ISLANDS, ARCHIPELAGOS, AND BEYOND: POPULATION GENETICS AND PHYLOGEOGRAPHY OF HAWAIIAN CORAL REEF

A DISSERTATION SUBMITTED TO THE GRADUATE DIVISION OF THE UNIVERSITY OF HAWAI‘I AT MĀNOA IN PARTIAL FULFILLMENT OF THE REQUIREMENT FOR THE DEGREE OF

DOCTOR OF PHILOSOPHY

IN

ZOOLOGY

AUGUST 2012

By

Derek J. Skillings

Dissertation Committee:

Robert Toonen, Chairperson Brian Bowen Charles Birkeland Andrew Taylor Ronald Bontekoe

i

DEDICATION

This dissertation is dedicated to my wife, Melissa Kay Skillings.

ii

ACKNOWLEDGEMENTS

First, I would like to thank my committee members who provided essential guidance and encouragement throughout my graduate career. Foremost, I would like to thank my advisor and committee chair Rob Toonen. He has generously offered me a near endless supply of advice and guidance, as he does for anyone who knocks on his door. He also gave me the flexibility and encouragement needed to make getting two simultaneous graduate degrees possible. My graduate career has been very unconventional, and Rob has supported me every step of the way. I would like to thank Brian Bowen for giving me the structure I needed to succeed. Given my tendency to get lost in an always increasing number of projects, I would have never finished in a reasonable amount of time without his firm hand at setting deadlines and his enthusiastic encouragement to meet those deadlines. Rob and Brian gave me the perfect balance of freedom and focus that I needed to succeed. I would like to thank Chuck Birkeland for helping me to put my work in the larger perspective of coral reef ecosystems. Chuck also encouraged my philosophical and historical investigations into biology through insightful conversation; every time I saw he seemed to have a valuable and important text that he wanted to give me for my collection, many from his personal library. I would like to thank Andy Taylor for his critical eye, for being able to see through the details to get at the heart of the problem. Andy first taught me statistics, a field that I didn’t appreciate at first, but have since come to focus on greatly because of his influence. Even though we have come to take different approaches to statistics, he has always challenged me in a way that has made my thinking clearer and has made me a better academic. Finally, I would to thank Ron Bontekoe for his dual service as a committee member for my PhD and as my advisor for my philosophy MA. His insightful questions and critical feedback have pushed my writing leaps and bounds ahead of where I thought it could be.

iii

Thank you to the facility, staff, and postdoctoral researchers at the Hawai‘i Institute of

Marine Biology and the University of Hawai‘i at Mānoa Zoology department. I would like to thank: Matt Iacchei, Michelle Gaither, Jon Puritz, Joseph DiBattista, Mike Stat, Zoltan Szabo,

Greg Concepcion, Kim Andrews, Marc Crepeau, Steve Karl and the rest of the extended ToBo lab for helpful discussions and support. I would especially like to thank Chris Bird, who spent many hours teaching me the skills I needed to be a successful molecular biologist. This work was made possible by the financial support of several agencies including: National Science

Foundation, Papahānaumokuākea Marine National Monument, Northwestern Hawaiian Islands

Coral Reef Reserve, University of Hawai‘i Art and Sciences Council, and the Jessie D. Kay

Fellowship.

iv

ABSTRACT

Genetic connectivity determines the evolutionary independence of populations and plays

a primary role governing intraspecies evolution. Patterns of genetic connectivity can also be used

to make inferences about macroevolutionary processes. For my dissertation I used mtDNA

sequences in phylogeographic and population level analyses to examine connectivity within and

between Pacific Archipelagos, utilizing four of tropical echinoderms. First, I review and

prescribe non-lethal sampling techniques for use in genetic studies of many species of marine

invertebrates. Second, I showed that two species of , atra and

Holothuria whitmaei, with a similar life history, species range, and habitat usuage cannot be used

as proxies for each other in order to predict phylogeographic patterns, degree of connectivity, and

population genetic structure within the Hawaiian Archipelago. Third, I examined the genetic population structure of H. atra across the central tropical Pacific to show that despite its large range, H. atra has hierarchical, fine-scale population structure driven primarily by between-

archipelago barriers, but with significant differences between sites within an archipelago. Finally,

I compared population genetic patterns in two congeneric brittle stars, Ophiocoma pica and

Ophiocoma erinaceous, across Hawai‘i and Central and Eastern Polynesia. I found contrasting

phylogeographic patterns in these two similar species, as was the case with H. atra and H.

whitmaei. Given the real-world constraints of limited time and money in marine ecosystem management it would be ideal if model species could stand in as proxies for a host of similar species. This dissertation shows that this ideal scenario is unlikely to be the case; similar life

histories and close phylogenetic relationships do not appear to predict population connectivity.

Generalizations based on a few representative taxa are unlikely to offer much in terms of

delineating boundaries for spatial management areas. Though a more inclusive multi-species

v approach is bound to cost more in terms of time and resources, it should ultimately payout as more informed, if complex, management.

vi

TABLE OF CONTENTS Dedication ………………………………………………………………………………… ii Acknowledgements ………………………………………………………………………... iii Abstract …………………………………………………………………………………... v List of Tables ……………………………………………………………………………... ix List of Figures ……………………………………………………………………………... x Chapter 1: Introduction …………………………………………………………………... 1 References ………………………………………………………………………... 5 Table ……………………………………………………………………………. 9 Chapter 2: It’s Just a Flesh Wound: Non-lethal Sampling for Conservation Genetics Studies……………………………………………………………………...…………... 12 Abstract …………………………………………………………………………... 13 Introduction ………………………………………………………………………... 13 Methods ...………………..………………………………………………………... 15 Discussion ………………………………………………………………………... 20 Conclusions ……………………………………………………………………….. 23 Acknowledgements ……………………………………………………………….. 23 References ………………………………………………………………………... 25 Figure ..…………………………………………………………………………...... 27 Chapter 3: Contrasting Phylogeography of Two Coral Reef Sea Cucumbers, Holothuria atra and Holothuria whitmaei ………………….…………………………………….. 28 Introduction ……………………………………………………………………... 29 Methods ……………………………………………………………………...……. 35 Results ……………………………………………………………………...……… 38 Discussion …………………………………………………………………...…… 42 Conclusions ……………………………………………………………………….. 46 References …………………………………………………………………...…… 47 Tables ……………………………………………………………………...……… 60 Figures …………...……………………………………………………...………… 65 Chapter 4: Gateways to Hawai‘i – Genetic population structure of the tropical sea cucumber Holothuria atra ……………………………………………….…………… 68

vii

Abstract ………………………………………………………...………………… 69 Introduction ………………………………………………………………...……… 69 Methods ……………...……………………………………………………..……. 73 Results ……………………………………………………………………………... 78 Discussion ………………………………………………………………………… 82 Conclusions ……………………………………………………………………….. 88 Acknowledgments ………………………………………………………………… 89 References ………………………………………………………………………... 91 Figures ………………………………………………………...………………… 104 Tables ………………………………………………………………………...…… 106 Appendices ………………………………………………………………………... 110 Chapter 5: Contrasting Phylogeography Patterns in Two Pacific Brittle Stars, Ophiocoma erinaceous and Ophiocoma pica …………..………………………………………...… 113 Introduction ………………………………………………………………...……… 114 Methods ……………...……………………………………………………..……. 119 Data Analysis ……………………………………………………………………... 121 Results ……………………………………………………………………………... 123 Discussion ………………………………………………………………………… 127 Conclusions ……………………………………………………………………….. 131 References ………………………………………………………………………... 132 Tables ………………………………………………………………………...…… 146 Figures ………………………………………………………...………………… 150 Chapter 6: Summary and Synthesis ………………..…………………...………………… 152 Summary ………………………………………………………………...………… 153 Synthesis ………...…...……………………………………………………..……. 155

viii

LIST OF TABLES

Chapter 3: Contrasting Phylogeography of Two Coral Reef Sea Cucumbers, Holothuria atra and Holothuria whitmaei

Table 3.1 Molecular diversity indices and sample sizes ………...... …………… 61

Table 3.2 Pairwise F statistic comparisons ………………………………...…...… 62

Table 3.3 Pairwise Dest_chao comparisons …………………………………………. 63

Table 3.4 AMOVAs by region ……………………………………………………. 64

Table 3.5 Pairwise population migration rate estimates …………………………... 65

Chapter 4: Gateways to Hawai‘i – Genetic population structure of the tropical sea cucumber Holothuria atra

Table 4.1 Molecular diversity indices and sample sizes ………………………… 107

Table 4.2 Pairwise Dest_chao and FST comparisons ……………………….……….. 108

Table 4.3 AMOVAs for different population groupings ……………………….… 109

Table 4.4 Pairwise population migration rate estimates ………………..………… 110

Chapter 5: Contrasting Phylogeography Patterns in Two Pacific Brittle Stars, Ophiocoma erinaceous and Ophiocoma pica

