 0.02 0.01 0.04 0.03 0.01 0.04 PH R FFS vs. 0.03 0.01 0.04 0.01 0.00 0.01 0.01 0.01 0.02 0.03 0.02 0.05 0.05 0.01 0.06 Mid FFS vs. 0.02 0.01 0.03 0.00 0.00 0.01 0.01 0.01 0.02 0.03 0.01 0.04 0.04 0.01 0.04 Kur Lay vs. 0.01 0.00 0.01 0.00 0.00 0.00 0.00 0.00 0.01 0.01 0.00 0.02 0.01 0.00 0.02 Lis Lay vs. 0.02 0.01 0.03 0.00 0.00 0.01 0.01 0.01 0.02 0.02 0.01 0.04 0.03 0.01 0.04 PH R Lay vs. 0.02 0.01 0.03 0.00 0.00 0.01 0.01 0.00 0.01 0.02 0.01 0.03 0.03 0.00 0.04 Mid Lay vs. 0.02 0.01 0.03 0.00 0.00 0.00 0.01 0.00 0.01 0.02 0.01 0.03 0.03 0.00 0.04 Kur Lis vs. 0.02 0.01 0.03 0.00 0.00 0.01 0.01 0.00 0.01 0.02 0.01 0.04 0.03 0.00 0.04 PH R Lis vs. 0.01 0.00 0.02 0.00 0.00 0.00 0.01 0.00 0.01 0.01 0.00 0.03 0.02 0.00 0.03 Mid

Lis 0.02 0.01 0.03 0.00 0.00 0.01 0.01 0.00 0.02 0.02 0.01 0.04 0.03 0.00 0.05 vs.

142

Kur

PH R 0.01 0.00 0.02 0.00 0.00 0.00 0.01 0.00 0.01 0.01 0.00 0.03 0.02 0.00 0.03 vs. Mid PH R 0.02 0.01 0.03 0.00 0.00 0.01 0.01 0.00 0.02 0.02 0.01 0.04 0.03 0.01 0.04 vs. Kur Mid vs. 0.02 0.01 0.03 0.00 0.00 0.01 0.01 0.00 0.02 0.02 0.01 0.04 0.03 0.00 0.04 Kur

143

(D) Fst D gst Gst GGst 03 to 07 Est - + Est - + Est - + Est - + Est - +

Mol - - - - - vs. 0.01 0.09 0.00 0.00 0.01 0.05 0.02 0.10 0.03 0.14 0.06 0.01 0.03 0.06 0.10 Kau

Mol - - - - vs. 0.03 0.10 0.00 0.00 0.00 0.02 0.06 0.05 0.13 0.07 0.16 0.03 0.01 0.02 0.05 FFS

Mol - - - - vs. 0.03 0.08 0.00 0.00 0.01 0.02 0.05 0.04 0.10 0.06 0.12 0.01 0.01 0.01 0.04 Lay

Mol - vs. 0.06 0.00 0.11 0.00 0.00 0.01 0.03 0.00 0.06 0.07 0.01 0.14 0.10 0.17 0.02 Lis

Mol - vs. 0.08 0.01 0.16 0.00 0.00 0.02 0.04 0.00 0.09 0.10 0.01 0.19 0.14 0.22 0.03 PHR Mol - - - - vs. 0.03 0.09 0.00 0.00 0.01 0.02 0.05 0.04 0.11 0.06 0.14 0.02 0.01 0.02 0.05 Mid

Mol - - vs. 0.05 0.00 0.10 0.00 0.00 0.01 0.02 0.00 0.05 0.05 0.11 0.07 0.14 0.01 0.04 Kur

Kau - - - - vs. 0.01 0.06 0.00 0.00 0.00 0.01 0.04 0.02 0.08 0.02 0.11 0.04 0.02 0.04 0.06 FFS Kau - - - - vs. 0.01 0.05 0.00 0.00 0.00 0.00 0.03 0.01 0.07 0.02 0.10 0.03 0.02 0.04 0.06 Lay

Kau - - - - vs. 0.01 0.06 0.00 0.00 0.00 0.01 0.03 0.01 0.08 0.02 0.10 0.03 0.02 0.04 0.06 Lis

Kau - - - - vs. 0.03 0.08 0.00 0.00 0.00 0.01 0.05 0.03 0.10 0.05 0.13 0.02 0.01 0.03 0.05 PHR Kau - - - - - vs. 0.01 0.05 0.00 0.00 0.00 0.03 0.01 0.06 0.01 0.08 0.03 0.01 0.02 0.04 0.06 Mid

144

(D) Fst D gst Gst GGst 03 to 07 Est - + Est - + Est - + Est - + Est - +

Kau - - - - vs. 0.01 0.04 0.00 0.00 0.00 0.00 0.02 0.00 0.04 0.00 0.06 0.02 0.01 0.03 0.05 Kur

FFS vs. 0.01 0.00 0.01 0.00 0.00 0.00 0.00 0.00 0.01 0.01 0.00 0.02 0.01 0.00 0.02 Lay FFS vs. 0.01 0.01 0.02 0.00 0.00 0.00 0.01 0.00 0.01 0.01 0.01 0.02 0.02 0.00 0.03 Lis

FFS vs. 0.01 0.00 0.02 0.00 0.00 0.00 0.01 0.00 0.01 0.01 0.00 0.02 0.02 0.00 0.03 PHR

FFS - vs. 0.01 0.00 0.01 0.00 0.00 0.00 0.00 0.00 0.01 0.01 0.00 0.02 0.01 0.02 0.01 Mid FFS vs. 0.02 0.01 0.03 0.00 0.00 0.01 0.01 0.00 0.02 0.02 0.01 0.03 0.03 0.01 0.04 Kur

Lay vs. 0.02 0.01 0.02 0.00 0.00 0.01 0.01 0.00 0.01 0.02 0.01 0.03 0.03 0.01 0.03 Lis

Lay vs. 0.02 0.01 0.04 0.00 0.00 0.01 0.01 0.01 0.02 0.03 0.01 0.04 0.04 0.01 0.05 PHR Lay vs. 0.01 0.00 0.02 0.00 0.00 0.00 0.01 0.00 0.01 0.01 0.00 0.03 0.02 0.00 0.03 Mid

Lay vs. 0.03 0.02 0.04 0.01 0.00 0.01 0.01 0.01 0.02 0.03 0.02 0.04 0.04 0.01 0.05 Kur

Lis vs. 0.02 0.00 0.04 0.00 0.00 0.00 0.01 0.00 0.02 0.02 0.00 0.04 0.03 0.00 0.05 PHR Lis vs. 0.01 0.01 0.02 0.00 0.00 0.01 0.01 0.00 0.01 0.01 0.01 0.03 0.02 0.00 0.03 Mid

Lis vs. 0.02 0.01 0.04 0.00 0.00 0.01 0.01 0.01 0.02 0.03 0.01 0.04 0.04 0.01 0.05 Kur

145

(D) Fst D gst Gst GGst 03 to 07 Est - + Est - + Est - + Est - + Est - +

PHR - vs. 0.00 0.00 0.01 0.00 0.00 0.00 0.00 0.00 0.01 0.01 0.00 0.01 0.01 0.02 0.01 Mid

PHR vs. 0.01 0.00 0.02 0.00 0.00 0.00 0.00 0.00 0.01 0.01 0.00 0.02 0.02 0.00 0.03 Kur Mid vs. 0.01 0.00 0.02 0.00 0.00 0.00 0.00 0.00 0.01 0.01 0.00 0.02 0.02 0.00 0.02 Kur

146

Supplementary Material

Data management and quality control for consensus genotyping

Including replicates and 89 controls, 3,567 samples were processed and yielded genotype data for 2,404 individuals using 24 primers characterized by Mihnovets et al. (2016). Of these samples, 23 failed to yield results, and 768 were removed due to evidence of contamination. It was not feasible to generate a uniform number of PCR replicates for all samples. However, between two and five replicate samples were amplified for 291 individuals, which constitute 12.10% of all individuals sampled. Out of 2,404 individuals, 1192 were completely genotyped at all 24 loci.