Table 5.1 Molecular diversity indices and sample sizes ………………………… 147

Table 5.2 AMOVAs by region …………………….…………………….……….. 148

Table 5.3 Pairwise Dest_chao and FST comparisons ……...……………………….… 149

Table 5.4 Pairwise population migration rate estimates ………………..………… 150

ix

LIST OF FIGURES

Chapter 2: It’s Just a Flesh Wound: Non-lethal Sampling for Conservation Genetics Studies

Figure 1.1 Non-lethal sampling of a Crown of Thorns sea star (Acanthaster planci) without removing the from the substrate …………………………………… 28

Chapter 3: Contrasting Phylogeography of Two Coral Reef Sea Cucumbers, Holothuria atra and Holothuria whitmaei

Figure 3.1 Map of the Hawaiian Archipelago …..………………………………… 66

Figure 3.2 Haplotype network for Holothuria atra …………….……………….... 67

Figure 3.3 Haplotype network for Holothuria whitmaei …………….…………… 68

Chapter 4: Gateways to Hawai‘i – Genetic population structure of the tropical sea cucumber Holothuria atra

Figure 4.1 Map of Central Pacific with major barriers and migration rates………. 105

Figure 4.2 Haplotype network …………………….………….…………………… 106

Chapter 5: Contrasting Phylogeography Patterns in Two Pacific Brittle Stars, Ophiocoma erinaceous and Ophiocoma pica

Figure 5.1 Haplotype network for Ophiocoma erinaceous ……………………… 151

Figure 5.2 Haplotype network for Ophiocoma pica ………………...…………… 152

x

CHAPTER ONE

Introduction

2

Echinoderms play a major role in structuring many marine ecosystems, and many

are described as "keystone species" because of their profound influence on benthic

community structure (e.g., Paine 1969, Power et al. 1996, Lessios et al. 2001, reviewed

by Uthicke et al. 2009). In addition to their important ecosystem functions, many

species are also the focus of artisanal or commercial fishing efforts,

particularly the sea urchins and sea cucumbers (Sloan 1984, Sala et al. 1998, Purcell

2010). The influence of echinoderm harvest on a wide range of other commercial fisheries, such as abalone, lobster, kelp and kelp-associated fin fish, has long stimulated

discussions of multispecies approaches to managing their exploitation (e.g., Sloan 1984,

reviewed by Purcell 2010). Delineation of the appropriate spatial scales for management

zones within a spatial management network requires a detailed understanding of dispersal

pathways and population connectivity (reviewed by Hedgecock et al. 2007; Thorrold et

al. 2007; Fogarty & Botsford 2007).

Understanding connectivity in the sea is complicated by the fact that many marine organisms share a biphasic life cycle typified by an adult form that is relatively sedentary and a larval form that can potentially disperse across large expanses of open ocean

(Thorson, 1950; Strathmann, 1993; Kinlan & Gaines, 2003; Kinlan et al., 2005; Paulay &

Meyer, 2006). For example, in the sea urchin genus Tripneustes some well-known biogeographic barriers, such the Isthmus of Panama or the long stretch of deep water in the western Atlantic, are important barriers to dispersal whereas others, such as the

Eastern Pacific Barrier show no evidence for limiting dispersal (Lessios et al. 2003).

However, the geographic limits of such dispersal are uncertain because it is virtually impossible with current technology to directly track these microscopic juveniles during

3

the pelagic phase (reviewed by Levin 2006) making indirect methods of quantifying

larval dispersal particularly attractive (reviewed by Grosberg & Cunningham 2001;

Hedgecock et al. 2007; Selkoe et al. 2009; Hellberg 2009). Proxies for dispersal, such as

pelagic larval duration (PLD) and geographic range have generally been used as rules of

thumb in the absence of a detailed understanding of connectivity for most marine species.

Unfortunately, intuitive expectations of larval dispersal potential as a function of PLD

and range size are not upheld in recent meta-analyses of the existing literature (Lester et al. 2007; Bradbury & Bentzen 2007; Bradbury et al. 2008; Weersing & Toonen 2009;

Shanks 2009; Ross et al. 2009). Realized dispersal distance is typically less than potential dispersal distance because of the presence of biophysical or biogeographical barriers (Burton & Feldman, 1981; Knowlton & Keller, 1986; Shanks et al., 2003;

Severance & Karl, 2006; Dawson & Hamner, 2008). Barriers that limit dispersal between marine populations include obvious geographical features such as land masses, i.e. the

Isthmus of Panama (Bermingham & Lessios, 1993), but also more subtle factors such as currents and oceanographic regimes (Dawson, 2001; Barber et al., 2002; Sotka et al.,

2004; Baums et al., 2006; Treml et al., 2008). The correlation between geographic distance and the probability of larval exchange among sites is low in many marine systems (e.g., White et al. 2010), and thus quantitative estimates of connectivity are an important prerequisite for delineating the appropriate scale over which marine populations ought to be managed.

In this dissertation I use genetic and statistical tools to examine marine genetic population connectivity in two echinoderm groups: the sea cucumbers (Holothuroidea)

Holothuria atra and Holothuria whitmaei, and brittle stars (Ophiuroidea) Ophiocoma

4

erinaceous and Ophiocoma pica. The first question I examine is whether congeners with a similar life history, species range, and habitat usuage can be used as proxies for each other in order to predict phylogeographic patterns, degree of connectivity, and population genetic structure. I investigate the usefulness of model species for predicting population structure, population connectivity, and dispersal for use in marine spatial management initiatives for the delineation of marine protected areas and for determining the appropriate scale for coral reef management.

The Hawaiian Archipelago lies at the periphery of the tropical Central Pacific and is the most isolated island chain in the world, making it biogeographically partitioned from the rest of the Pacific Islands (reviewed by Ziegler 2002). This isolation results in one of the highest proportions of endemism in the world (e.g., Briggs 1974; Kay 1980;

Grigg 1983; reviewed by Ziegler 2002; Eldredge & Evenhuis, 2003). Johnston Atoll is believed to be a stepping-stone into Hawai‘i, and simulations of larval dispersal suggest that larvae from Johnston atoll can reach French Frigate Shoals or Kaua‘i along two separate larval corridors (Kobayashi 2006; Kobayashi & Polovina 2006). The second question I examine is whether Johnston Atoll is a stepping-stone into Hawai‘i, whether it tends to be an outpost of Hawai‘i. I also elucidate the historical connectivity pathways between the Hawaiian Archipelago and Johnston Atoll.

Though there are many examples of pan-pacific coral reef organisms in Hawai‘i, the isolation of the Hawaiian Archipelago is thought to limit larval exchange sufficiently that colonization is rare (Jokiel & Martinelli 1992). For example, Kay (1984) estimated that Western Pacific marine species successfully colonize the Hawaiian Archipelago about once every 13,000 years. Unlike the terrestrial fauna, however, the Hawaiian

5

marine fauna contains a large proportion of endemics that are differentiated but not

diversified from their Indo-West Pacific roots (Kay & Palumbi 1987; Hourigan & Reese

1987; Jokiel 1987; Ziegler 2002). In this scenario Hawai‘i is seen primarily as a dead-

end, an isolated land-mass that does not contribute in a significant way to the overall

diversity of the tropical pacific. Counter to the island biogeography hypotheses of

Hawaiian diversity, Jokiel and Martinelli (1992) proposed the Vortex model of speciation,

wherein the stunning biodiversity of the Coral Triangle is a result of centrifugal

accumulation of species from the peripheral habitats around the Pacific. Though these

two models primarily make predictions about speciation-level processes and do not speak

directly to gene-flow within a species, they do make opposite claims about the dominant

direction of gene-flow and dispersal. The final questions I examine are whether periphery

archipelagos act as a source of genetic diversity in the Pacific, and the likely colonization

routes, into and out of, the extremely isolated Hawaiian Archipelago.

REFERENCES

Barber PH, Palumbi SR, Erdmann MV, Moosa MK (2002) Sharp genetic breaks among

populations of Haptosquilla pulchella (Stomatopoda) indicate limits larval transport:

patterns, causes, and consequences. Molecular Ecology. 659-674.

Baums IB, Paris CB, Cherubin LM (2006) A bio-oceanographic filter to larval dispersal in a reef-building coral. Limnology and Oceanography. 51:1969–1981.

Bermingham E, and Lessios H (1993) Rate variation of protein and mtDNA evolution as revealed by sea urchins separated by the Isthmus of Panama. Proceedings of the National.

Academy of Science. 90: 2734-2738

6

Bradbury IR, Bentzen P (2007) Non-linear genetic isolation by distance: implications for dispersal estimation in anadromous and marine fish populations. Marine Ecology

Progress Series. 340:245-257.

Bradbury IR, Laurel B, Snelgrove PVR, Bentzen P, Campana SE (2008) Global patterns in marine dispersal estimates: the influence of geography, taxonomic category and life history. Proceedings of the Royal Society London B 275:1803-1810.

Briggs JC (1974) Marine Zoogeography. McGraw Hill Publishers, New York, NY.

Burton R, Feldman M (1981) Population genetics of coastal and estuarine invertebrates: are behavioral differences among species reflected in population structure? Estuaries.

4:239.

Dawson MN (2001) Phylogeography in coastal marine : solution from California?

Journal of Biogeography. 28:723-736.