A database was created using Microsoft Access (v2007) to archive all individual genotypes generated using the 24 primers of Mihnovets et al. (2016) and 18 primers of Schultz et al. (REFS) (Table 2), and to facilitate integrative queries for genetic and population demography data. Any combination of variables can be queried according to a unique identifier assigned to each individual seal. The database contains datasets consisting of five demographic survey variables (year first observed, year sample collected, age, sex, size first observed and at sample collection), seven spatial variables (island first observed, island first sampled, translocation (island from/to), island code, island name, longitude/latitude coordinates), seven spatial and temporal variables associated with individual translocations between islands, DNA archive storage locations for each DNA sample, and all available genotype data associated with unique identification numbers, for a total of 2,409 individuals (Figure S1). The database also contains data for an additional 1,199 samples that were either replicates of certain individuals or that originated from individuals of unknown ID.

Seal IDs were used to query replicate genotypes in Access, and the results were exported to an Excel spreadsheet. Replicate data were then compared by using an automated command in Excel to identify and flag conflicting allele calls between replicate

147

genotypes of each locus. These discrepancies were then individually examined by eye and were compared by re-examining their associated spectrogram data in GENEMAPPER. When conflicting calls could not be resolved between two replicate genotypes due to conflicting heterozygote and homozygote calls at a given locus, the discrepancies were re-coded as “missing data” at the locus in question. Final consensus genotypes for replicated individuals were then archived in a master Access database file, which contains a single final genotype for each individual seal. The master file was queried to generate a table consisting of individuals for which complete genotypes at all 42 loci were available (to minimize confounding effects associated with missing data).

148

Figure 4S-1. Configuration of relational database created in Microsoft Access to manage genetic and population demography data.

149

blah149

CHAPTER 5. CONCLUSION

In this dissertation, evolutionary relationships, genetic variation, and gene flow were investigated in the genus Neomonachus, which is comprised of two New World, warm-water pinnipeds: the extinct Caribbean monk seal (N. tropicalis) and the endangered Hawaiian monk seal (N. schauinslandi). This study included three main objectives: (1) to resolve outstanding molecular systematic questions pertaining to the extinct N. tropicalis and closely related species, (2) to develop new microsatellite marker tools that increase capabilities for individual genotyping and population genetic analyses of the endangered N. schauinslandi, and (3) to assess the genetic status (e.g., genetic diversity and population differentiation) of N. schauinslandi, with consideration of relevant conservation and management practices.

To achieve the first objective, minimally destructive collection techniques were used to obtain tissue samples from museum specimens of N. tropicalis dating back over

150 years, laboratory protocols were optimized to extract total genomic DNA from the samples, and new mitochondrial genome sequence data generated by collaborators at

Johns Hopkins University were annotated and characterized. The resulting mitochondrial genome was then used in molecular systematic analyses to resolve the phylogeny of tribe

Monachini, with strong support for the placement of M. monachus as basal to sister species N. tropicalis and N. schauinslandi. The added molecular resolution provided by the complete mitogenome of N. tropicalis also clarified the phylogenetic placement of the warm water-loving tribe Monachini as basal to cold water- loving sister tribes Miroungini and Lobodontini.

150

To achieve objectives 2 and 3, novel molecular laboratory protocols were optimized for 24 new markers, to facilitate the efficient microsatellite genotyping of N. schauinslandi tissue samples collected and archived by the Hawaiian Monk Seal

Research Program over the past four decades. A total of 3,567 tissue samples were processed, ultimately yielding varying degrees of genotype coverage for 2,404 individuals. Using the new microsatellite markers alone, 1192 individuals were completely genotyped at all 24 loci. By integrating the new microsatellite marker data with a previous data set that included 18 other loci, genotypes were generated for 2404 individuals. Of these, there were complete genotypes (no missing data) for 785 individuals. The complete genotype data set was used to assess genetic diversity and population differentiation in a battery of genetic analyses, and some of the new analytical approaches prompted interest among managers in testing with more uniform sample sizes. Sample sizes suffered due to the limitation of analyses including only individuals with complete coverage across all 42 loci. However, doing so enabled a contrast between using complete and incomplete genotype data for population level analyses. Overall, results corroborated previous findings of high levels of gene flow and low levels of genetic diversity. Thus, there was no evidence to generate concern regarding the potential impact of translocation (used as a management tool to facilitate survival) on the population genetic processes of N. schauinslandi. However, the resulting genotype data will be useful for further consideration of individual-level genetic processes within a conservation and management context.