Dawson MN, Hamner WM (2008) A biophysical perspective on dispersal and the geography of evolution in marine and terrestrial systems. Journal of the Royal Society

Interface. 5:135-150.

Eldredge LG, Evenhuis NL (2003) Hawai‘i’s biodiversity: a detailed assessment of the numbers of species in the Hawaiian Islands. Bishop Museum Occ Pap. 76:1-28.

Fogarty MJ, Botsford LW (2007) Population connectivity and spatial management of marine fisheries. Oceanography. 20(3):112-123.

7

Grigg RW (1983) Community structure, succession and development of coral reefs in

Hawai‘i. Marine Ecology Progress Series 11:1-14.

Grosberg RK, Cunningham CW (2001) Genetic Structure in the Sea: From Populations to

Communites. In: Marine Community Ecology. Bertness MD, Gaines SD, Hay ME (Eds.).

Sinauer Associates, Inc., Sunderland, MA. 61-84.

Hedgecock D, Barber PH, Edmands S (2007) Genetic approaches to measuring connectivity. Oceanography. 20(3):70-79.

Hellberg ME (2009) Gene flow and isolation among populations of marine animals.

Annual Review of Ecology, Evolution, and Systematics. 40:291-310.

Hourigan TF, Reese ES (1987) Mid-ocean isolation and evolution of Hawaiian reef fishes. Trends in Ecology and Evolution. 2:187-191.

Jokiel PL (1987) Ecology, biogeography and evolution of corals in Hawai;i. Trends in

Ecology & Evolution. 2:179-182.

Jokiel PL, Martinelli FJ (1992) The Vortex Model of Coral Reef Biogeography. Journal of Biogeography, 19(4):449-458.

Kay EA (1980) Little worlds of the Pacific. An essay on Pacific Basin Biogeography.

Harold L. Lyon Arboretum, University of Hawai‘i, Lecture No. 9:1-40.

Kay EA (1984) Patterns of speciation in the Indo-West Pacific. In: Radovsky FJ, Raven

PH, Sohmer SH (Eds.). Biogeography of the Tropical Pacific. Assoc. of Systematics

Collections and B.P. Bishop Museum. 15-31.

8

Kay EA, Palumbi SR (1987) Endemism and evolution in Hawaiian marine invertebrates.

Trends in Ecology and Evolution. 2(7):183-186.

Kinlan BP, Gaines SD (2003) Propagule dispersal in marine and terrestrial environments: a community perspective. Ecology. 84:2007-2020.

Kinlan BP, Gaines SD , Lester SE (2005) Propagule dispersal and the scales of marine community process. Diversity & Distribution. 11:139–148

Knowlton N, Keller B (1986) Larvae which fall short of their potential: highly localized recruitment in an alpheid shrimp with extended larval development. Bulletin of Marine

Science. 39:213-223.

Kobayashi DR (2006) Colonization of the Hawaiian Archipelago via Johnston Atoll: A characterization of oceanographic transport corridors for pelagic larvae using computer simulation. Coral Reefs. 25:407–417.

Kobayashi DR, Polovina JJ (2006) Simulated seasonal and interannual variability in larval transport and oceanography in the Northwestern Hawaiian Islands using satellite remotely sensed data and computer modeling. Atoll Research Bulletin. 543:365–390.

Lessios HA, Garrido M, Kessing BD (2001) Demographic history of Diadema antillarum, a keystone herbivore on Caribbean reefs. Proceedings of the Royal Society.

268:1-7.

Lessios HA, Kane J, Robertson DR (2003) Phylogeography of the pantropical sea urchin

Tripneustes: contrasting patterns of population structure between oceans. Evolution.

57(9):2026-2036.

9

Lester SE, Ruttenberg BI, Gaines SD, Kinlan BP (2007) The relationship between

dispersal ability and geographic range size. Ecology Letters. 10:745-758.

Levin L (2006) Recent progress in understanding larval dispersal: new directions and

digressions. Integrative and Comparative Biology. 46(3):282-297.

Paine RT (1969) A note on trophic complexity and community stability. American

Naturalist 103: 91-93.

Paulay G, Meyer C (2006) Dispersal and divergence across the greatest ocean region: Do larvae matter? Integrative and Comparative Biology. 46(3):269-281.

Power ME, Tilman D, Estes JA, Menge BA, Bond WJ, Mills LS, Daily G, Castilla JC,

Lubchenco J, Paine RT (1996) Challenges in the Quest for Keystones. BioScience. 46:

609-620.

Purcell SW (2010) Managing sea cucumber fisheries with an ecosystem approach. In:

FAO Fisheries and Aquaculture Technical Paper No. 520. Lovatelli A, Vasconcellos M,

Yimin Y (Eds.). Rome, FAO. 157 pp.

Ross PM, Hogg ID, Pilditch CA, Lundquist CJ (2009) Phylogeography of New

Zealand’s coastal benthos. New Zealand Journal Marine and Freshwater Research. 43:

1009-1027.

Sala E, Boudouresque CF, Harmelin-Vivien M (1998) Fishing, Trophic Cascades, and the

Structure of Algal Assemblages: Evaluation of an Old but Untested Paradigm. Oikos. 82:

425-43.

10

Selkoe KA, Henzler C, Gaines SD (2009) Seascape genetics and the spatial ecology of marine populations. Fish and Fisheries 9:363-377.

Severance E, Karl S. Contrasting population genetic structures of sympatric, mass- spawning Caribbean corals. Marine Biology. 150:57-68.

Shanks A, Grantham B, Carr M (2003) Propagule dispersal distance and the size and spacing of marine reserves. Ecological Applications. 13:S159-S169.

Shanks AL (2009) Pelagic Larval Duration and Dispersal Distance Revisited.

Biolological Bulletin. 216:373-385.

Sloan, SA (1984) Echinoderm fisheries of the world: A review. In: Echinodermata:

Proceedings of the Fifth International Echinoderm Conference. Keegan BF, O’Connor

BDS (Eds.). A.A. Balkema, Rotterdam, Netherlands. 109-124.

Sotka EE, Wares JP, Barth JA, Grosberg RK, Palumbi SR (2004) Strong genetic clines and geographical variation in gene flow in the rocky intertidal barnacle Belanus glandula.

Molecular Ecology. 13:2143-2156.

Strathmann RR (1993) Hypotheses on the origins of marine larvae. Annual Review of

Ecology and Systematics. 24:89-117.

Thorrold SR, Zacherl DC, Levin LA (2007) Population connectivity and larval dispersal using geochemical signatures in calcified structures. Oceanography. 20(3):80-89.

Thorson G (1950) Reproductive and larval ecology of marine bottom invertebrates.

Biological Review of the Cambridge Philosophical Society. 25:1-45.

11

Treml EA, Halpin PN, Urban DL, Pratson LF (2008) Modeling population connectivity

by ocean currents, a graph-theoretic approach for marine conservation. Landscape

Ecology. 23:19-36.

Uthicke S, Schaffelke B, Byrne M (2009) A boom-bust phylum? Ecological and

evolutionary consequences of density variations in echonoderms. Ecological

Monographs. 79:3-24.

Weersing KA, Toonen RJ (2009) Population genetics, larval dispersal, and demographic

connectivity in marine systems. Marine Ecology Progress Series 393:1-12 .

White C, Selkoe KA, Watson J, Siegel DA, Zacherl DC, Toonen RJ (2010) Ocean

currents help explain population genetic structure. Proceedings of the Royal Society B:

Biological Science. 277: 1685-1694.

Ziegler AC (2002) Hawaiian natural history, ecology & evolution. University of Hawai‘i

Press, Honolulu HI, USA. 477 p.

12

CHAPTER TWO

It’s Just a Flesh Wound: Non-lethal Sampling for Conservation Genetics Studies

Published as:

Skillings DJ and Toonen RJ (2010) It’s Just a Flesh Wound: Non-lethal Sampling for Conservation Genetics Studies. in Diving for Science. Proceedings of the 29th American Academy of Underwater Sciences Symposium, N. W. Pollock, Ed., Proceedings of the American Academy of Underwater Sciences, AAUS, Dauphin Island, Ala, USA.

13

ABSTRACT

The Papahānaumokuākea Marine National Monument (PMNM) is a chain

of coral reef atolls and small basalt pinnacles, home to the most beautiful

reefs in Hawai‘i. The size, remoteness, and near “pristine” condition of

PMNM make it a unique and ideal natural laboratory for conservation and

scientific studies. Because PMNM enjoys the highest level of protection

under US law, every effort is made to minimize impacts when performing

research and collecting samples within the borders of the Monument. An

overview of underwater, non-lethal, tissue-sampling techniques is given

for a wide range of invertebrates used for genetic population studies.

Many invertebrates pose a special challenge for non-lethal sample

collecting as defensive chemical and physical structures can make

downstream genetic procedures difficult or impossible if they are not

sampled correctly.

INTRODUCTION

The Hawaiian archipelago stretches more than 2500 km, from the Big Island of

Hawai‘i at its southeast point, to Kure Atoll in the northwest. The Archipelago is commonly divided into two regions: the main Hawaiian Islands (MHI) consisting of populated, high volcanic islands and the Northwestern Hawaiian Islands (NWHI), an uninhabited chain of basalt pinnacles, atolls, shoals, and banks. Although the inhabited

MHI are about on par with reefs in other inhabited areas, partly because of their isolation, the coral reefs of the NWHI are among the healthiest and most extensive remaining in the

14 world (Pandolfi et al. 2003). The reefs of the NWHI represent a nearly undamaged coral reef ecosystems with virtually no human impacts by comparison to the heavily populated

MHI (Selkoe et al. 2009), abundant and large apex predators, and an extremely high proportion of endemic species across many taxa (DeMartini & Friedlander 2004;

Friedlander & DeMartini 2002), . The coral reefs in the MHI, by comparison, are under considerable anthropogenic pressure, reefs near many of the urban areas and popular tourist destinations show significant negative impacts (Friedlander et al. 2004; Rodgers et al. 2009).

The Papahānaumokuākea Marine National Monument (PMNM) was created to protect the nearly 140,000 square miles of the NWHI; home to more than 7,000 species of fishes, invertebrates, marine mammals, and birds. While the full extent of its biodiversity is still unknown, an average of about 27% of the known species are endemics found nowhere else on Earth (Eldredge & Evenhuis 2003). The diverse habitats found in the Papahānaumokuākea Marine National Monument present both opportunity and duty. Its size, remoteness, and near pristine condition make it a unique and ideal natural laboratory for conservation and scientific studies. As one of the last bastions of minimally impacted coral reefs we have a duty to preserve the NWHI. In order to inform ecosystem-based management of the Papahānaumokuākea Marine National Monument, understand the role of NWHI in the management of the MHI, and evaluate the potential for spillover of fisheries species from the protected area, our group has embarked on a genetic survey of reef fishes, corals, and other invertebrates. This survey is designed to address the issue of population connectivity between the Northwestern and Main

15

Hawaiian Islands. An extended overview of this effort and preliminary research results

can be found in Toonen et al. (2010).

In keeping with conservation ethic every effort is made to minimize impacts when

performing field research and collecting samples for laboratory analysis. This paper is an

overview of underwater, non-lethal, tissue-sampling techniques for a wide range of invertebrates used by our research group for population genetics studies. Many invertebrates pose a special challenge for non-lethal sample collecting; defensive chemicals and physical structures can make downstream procedures difficult or impossible if an appropriate sampling methodology is not followed. We present here common downstream problems and methods for navigating the line between safe, non- lethal sampling and the effective and efficient use of collecting time.

METHODS

Sampling preparation for genetic population studies starts well before getting into the field. The end needs for the study in question must be kept in mind when deciding on the appropriate sampling methodology. This is especially important for our research group because of the isolation of the NWHI. Following a roughly 6 month permit application process, any field expedition is expensive and extremely time-constrained with typically only 1 day to perform collections, and at best 4 days possible at any given site. Thus, in addition to the ethical responsibility to avoid wasting anything collected, it is also important to know exactly what is needed for the study because there may not be a second chance for getting samples. Final sampling methodology is dictated by the following questions, each addressed in turn:

16

What laboratory procedures will be used?

Lab methodologies for population genetics studies utilize genetic marker variability, or infer genetic variability through fragment length or electrophoretic differences, in order to understand population structure. Genetic variability is generally determined using either direct sequencing of DNA, or size-based scoring of the allele frequencies of genes or proteins. Our research group uses both direct sequencing of mitochondrial (mtDNA) and nuclear genes (nDNA), and the scoring of length variability within nDNA microsatellite markers. The required sample size for using the different types of markers is dependent on the variability both between, and within the chosen markers, and is beyond the scope of this paper (reviewed by Ruzzante 1998; Kalinowski

2005).

Sampling considerations come from three laboratory processing stages necessary for obtaining genetic information: 1) the preservation of the DNA molecules within the whole tissue; 2) the clean extraction in order to isolate molecules of interest; and 3) the amplification of target genetic material. Details for these procedures can be found in

Iacchei and Toonen (2010). The quality of the final product depends largely on type of tissue that is used in the initial extraction. Most extraction procedures are optimized for vertebrate skin, muscle or blood. Invertebrate muscle or epithelial tissue is likewise usually suitable for DNA extraction, depending on the organism in question. Protein extraction may call for specific tissues that have a higher concentration of the protein of interest, and requires fresh tissue or specialized storage conditions (-80ºC).

17

How much tissue is needed?

Most DNA extraction protocols call for only a few milligrams of tissue per

extraction. Generally the smaller the amount of tissue that is biopsied the greater the

chances are for a quick recovery for the organism, but this is also a trade-off in that very tiny biopsies do not leave a tissue backup for re-extraction if the first sample is lost or contaminated. Thus, it is almost always a good idea to take moretissue than is needed for a single laboratory procedure. A biopsy of 50-500 milligrams of tissue generally allows a reserve to be stored in the case of failed extractions, lost product, multiple analyses, or for an independent extraction should the results need secondary confirmation.

Small tissue biopsies can also make it very difficult or impossible to correctly match a tissue sample to the species from which it was taken at any point down-stream.

This easily overlooked consequence of tissue biopsy sampling can completely derail a research project. Larger tissue samples or samples with important identifying features can help for later identification. Regardless, good field photos, notes, and proper labeling of all tissue are essential and can mitigate this risk almost entirely.

A greater amount of tissue is needed for protein extraction than is needed for

DNA extraction; the total amount of tissue needed will depend on the concentration of the protein in the tissue. Preliminary preservation and extraction tests should be run prior to field expeditions, especially for marine invertebrates which frequently show taxon- specific difficulties for preservation, extraction or DNA amplification (reviewed by

Dawson et al. 1998; Gaither et al. in review). Other considerations include whether the tissue will be shared across multiple labs, if multiple types of analyses will done using the

18

tissue and if so, how much tissue will be needed for each analysis. Any of these factors

will increase the amount of tissue that will be needed.

What type of tissue is needed?

The type of tissue needed will depend on both the desired end product and the

organism in question. The tissue needed for protein extraction will depend on the type of

protein that is targeted. If DNA is the target then cellular tissue is required generally in

the form of epithelial, muscular, visceral, or reproductive tissue. Structural, skeletal, and

connective tissues rarely contain enough DNA or protein to work with easily. Although it

has been done successfully with special effort, coral skeletons, crustacean

exoskeletons/molts, molluscan shells, and echinoderm tests or spines are not ideal

samples for subsequent genetic analyses.

Many invertebrates also have defensive chemicals that can interfere with the

amplification or sequencing of DNA. These chemicals can be concentrated in tissues or

diffuse throughout the body. The tissue selected for biopsy will then depend on whether

the tissue contains compounds that interfere with downstream procedures and whether

the interference can be removed or mitigated.

Can the study organism be non-lethally sampled?

The amount and type of tissue needed for molecular procedures, taken together

with the idiosyncrasies of the target organism, will determine if non-lethal sampling is

viable. Tiny organisms are more difficult to sample non-lethally as the amount of tissue that is needed might be immediately lethal or significantly decrease the survival rate post- biopsy. Molting, regenerating, or colonial organisms are prime candidates for non-lethal

19

sampling because they can usually recover quickly from biopsy injuries. Segmented

organisms might also be good candidate if there are replicates of the body part chosen for

sampling or the natural defense strategy of the organism includes autotomy (self

amputation). Finally, the ability to successfully capture, biopsy, and safely release the

organism in a reasonable timeframe will also determine if non-lethal sampling is possible. For example, this last factor is one of the key reasons why small reef fishes are difficult to sample non-lethally.

How will the tissue be preserved?

After the size and location of tissue biopsy is decided, the method of tissue

preservation must be determined. Two of the more common preservatives for genetic

studies are dimethyl sulfoxide (DMSO) salt-saturated buffer (Seutin et al. 1991) and

>70% ethanol (EtOH). Although specimen collection is time consuming and expensive, few laboratories test preservation methods for tissues before setting out on field expeditions. Particularly for marine invertebrates, there are substantial differences among taxa in the efficacy of tissue preservation among methods, and species-specific

differences indicate that preservation comparisons should be undertaken for any long-

term storage of samples destined for PCR study (Dawson et al. 1998; Gaither et al. in

review). In general, DMSO is non-flammable and makes a good preservative for samples

that require shipping or denser tissue; soft tissues can be broken down into a thick,

mucous-like texture, making tissue handling difficult. Ethanol dehydrates tissue, often

hardening it in the process, making it good for soft tissues that must be manipulated

downstream. Ethanol is flammable, however, so samples preserved in EtOH require

special storage, and shipping samples stored in ethanol can be difficult. Protein analyses

20

usually require deep-frozen (-80ºC) tissue that has not been chemically preserved; this

fact alone makes studies focusing on proteins impractical in many tropical locations,

especially when sampling takes place in remote locations.

DISCUSSION

Marine organisms comprise an extremely diverse category for study, and each

new group presents new challenges. Our research group has experience with about 1/3 of

marine phyla, but this represents only small fractionof the multitude of marine species.

Despite this we have found some general rules-of-thumb that have been helpful when starting work on new organisms.