Contributions and Implications

This research involved evolutionary and population genetic research at several hierarchical levels, and across a scale unlike that in previous research for New World

151

monk seals, with various implications for conservation and future evolutionary and ecological research.

Chapter Two contributes to a rich body of pinniped studies by using ancient DNA methods in combination with cutting-edge high throughput sequencing technologies to sequence the first complete mitochondrial genome of an extinct monk seal, N. tropicalis.

This provides long sought-after molecular resolution and perspective on the evolutionary history of pinnipeds, particularly for tribe Monachini. Furthermore, the mitogenome data will be useful to future total-evidence-based evolutionary studies seeking to resolve discrepancies between fossil, neo-morphological, and molecular evidence. In a more contemporary context, since extinction risk and loss of evolutionary history in the age of the Anthropocene are often not phylogenetically random (Cadotte and Davies, 2010;

Faith, 1992; Moritz and Faith, 1998; Purvis et al., 2000), the results of this study make pinnipeds a compelling case for future studies of biodiversity loss among closely related species. For example, the complete mitogenome sequence will serve as an important reference in future efforts to examine mitochondrial haplotypes from other museum specimens of N. tropicalis, in order to capture a snapshot of the extent of genetic diversity that remained just prior its extinction. Additionally, the technical pipeline used to sequence the mitogenome also generated raw data for total genomic fragments as a byproduct, thereby offering a rare opportunity to piece together chromosomal sequences of N. tropicalis in future inter- and intraspecific comparative genomic investigations.

In Chapter Three, a set of 24 new nuclear microsatellite markers was optimized to expand genotype profiles for 2,404 individuals of N. schauinslandi, to support population level investigations as part of this dissertation and additional individual level

152

investigations in the near future. The results of this work constitute a more than two-fold increase in the number of molecular markers now available for population genetic analyses, and thus, the largest genetic data set ever generated for N. schauinslandi. Yet summary statistics of the new markers indicate comparable levels of genetic variation to those reported in previous studies, which reinforces concerns about the possible threats to species survival associated with low genetic diversity.

In Chapter Four, results of the first exhaustive investigation to use strictly complete genotypes (no missing data) for N. schauinslandi suggest that the current extent of genotypic coverage is sufficient for population level analyses of genetic diversity and gene flow. However, additional fine scale studies are needed to better understand individual level processes that may be contributing to observed genetic patterns of interest for population recovery efforts. For example, it is generally accepted that translocation helps to maintain genetic diversity and does not introduce genetic threats such as outbreeding depression (Schultz, 2011). By facilitating migration and dispersal in small populations, translocation is used as a management tool to maintain meta- population structure, to increase individual fitness (e.g., reproductive potential and juvenile survival), and to improve species’ status (Griffith et al., 1989; Storfer, 1999).

Nevertheless, specific genetic factors that are not explicitly accounted for in current N. schauinslandi translocation protocols are worthy of further consideration by managers seeking to tailor protocols for optimal success.

The practice of translocating individuals to promote survival commonly includes broad assumptions associated with underlying ecological and evolutionary processes

(e.g., population equilibrium, the demographic equality of migrants versus non-migrants,

153

relationships between population size and effective population size), which may contribute to observed genetic patterns. Such assumptions prompt questions that are relevant to the genetic status of N. schauinslandi (Table 1). To address these and other questions of conservation and management interest, work is currently underway to include the use of presently available microsatellite data in order to elucidate relatedness among individuals, male and female breeding patterns, and fitness. Resulting answers are expected to help managers by enabling them to optimize for potential adaptive capacity and fitness when prioritizing conservation efforts or selecting candidates for translocation. Continuous genetic monitoring could be implemented as a standard post- translocation protocol to enable detection of genetic diversity declines and subsequent adjustment of translocation strategies to reduce further loss (Goossens et al., 2002).

Next Steps

Genetic analyses conducted to date constitute a foundation upon which to develop a further understanding of the importance of genetic diversity for species persistence.

Additional research efforts currently underway aim to understand how limited genetic diversity affects individual fitness, and to elucidate variables associated with male mating strategies and reproductive output that impact population viability. This requires individual-level relatedness and parentage analyses, and an important outcome will be the development of a pedigree with which to directly estimate fitness. Preliminary simulations and power analyses based on 785 individuals suggest that the newly expanded genotypic coverage is sufficient for inference of parental relationships within a strict confidence range. However, preliminary parentage assignment results indicate that additional genotyping is necessary to increase samples sizes of offspring and candidate fathers in order to construct a pedigree for further fitness-related investigation.