Sponges and Corals

Sponges and corals are colonial organisms that quickly regenerate damaged tissue and are straightforward to sample non-lethally. Sponges are totipotent and contain equivalent tissue throughout whereas corals are usually comprised of thin tissue covering a dense skeletal structure unsuitable for genetic study. Care must be taken so that the colony attachment point to the substrate is not damaged. Most of the downstream problems involving sponges come from skeletal elements and the myriad array of defensive chemicals that interfere with DNA extraction and amplification. Additionally, we have found that EtOH preservation of both corals and sponges yields poor quality

DNA relative to DMSO (Gaither et al. in review). Troubleshooting generally involves extended cycles of cleaning the DNA or diluting the extraction solution to a point where inhibition drops off but there is still enough template DNA for amplification.

21

Downstream problems for corals include extreme amounts of mucous production and similar interference from defensive chemicals or metabolites (Concepcion et al. 2006).

Crustaceans

All crustaceans are segmented with articulated joints, and most use autotomy as a defensive mechanism of self-amputation to escape predators. Likewise, most shed their exoskeletons multiple times throughout their lifetimes, and have the ability to completely regenerate lost limbs. Many small crustaceans are difficult to identify to species in the field, and are too small to effectively sample non-lethally. Larger benthic crustaceans such as lobsters, crabs, and shrimp can be often be hand-caught, and many will drop a limb for distraction in the effort to escape capture. This makes sample-collecting easier because the whole organism need not be captured. Even for those who do not autotomize a limb, once caught, a leg joint can be clipped and will often regenerate within a single molt cycle. The sampling goal is obtain muscular tissue and not the strong, chitinous exoskeleton. We often come across empty exoskeletons, or molts, that have been shed in the growing process. Though DNA can occasionally be salvaged from these molts it takes special effort, and is not a preferred sampling strategy. Muscle tissue from a live specimen is the ideal tissue sample.

Echinoderms

Echinoderms are a very diverse group with both physical and chemical defenses to be overcome in the post-sampling analyses. Most echinoderms are capable of extensive regeneration. In the case of urchins and sea cucumbers, most regenerate quickly so long as the body cavity is not exposed and damaged, and in the case of brittle stars and

22

sea stars, most regenerate quickly so long as the oral disk is left intact. Many sea cumbers

also have defensive chemicals in their skin that interfere with DNA extraction and/or

amplification so it is important to get muscle tissue from the body wall when biopsying.

Care must be taken to not puncture the muscular body wall completely, which will

expose the viscera and frequently result in death via infection or as a result.

Whole arms or individual tube feet can be taken in the case of sea stars or brittle stars

(Figure 1). Urchins can be difficult to sample non-lethally, because most of the exposed portion of an urchin is skeletal material which does not contain much tissue. One strategy for urchins that can be picked up is to place them on a hard, clean surface until they attach themselves with their tube feet. They can then be quickly pulled away from the surface; this usually leaves behind a few tube feet which can be collected for subsequent

DNA analyses. Alternatively, urchins with large spines will have muscle tissue and skin at the base attachment point which can be taken after pulling or twisting out spines. It is important to not puncture the test while removing spines or tube feet; this will usually kill the urchin because in our experience damaged urchins on a reef usually suffer predation before the diver reaches the surface.

Molluscs

By comparison to the groups above, most molluscs are fleshy and either slowly regenerating or non-regenerating. Biopsies can create permanent damage but if care is taken the damage can be repaired or very minor. For example, seemingly healthy individuals are seen in the field with old wounds from predators that are far more extreme than damage done by tissue biopsy. Thus small, non-lethal tissue sampling is possible,

and pieces of mantle, foot, or in the case of cephalopods, arm tips are preferred.

23

CONCLUSIONS

Studies of genetic population connectivity are important for the successful

management and conservation of coral reefs (see Toonen et al. 2010). When performing a

study aimed at conservation and management, we believe every effort should be made to

minimize impacts on the natural resource we are studying. In the case of molecular

genetic studies, the tiny tissue sample needed for research makes non-lethal sampling possible for many species. In order for any study to be successful, it is important to plan sampling methodology and preservation method ahead of time, especially when considering the expense associated with field research across a large and remote geographical area. In the case of remote locations such as the National Monuments in the

Pacific, field expeditions are often one-time opportunities within the career of a graduate student. Thus, knowing the physical and chemical characteristics of target organisms is important to minimize the impact of collections and the chance of insurmountable problems with downstream laboratory procedures while simultaneously maximizing the scientific return on the time and financial investment in the research.

ACKNOWLEDGEMENTS

We thank the Papahānaumokuākea Marine National Monument, US Fish and

Wildlife Services, and Hawai‘i Division of Aquatic Resources (DAR) for coordinating research activities and permitting, and the National Oceanic and Atmospheric

Administration (NOAA) research vessel Hi‘ialakai and her crew for years of outstanding service and support. Special thanks go to the members of the ToBo Lab, UH Dive

Program, NMFS, PIFSC, CRED, A. Wilhelm, R. Kosaki, H. Johnson, M. Pai, D. Carter,

24

C. Kane, S. Karl, C. Meyer, S. Godwin, D. Minton, P. Reath, J. Zardus, D. Croswell, K.

Holland, M. Stat, X. Pochon, R. Gates, M. Rivera, E. Brown, M. Ramsay, J. Maragos, B.

Walsh, B. Carmen, I. Williams, S. Cotton, T. Montgomery, S. Pooley, M. Seki, E.

DeMartini, J. Polovina, B. Humphreys, D. Kobayashi, F. Parrish, B. Moffit, G. DiNardo,

J. O’Malley, R. Brainard, M. Timmers, J. Kenyon, S. Daley, M. Crepeau, K. Schultz, M.

Duarte, H. Kawelo, T. Daly-Engel, L. Sorenson, L. Basch, A. Alexander, M. Craig, L.

Rocha, C. Musberger, D. White, M. Gaither, G. Conception, Y. Papastamatiou, M.

Crepeau, Z. Szabo, and the HIMB EPSCoR Core genetics Facility.

This work was funded in part by grants from the National Science Foundation

(DEB#99-75287, OCE#04-54873, OCE#06-23678, OCE#09-29031), National Marine

Sanctuaries NWHICRER-HIMB partnership (MOA-2005-008/6882), University of

Hawai‘i Sea Grant College Program, National Park Service PICRP, National Marine

Fisheries Service, Western Pacific Regional Fishery Management Council, NOAA's

Coral Reef Conservation Program, the Hawai‘i Coral Reef Initiative, NSF EPSCoR, EPA

STAR Fellowship, the Watson T. Yoshimoto Foundation, the Jessie D. Kay Memorial

Fellowship, PADI Foundation Research Grant, Charles and Margaret Edmondson

Research Fund, American Malacological Society Student Award, Conchologists of

America Research Grant, Sigma Xi Grants-in-Aid, Society for Integrative and

Comparative Biology Student Award, Western Society of Malacologists Student Award, and the Ecology, Evolution, and Conservation Biology (EECB) NSF GK-12 fellowships.

25

REFERENCES

Concepcion G, Medina M, Toonen RJ (2006) Non-coding mitochondrial loci for corals.

Molecular Ecology Notes. 6:1208–1211.

Dawson MN, Raskoff KA, Jacobs DK (1998) Field preservation of marine invertebrate tissue for DNA analyses. Molecular Marine Biology and Biotechnology. 7; 145-152

DeMartini EE, Friedlander AM (2004)Spatial pattern of endemism in shallow water reef

fish populations of the Northwestern Hawaiian Island. Marine Ecology Progress Series.

271: 281-296.

Eldredge LG, Evenhuis NL (2003) Hawai‘i’s biodiversity: a detailed assessment of the numbers of species in the Hawaiian Islands. Bishop Museum Occasional Papers. 76:1–

28. Available at http://hbs.bishopmuseum.org/pdf/op76.pdf

Friedlander AM, DeMartini EE (2002) Contrasts in density, size and biomass of reef

fishes between the Northwestern and Main Hawaiian Islands: the effects of fishing down apex predators. Marine Ecology Progress Series. 230:253-264.

Friedlander AM, Kobayashi D, Bowen BW, Meyers C, Papastamatiou Y, DeMartini EE,

Parrish F, Treml E, Currin C, Hilting A, Weiss J, Kelley C, O’Conner R, Parke M, Clark

R, Toonen R, Wedding L. (2009) Chapter 9. Connectivity and Integrated Ecosystem

Studies. In: Papahānaumokuākea Monument Biogeographic Assessment Report. Pp. 291-

330. NOAA Technical Memorandum.

26

Gaither M, Bowen BW, Toonen RJ, Planes S, Messmer V, Earle J, Robertson DR (2010)

Genetic consequences of introducing two divergent, allopatric lineages of Bluestripe

Snapper (Lutjanus kasmira) to Hawai‘i. Molecular Ecology. 19:1107-1121.

Gaither M, Szabó Z, Crepeau M, Bird CE, Toonen RJ (In review) Preservation of corals

in salt-saturated DMSO buffer is superior to ethanol for PCR experiments. Coral Reefs.

Kalinowski ST (2005) Do polymorphic loci require large sample sizes to estimate genetic

distances? Heredity. 94; 33–36.