154

Continuation of these efforts will enable conservation practitioners to further consider ecological and evolutionary processes that influence population viability (e.g., dispersal, kinship, sexual selection, and their influences on fitness (Kruuk and Hill, 2008; Wilson and Ferguson, 2002)).

Dispersal

Translocation has been used as a management technique for several decades to minimize mortality. This creates a challenge when using genetic data to infer non- assisted movement patterns. However, pedigree information could be useful for investigating fitness implications for individuals that remain in natal territory when breeding or that reproduce in different breeding sites throughout their lifetimes.

Kinship

Adult monk seals exhibit energetically costly behavior such as alloparenting and infanticide, which may be positively or negatively associated with kinship, and which may ultimately confer selective advantages. Although numerous cases of alloparenting have been observed for nursing females, it is not clear whether these are random occurrences due to confusion (as breeding grounds support numerous mother-pup pairs), or whether alloparents that foster their young acquire selective advantages associated with increased inclusive fitness (Riedman, 1982).

Cases of infanticide by adult males may be explained by sexual selection theory, which suggests that by killing unrelated offspring, reproductive adults improve the probability of survival for their own offspring. Signaling the magnitude of concern over infanticide, in an effort to promote juvenile survival, managers have intervened on rare occasions by euthanizing particularly aggressive males. Yet the long-term demographic

155

implications of such actions are unknown. By investigating the influence of genetic relatedness in alloparenting and infanticide, it may be possible to tailor management strategies to optimize survival and minimize mortality.

Sexual selection

Because monk seals mate in the water, it is not readily possible to observe their mating behavior. However, it is generally believed that enables females to choose their mates. Understanding mate choice is particularly important since it directly influences levels of genetic variation within populations. Inbreeding avoidance may be a determining factor; females may prefer to mate with sires that have high reproductive success to ensure that male offspring in particular, or all offspring, will also be reproductively fit (the ‘sexy sons’ and ‘good genes’ hypotheses), or females may prefer mates with specific genetic profiles that confer an adaptive advantage (Clutton-

Brock and McAuliffe, 2009). Development of a pedigree would facilitate detection of such factors by enabling quantification of individual reproductive output.

Emerging opportunities

Additionally, the ever-increasing accessibility and feasibility of high throughput

“omics” technologies offer more holistic opportunities to tackle urgent evolutionary and population genetic questions that directly relate to N. schauinslandi conservation. Whole genome sequencing can expand the capacity to test for the influence of historical evolutionary processes such as drift or selection (Griffith et al., 1989; Leffler et al.,

2012), as well to as identify gene complexes that regulate innate and adaptive immunological responses (Grueber, 2015; Larsen et al., 2014; Longo et al., 2014).

Building on Schultz’s (2011) comments, transcriptomic studies that identify patterns in

156

gene expression could be complemented by proteomic studies that seek to characterize the form and function of protein expression (Horgan and Kenny, 2011), thereby revealing critical elements of resilience within specific or response to rapid environmental change. Genomic characterization of individual microbiomes can facilitate fine-scale understanding of an individual’s interaction with its environment, and can elevate consideration of co-evolutionary relationships between hosts and pathogens or symbionts

(Amato, 2013), as well as yield insights into multi-factorial processes affecting animals’ health and tolerance or susceptibility to disease.

Regardless of scale or technological approach, genetic studies to date have provided managers with important perspectives on the genetic processes underlying observed population demography patterns in species of conservation concern, and will provide a valuable guide for future journeys into the uncharted ‘omics’ of N. schauinslandi.

157

References Amato, K.R., 2013. Co-evolution in context: The importance of studying gut microbiomes in wild animals. Microbiome Science and Medicine 1. doi:10.2478/micsm- 2013-0002

Cadotte, M.W., Davies, J.T., 2010. Rarest of the rare: advances in combining evolutionary distinctiveness and scarcity to inform conservation at biogeographical scales. Diversity and Distributions 16, 376–385. doi:10.1111/j.1472-4642.2010.00650.x

Clutton-Brock, T., McAuliffe, K., 2009. Female mate choice in mammals. The Quarterly Review of Biology 84, 3–27. doi:10.1086/596461

Faith, D.P., 1992. Conservation evaluation and phylogenetic diversity. Biological Conservation 61, 1–10.

Goossens, B., Funk, S.M., Vidal, C., Latour, S., Jamart, A., Ancrenaz, M., Wickings, E.J., Tutin, C.E.G., Bruford, M.W., 2002. Measuring genetic diversity in translocation programmes: principles and application to a chimpanzee release project. Animal Conservation 5, 225–236. doi:10.1017/S1367943002002275

Griffith, B., Scott, J.M., Carpenter, J.W., Reed, C., 1989. Translocation as a species conservation tool: status and strategy. Science 245, 477–480. doi:10.1126/science.245.4917.477

Grueber, C.E., 2015. Comparative genomics for biodiversity conservation. Computational and Structural Biotechnology Journal 13, 370–375. doi:10.1016/j.csbj.2015.05.003

Horgan, R.P., Kenny, L.C., 2011. “Omic” technologies: genomics, transcriptomics, proteomics and metabolomics. The Obstetrician & Gynaecologist 13, 189–195. doi:10.1576/toag.13.3.189.27672

Kruuk, L.E., Hill, W.., 2008. Introduction. Evolutionary dynamics of wild populations: the use of long-term pedigree data. Proceedings of the Royal Society B: Biological Sciences 275, 593–596. doi:10.1098/rspb.2007.1689

Larsen, P.A., Campbell, C.R., Yoder, A.D., 2014. Next-generation approaches to advancing eco-immunogenomic research in critically endangered primates. Mol Ecol Resour 14, 1198–1209. doi:10.1111/1755-0998.12274

Leffler, E.M., Bullaughey, K., Matute, D.R., Meyer, W.K., Ségurel, L., Venkat, A., Andolfatto, P., Przeworski, M., 2012. Revisiting an old riddle: what determines genetic diversity levels within species? PLOS Biol 10, e1001388. doi:10.1371/journal.pbio.1001388

158

Longo, A.V., Burrowes, P.A., Zamudio, K.R., 2014. Genomic studies of disease-outcome in host–pathogen dynamics. Integr. Comp. Biol. 54, 427–438. doi:10.1093/icb/icu073

Moritz, C., Faith, D.P., 1998. Comparative phylogeography and the identification of genetically divergent areas for conservation. Molecular Ecology 7, 419–429. doi:10.1046/j.1365-294x.1998.00317.x

Purvis, A., Agapow, P.-M., Gittleman, J.L., Mace, G.M., 2000. Nonrandom extinction and the loss of evolutionary history. Science 288, 328–330. doi:10.1126/science.288.5464.328

Riedman, M.L., 1982. The evolution of alloparental care and adoption in mammals and birds. The Quarterly Review of Biology 57, 405–435.

Schultz, J.K., 2011. Population genetics of the monk seals (Genus Monachus): a review. Aquatic Mammals 37, 227–235.

Storfer, A., 1999. Gene flow and endangered species translocations: a topic revisited. Biological Conservation 87, 173–180. doi:10.1016/S0006-3207(98)00066-4

Wilson, A.J., Ferguson, M.M., 2002. Molecular pedigree analysis in natural populations of fishes: approaches, applications, and practical considerations. Canadian Journal of Fisheries and Aquatic Sciences 59, 1696–1707. doi:10.1139/f02-127

159

Table 5-1. Common assumptions underlying the use of translocation as a management tool for imperiled populations, and associated questions of relevance to management and genetic research of the Hawaiian monk seal.

ASSUMPTIONS QUESTIONS Can small declining populations be in Migration under population equilibrium equilibrium? What is the time to equilibrium following bottleneck? Are translocated individuals equally as Migrants are demographically equal (and likely to mate successfully (and become by extension, there are no impacts on part of the gene pool) as other their fitness) individuals? Effective population size = population What is the effective population size and size (number of breeding males must the proportion of breeding males and equal number of breeding females; females for the Hawaiian monk seal otherwise a larger number of immigrants across its range, and on individuals’ is required to maintain population size) islands of origin?

160