Levin LA (2006) Recent progress in understanding larval dispersal: new directions and

digressions. Integrative and Comparative Biology. 46:282-297.

Pandolfi JM, Bradbury RH, Sala E, Hughes TP, Bjorndal KA, Cooke RG, McArdle D,

McClenachan L, Newman MJH, Paredes G, Warner RR, Jackson JBC (2003) Global

trajectories of the long-term decline of coral reef ecosystems. Science. 301: 955–958.

Rivera MA, Kelley CD, Roderick GK (2004) Subtle population genetic structure in the

Hawaiian grouper, Epinephelus quernus (Serranidae) as revealed by mitochondrial DNA

analyses. Biological Journal of the Linnean Society. 81: 449-468.

Rodgers K, Jokiel PL, Bird C, Brown EK (2009) Quantifying the condition of Hawaiian

coral reefs. Aquatic Conservation: Marine and Freshwater Ecosystems. 20: 93–105.

Ruzzante DE (1998) A comparison of several measures of genetic distance and

population structure with microsatellite data: bias and sampling variance. Canadian

Journal of Fish and Aquatic Science. 55; 1−14.

27

Selkoe KA, Halpern BH, Ebert C, Franklin E, Selig E, Casey K, Bruno J, Toonen RJ

(2009) A map of cumulative impacts to a “pristine” coral reef ecosystem, the

Papahānaumokuākea Marine National Monument. Coral Reefs. 28:635-650.

Seutin G, White BN, Boag PT (1991) Preservation of avian blood and tissue samples for

DNA analyses. Canadian Journal of Zoology. 69:82-90.

Figure 1.1 Non-lethal sampling of a Crown of Thorns sea star (Acanthaster planci) without removing the animal from the substrate. Roughly 1cm is cut off the tip of an arm

in situ to minimize impact to the animal.

28

CHAPTER THREE

Contrasting Phylogeography of Two Coral Reef Sea Cucumbers, Holothuria atra and

Holothuria whitmaei

29

INTRODUCTION

Coral reef ecosystems are in peril, some estimates predict that as much as 70% of

the world’s reefs will be gone by 2050 (Wilkinson 2000) or undergo a major shift from

coral-dominated to algal-dominated environments (Pandolfi et al. 2005). The primary

threats to marine ecosystems include overfishing, destructive fishing, pollution, coastal

development and climate change (McCleod and Leslie 2009). In order to help protect

community-wide coral reef biodiversity and the associated ecosystem services it provides

against this host of threats, marine protected areas (MPAs) have been, and are continuing

to be, established worldwide (McCleod and Leslie 2009). Successful spatial management

requires a complex system of zones, each of which seeks to match exploitation to

differences in biological productivity, population levels and socio-economic payoffs

(Sanchirico & Wilen 2005). Delineation of the appropriate spatial scales for management

zones within a spatial management network requires a detailed understanding of dispersal

pathways and population connectivity (reviewed by Hedgecock et al. 2007; Thorrold et

al. 2007; Fogarty & Botsford 2007).

Understanding connectivity in the sea is complicated by the fact that many marine organisms share a biphasic life cycle in which benthic or sedentary adults reproduce through an oceanic phase (eggs and/or larvae), which may be pelagic for as little as a few minutes to in excess of a year depending on the species. Following the pelagic development phase, these larvae settle onto a reef where they may remain through their entire lives, and in cases of sessile organisms, such as corals, the act of settlement includes permanent attachment to a single site. Thus, long distance dispersal is accomplished almost exclusively during the pelagic larval phase, which can potentially

30

disperse across large expanses of open ocean (Thorson, 1950; Strathmann, 1993; Kinlan

& Gaines, 2003; Kinlan et al., 2005; Lessios & Robertson 2006; Paulay & Meyer, 2006;

Dawson & Hamner 2008). For example, in the sea urchin genus Tripneustes some well-

known biogeographic barriers, such the Isthmus of Panama or the long stretch of deep

water in the western Atlantic, are important barriers to dispersal whereas others, such as

the Eastern Pacific Barrier show no evidence for limiting dispersal (Lessios et al. 2003).

However, the geographic limits of such dispersal are uncertain because it is virtually

impossible to track these microscopic juveniles during the pelagic phase (reviewed by

Levin 2006) making indirect methods of quantifying larval dispersal particularly

attractive (reviewed by Grosberg & Cunningham 2001; Hedgecock et al. 2007; Selkoe et

al. 2009; Hellberg 2009).

Proxies for dispersal, such as pelagic larval duration (PLD) and geographic range

have generally been used as rules of thumb in the absence of a detailed understanding of

connectivity for most marine species. However, intuitive expectations of larval dispersal

potential as a function of PLD are not upheld in detailed meta-analyses of the existing literature (Lester et al. 2007; Bradbury & Bentzen 2007; Bradbury et al. 2008; Weersing

& Toonen 2009; Shanks 2009; Ross et al. 2009). Realized dispersal distance is typically less than potential dispersal distance because of the presence of biophysical or biogeographical barriers (Burton & Feldman, 1981; Knowlton & Keller, 1986; Shanks et al., 2003; Severance & Karl, 2006; Dawson & Hamner, 2008). Barriers that limit dispersal between marine populations include obvious geographical features such as land masses, i.e. the Isthmus of Panama (Bermingham & Lessios, 1993), but also more subtle factors such as currents and oceanographic regimes (Dawson, 2001; Barber et al., 2002;

31

Sotka et al., 2004; Baums et al., 2006; Treml et al., 2008). The correlation between geographic distance and the probability of larval exchange among sites is low in many marine systems (e.g., White et al. 2010), and thus quantitative estimates of connectivity are an important prerequisite for delineating the appropriate scale over which marine populations ought to be managed.

The Hawaiian archipelago stretches more than 2500 km in length, and consists of two regions: the main Hawaiian Islands (MHI) which are populated, high volcanic islands; and the Northwestern Hawaiian Islands (NWHI) which are an uninhabited string of low islands, atolls, shoals, and banks. Thanks to their isolation, the roughly 4,500 square miles of wild coral reefs found throughout the NWHI are among the healthiest and most extensive remaining in the world (Pandolfi et al. 2003). The reefs of the NWHI represent almost undamaged coral reef ecosystems with abundant and large apex predators and an extremely high proportion of endemic species across many taxa

(DeMartini and Friedlander 2004; Friedlander and DeMartini 2002), and few human impacts compared to the heavily populated MHI (Selkoe et al. 2009). In contrast, coral reefs in the MHI are under considerable anthropogenic pressure from 1.29 million residents (with over 900,000 of those living on the island of Oahu) and more than 7 million tourists visiting the state each year. Coral reefs in many of the urban areas and popular tourist destinations have significant impacts and many show ongoing declines

(Friedlander et al. 2005; Rodgers et al. 2009). The primary anthropogenic impacts to coral reefs in the MHI include coastal development and land-based sources of pollution, overfishing, recreational overuse, and alien species, whereas in the NWHI global impacts

32

such as climate change, ocean acidification and marine debris are the primary stressors

(Friedlander et al. 2005; Rodgers et al. 2009; Selkoe et al. 2009).

The Hawaiian Archipelago also lies at the periphery of the tropical Central Pacific

and is the most isolated island chain in the world, making it biogeographically partitioned from the rest of the Pacific Islands (reviewed by Ziegler 2002). This isolation results in one of the highest proportions of endemism in the world (e.g., Briggs 1974; Kay 1980;

Grigg 1983; reviewed by Ziegler 2002; Eldredge & Evenhuis, 2003). Though there are many examples of pan-pacific coral reef organisms in Hawai‘i, the isolation of the

Hawaiian Archipelago is thought to limit larval exchange sufficiently that colonization is rare (Jokiel & Martinelli 1992). For example, Kay (1984) estimated that Western Pacific marine species successfully colonize the Hawaiian Archipelago about once every 13,000 years. Unlike the terrestrial fauna, however, the Hawaiian marine fauna contains a large proportion of endemics that are differentiated but not diversified from their Indo-West

Pacific roots (Kay & Palumbi 1987; Hourigan & Reese 1987; Jokiel 1987; Ziegler 2002).

Johnston Atoll is believed to be a stepping-stone into Hawai‘i, and simulations of larval dispersal suggest that larvae from Johnston atoll can reach French Frigate Shoals or

Kaua‘i along two hypothetical dispersal corridors (Kobayashi 2006; Kobayashi &

Polovina 2006).

Delineation of management units is further complicated by the fact that single- species studies are frequently contradictory in their recommendations for management.

Analyses of connectivity frequently focus on single species exemplars which are then extrapolated to the level of the community, but the utility of exemplars in such cases is limited; even among closely-related taxa with similar life history, ecology and geographic

33 ranges, patterns of connectivity can be highly different (Rocha et al. 2002; Bird et al.

2007). In other cases, animals with highly divergent life history, ecology and geographic range (California spiny lobster, Kellet’s whelk & Kelp bass) can have surprisingly similar patterns of connectivity within the Southern California Bight (Selkoe et al. 2010). Such variability among species appears to be the rule rather than the exception, and has led to a call for explicit multi-species comparisons of connectivity across all trophic levels to broadly define the boundaries for management and determine shared borders to exchange among ecosystems. However, few such studies exist (Kelly & Palumbi 2010, Carpenter et al. 2011). The Hawaiian Archipelago makes an ideal system for a case study to determine areas of shared barriers to gene flow across many species at a variety of trophic levels to inform ecosystem-based management.

Echinoderms play a major role in structuring many marine ecosystems, and many are described as "keystone species" because of their profound influence on benthic community structure (e.g., Paine 1969, Power et al. 1996, Lessios et al. 2001, reviewed by Uthicke et al. 2009). In addition to their important ecosystem functions, many echinoderm species are also the focus of artisanal or commercial fishing efforts, particularly the sea urchins and sea cucumbers (Conand 1990; Sloan 1984, Sala et al.

1998, Purcell 2010). Stocks of the most valuable species are severely depleted throughout the Pacific with fishers continually moving to more remote locations and less valuable species (Conand 1990; Conand and Byrne 1993; Conand 1998, 2001; D’Silva 2001;

Buckner et al. 2003; Lovatelli et al. 2004; Uthicke 2004). The influence of echinoderm harvest on a wide range of other commercial fisheries, such as abalone, lobster, kelp and

34 kelp-associated fin fish, has long stimulated discussions of multispecies approaches to managing their exploitation (e.g., Sloan 1984, reviewed by Purcell 2010).

The lollyfish, Holothuria atra, and the black teatfish, Holothuria whitmaei, are two common shallow-water tropical sea cucumbers in the Indo-Pacific that extend to the

Hawaiian Archipelago, where they are locally known as loli (Clark and Rowe 1971;

Uthicke 2001; Uthicke et al. 2003; Conand 1994). Both species perform vital ecosystem services on coral reefs and they support an active fishery in many regions of the Pacific

(Bonham & Held 1963; Uthicke 1999; Purcell 2010). There has been a considerable amount of confusion regarding the names and identification of these Indo-Pacific teatfish

(reviewed by Uthicke et al. 2004). Black, white, and mottled teatfish were all included in the original description of Mülleria (= Holothuria) nobilis (Selenka, 1867). A black teatfish from Samoa was described as Holothuria whitmaei by Bell (1887) because he did not compare it to Selenka’s original description. Cherbonnier (1980) listed the mottled teatfish as a new distinct species, Holothuria fuscogilva, and treated the sympatric black specimens as H. nobilis. Rowe and Gates (1995) realized that the black and white H. nobilis specimens were separate species and the name H. nobilis was assigned to the white specimens and H. whitmaei was assigned to the black specimens. This confusion has led to many authors using H. nobilis for the black H. whitmaei (Uthicke and Benzie

2000, 2003). Uthicke et al. (2004) confirmed with genetic markers and a morphological study that the all-black teatfish is a distinct species and appropriately named H. whitmaei.

Echinoderms are described as a boom-bust phylum in which populations go through natural fluctuations in abundance (Uthicke et al. 2009), an attribute that can compound problems in a harvested population, but may hasten repopulation in previously

35

impacted areas. Holothurians are particularly susceptible to overexploitation because of their limited mobility, late maturity, density-dependent reproduction and average low rates of recruitment (Uthicke and Benzie 2000; Uthicke 2004; Uthicke et al. 2004). As such, there is a call for the management of sea cucumber harvests (Purcell 2010).

Furthermore, the boom-bust nature of echinoderms has important implications for connectivity in evolutionary time-frames where biological attributes can drive population structure to a greater extent than oceanographic processes as hypothesized in the

Tripneustes sea urchins (Lessios et al. 2003). Together these characteristics make H. atra and H. whitmaei ideal organisms to examine levels of connectivity and historical population dynamics to inform management and to test hypotheses about population connectivity within Hawai‘i connection and between Hawai‘i and its closest neighbor

Johnston Atoll. Here, we compare the mitochondrial genetic population structure of the congeners H. atra and H. whitmaei within Hawai‘i in an attempt to delineate the appropriate scales for management.

METHODS

Sampling, PCR, and Sequencing

Holothuria atra and Holothuria whitmaei were sampled from eleven sites on ten islands within the Hawaiian Archipelago and one site from Johnston Atoll, though both species were not found at every locality (fig. 3.1). Sampling in the Northwest Hawaiian

Islands took place within the Papahānaumokuākea Marine National Monument on research cruises aboard the NOAA R.V. Hi`ialakai. All other samples were collected on shore dives or while snorkeling. Sampling took place between spring 2006 and fall 2009.

36

Samples were obtained non-lethally through muscle-tissue biopsy and preserved in either

95% ethanol or DMSO-saturated salt buffer, and archived at the Hawai‘i Institute of

Marine Biology at room temperature. Skillings and Toonen (2010) contains an extended discussion of sampling and preservation protocol. No asexual morphs of H. atra --

distinguished by transverse scarring, smaller body size, and their location in lagoonal

habitats -- were found during sampling expeditions and no reports indicate the presence

of the asexual stage in the sampled locations. The asexual morph of H. atra appears to be

located only in the Southern and West Pacific (e.g., Ebert 1983; Conand 1994; Lee et al.,

2008). Asexual forms of H. whitmaei are not reported in the literature. Sample sizes and

locations are provided in Table 3.1

Total genomic DNA was extracted using DNeasy™ Blood and Tissue Kits

(QIAGEN) following the manufacturer’s instructions. Polymerase chain reaction (PCR)

was used to amplify a fragment of the mitochondrial cytochrome c oxidase subunit I gene

(COI) using custom primers created with Primer3 (Rozen & Skaletsky, 2000) targeting

Holothuria spp. : GenHol2L (5’- AACCAAATGGTTCTTGCTTACC -3’) and GenHol2R

(5’TTCTGATTAATCCCACCATCC -3’) (Skillings et al. 2011). The resolved fragment

was 423 base pairs long in H. atra and 446 base pairs in H. whitmaei. PCR was

performed using 15 µL reactions containing 1 µL of diluted DNA extract (one part

genomic extraction to 199 parts nanopure water), 1 µL each of 0.2 µM forward and

reverse primers, 0.6 µL of 0.5 µM BSA, 7.5 µL of Bioline (Bioline) Biomix Red diluted

as per manufacturer’s instructions, and 3.9 µL of nanopure water. PCR was done on Bio-

Rad Icycler™ thermocyclers (Bio-Rad Laboratories) with an initial denaturation at 95°C

for 7 min followed by 35 cycles of a denaturing step at 95°C for 1 min, annealing at 50°C

37

for 1 min, extension at 72°C for 1 min. A final extension at 72°C was held for 7 minutes.

PCR product (8 µL) was treated with 0.7 µL of Exonuclease I combined with 0.7 µL of

calf intestinal alkaline phosphatase (Exo-CIAP) and incubated at 37°C for 30 minutes

and with a final inactivation step at 85°C for 10 minutes. The treated PCR product was

sequenced using an ABI Prism 3730 automatic sequencer at the Hawai‘i Institute of

Marine Biology’s EPSCoR sequencing facility. All samples were sequenced in the

forward direction; uncertain sequences were also sequenced in the reverse direction for

confirmation. Sequences were compiled and trimmed using Sequencher 4.8 and aligned

using ClustalW implemented in Bioedit 7.0.5 (Thompson et al., 1994; Hall, 1999).

Data Analysis

Statistical parsimony networks of mitochondrial haplotypes were constructed by

creating median joining networks implemented in Network 4.610 (www.fluxus-

engineering.com; Bandelt et al., 1995; Bandelt et al., 1999). Networks were drawn using

Network Publisher 1.3.0.0 (www.fluxus-engineering.com).

Nei’s average pairwise genetic difference (π) (Nei & Li, 1979), haplotype diversity (h), Tajima’s D (Tajima, 1989), and Fu’s FS (Fu, 1997) were calculated in

DnaSP 4.1 (Rozas, 2003). The effective number of alleles (1/(1-h)) was calculated by hand following Jost (2008).

To assess levels of genetic differentiation between sites we calculated pairwise

FST and ΦST values using Arlequin 3.1 (Excoffier et al., 2005). Pairwise Dest_chao values

(Jost 2008) were calculated using SPADE (Chao, A. and Shen, T.-J. (2010) Program

SPADE (Species Prediction And Diversity Estimation Program and User’s Guide

38

published at http://chao.stat.nthu.edu.tw). ΦST is a fixation index incorporating genetic

distance that ranges from 0-1, where a zero indicates identical haplotypic composition

and a one signifies alternate fixation of alleles and a complete lack of gene flow. Dest_chao

is an index of genetic dissimilarity which does not account for genetic distance among

haplotypes, but also ranges from 0-1 (note that both ΦST and Dest_chao can be slightly

negative due to bias correction for sampling error). In the case of Dest_chao, a zero also

indicates identical haplotypic composition, but unlike ΦST, a one simply indicates that no haplotypes are shared between the populations. The primary difference in interpretation is that in the absence of gene flow ΦST values can be significantly less than one, while

this is not the case for Dest_chao, which is argued to be an advantage of this latter statistic

(Jost, 2008). The Morisita index of dissimilarity was calculated to measure overall

similarity between all sites (Chao 2008).

To correct the critical P value for statistical significance in pairwise comparisons, the family-wise false discovery rate (FDR) correction of (Benjamini et al., 2006) was

implemented. Analysis of molecular variance (AMOVA) was used for hierarchical

analysis of the partitioning of COI diversity among sites within archipelagic regions and

among archipelagic regions using Arlequin 3.1. The pairwise ΦST and AMOVA analyses were conducted using a distance matrix with 50,000 permutations and the Tamura-Nei mutational model (Tamura & Nei, 1993) with gamma = 0.0164. The mutational model

HKY+G was selected using AIC in Modeltest 3.7; the model hierarchy was used to select the closest available model when the best-fit model could not be implemented by the chosen program, as in the case of ARLEQUIN (Posada & Crandall, 1998). The inferences are robust to the mutational model and our conclusions are not altered regardless of

39

which model is chosen (data not shown). The transition-to-transversion ratio was

calculated using Modeltest.

RESULTS

Holothuria atra

A total of 252 individuals were sampled in this study. We observed 32 haplotypes,

of which 22 were private with 20 of them occurring in single individuals. Due to low

sample size, Niihau (n=5) and Gardner Pinnacles (n=2) were excluded from all between

site population analyses. The number of individuals (N), number of haplotypes (H),

number of unique haplotypes at site (Hu), nucleotide diversity averaged over sequence length (π), haplotype diversity (h), and effective number of alleles (AE) are presented in

the upper half of Table 3.1. Overall nucleotide diversity was relatively low (π = 0.0087 ±

0.0004) while the corresponding haplotype diversity was high (h = 0.88 ± 0.01).

Nucleotide diversity across all sites ranged from π = 0.0033 at Kauai to π = 0.0131 at

Johnston Atoll. Haplotype diversity ranged from h = 0.75 at Kauai to h = 0.89 at Laysan,

excluding Niihau because of low sample. There were no haplotypes shared across all

sites.

None of the site-by-site Tajima’s D values were significant, and only Laysan

deviated from expectation using Fu’s Fs; thus, there is no evidence to indicate that non- neutral processes are responsible for the pattern of COI haplotype diversity presented here.

Two Analysis of MOlecular VAriance were run on the H. atra COI haplotype data

(Table 3.4). Previous studies have shown significant population differentiation between

40 islands in the Northwest Hawaiian Islands and the Main Hawaiian Islands in multiple species (Toonen et al. 2011). We clustered sites into these two general regions, as well as pulling out Johnston Atoll as a separate region - because it lies outside of the Hawaiian

Archipelago - in order to test subdivision using AMOVAs. The two AMOVAs differed in the measure used to determine genetic diversity; FST, or strict haplotype counts, and ΦST, which incorporates genetic distance between haplotypes. Both measures displayed a similar partitioning of among-population-within-region variance, but the among-group partitioning of variance is over twice as high when measured using ΦST. Overall, all tests were significant (P < 0.0001), indicating significant genetic partitioning between regions.

Overall FST = 0.090 (P < 0.0001). Overall ΦST = 0.165 (P < 0.0001). The Morisita dissimilarity was 0.439 +/- 0.042. Pairwise comparisons between sites showed agreement between FST and ΦST measures (Table 3.3). In the Main Hawaiian Islands (MHI), O‘ahu and Kaua‘i are both significantly different from the Kona site on the Island of Hawai‘i.

These two sites were also significantly different than all other sites except Laysan Island.

There was little structure within the Northwestern Hawaiian Islands (NWHI), only

Laysan is significantly partitioned from the other sampling sites, with differentiation between the other sites within the NWHI and Johnston Atoll, but not significantly differentiated from sites within the MHI. Though outside the Hawiian Archipelago,

Johnston Atoll only showed significant differentiation from Laysan, Kauai and Oahu.

Holothuria whitmaei

We observed 27 haplotypes in 273 individuals, of which 15 were private and 15 occurring in single individuals. Due to low sample size, Lisianski (n=2) and Gardner

41

Pinnacles (n=1) were excluded from population analyses. The number of individuals (N), number of haplotypes (H), number of unique haplotypes at site (Hu), nucleotide diversity

averaged over sequence length (π), haplotype diversity (h), and effective number of

alleles (AE) are presented in the lower half Table 3.1. Overall nucleotide diversity was

relatively low (π = 0.0067 ± 0.0013) while the corresponding haplotype diversity was

relatively high (h = 0.83 ± 0.01). Nucleotide diversity ranged from π = 0.0036 at Nihoa to

π = 0.0251 at Midway. Haplotype diversity ranged from h = 0.52 at Nihoa to h = 0.90 at

Midway. There was one haplotype present at all sites and two haplotypes detected at all

sites except Gardner, Lisianski and Nihoa; three sites when combined had n = 13 with 3

haplotypes.

Tests for neutrality, Tajima’s D and Fu’s FS, are presented in Table 3.1. Midway

was the only site showing a significant Fu’s FS statistic, indicating an excess of low- frequency haplotypes, suggesting either selection or a recent demographic expansion

(Gaither, 2010). Pearl and Hermes deviated from expectation in Tajima’s D test, indicating the genetic distance between haplotypes is greater than expected, also indicating selection or recent demographic expansion. Tajima’s D and Fu’s FS were both

significant for the overall diversity.

Two AMOVAs were also run on the H. whitmaei COI haplotype data with

regional groupings identical to those of the H. atra analysis (Table 3.4). Almost all of the

variance was explained by within-population variance for both genetic measures. The

AMOVA using FST shows a small but significant amount of among-population within-

group variance.

42

Overall FST was 0.01704 (P = 0.00782). Overall ΦST was -0.00635 (P = 0.65103).

The Morisita dissimilarity index (Chao 2008), or nearly unbiased estimation of haplotypic differentiation, was 0.190 +/- 0.059. The only significant pairwise comparisons were between Nihoa and all other sites except for Kure and Kona, and only with FST (Table 3.3).

DISCUSSION

Holothuria atra and Holothuria whitmaei are the most ubiquitous sea cucumbers on Hawai‘i’s coral reefs, sharing a wide range across the central and Indo-Pacific, with

similar life histories. In this survey of genetic connectivity within the Hawaiian

Archipelago and Johnston Atoll, we were interested in any phylogeographic similarities

between these species that could be used for evaluating or planning marine management

actions. Studies have shown that a single representative or model species is seldom useful

as a proxy for estimating dispersal among marine communities, even in closely related

groups such as the endemic, sympatric, Hawaiian limpets, (Bird et al. 2007; Toonen et al.

2011). Would surveying one species give us the proper foundations for the management

of a group of wide-ranging fisheries animals? With distinct differences in population

structure, the loli, like the opihi, say ‘no’.

Biogeography and range size

If a large species range is a consequence of high dispersal potential, then both H.

atra and H. whitmaei should have little pronounced population structure, especially

across small scales (Thorson 1950; Gilman 2006; Paulay & Meyer 2006). Indeed, this is

the case for many species in the central West Pacific (Lessios et al., 2003; Craig et al.,

43

2007; Schultz et al., 2007; Eble et al., 2010; Gaither et al., 2010). Despite a species range

which stretches from the Western Red Sea to the eastern Central Pacific we found a

significant restriction on dispersal in H. atra. H. whitmaei, on the other hand, showed

almost no significant population structure despite a more restricted species range. This

pattern is consistent with population genetic survey of H. whitmaei on the Great Barrier

Reef and the Australasian Region, which found structure between regions, but not within

regions (Uthicke and Benzie 2003). These contrasting patterns highlight the dangers of

making predictions about population connectivity and diversity based solely on the

location and size of a species’ range.

The larval life history of H. atra and H. whitmaei are almost unknown, but they

require at least 18-25 days to reach competency to settle, and are capable of traversing

long oceanic distances with sufficient frequency to maintain species cohesion across a

very broad geographic range (Martinez and Richmond 1998; Laxminarayana, 2005).

Counter to intuition, the geographic distance among sites is a poor predictor of the ease

with which larvae can disperse among locations; the “oceanographic distance”

experienced by larvae between sites is uncorrelated with geographic separation between

them (Baums et al. 2006; White et al. 2010). Likewise, recent meta-analyses indicate the relationship between the length of pelagic larval development and dispersal ability is not as tight as has been generally assumed (Bradbury et al. 2008; Weersing & Toonen 2009;

Shanks 2009; Ross et al. 2009; Riginos et al. in press). Finally, a broad meta-analysis by

Lester et al. (2007) indicates the intuitive relationship between range size and larval dispersal potential are poorly correlated overall, but can play an important role in some taxa. Although the mechanism of isolation across small scales remains unknown, our data

44

clearly indicate that for H. whitmaei and H. atra relative range size does not predict

relative dispersal ability. Given that both species appear to share a similar life history,

have a similar minimum larval duration, occupy the same habitats, are both wide ranging,

and are closely related it would be expected that they would have similar levels of

population partitioning. The simplest explanation is a divergence in larval behavior or

maximum larval duration (Scheltema 1988; Shulman 1998; Riginos and Victor 2001). A

direct comparison of larval biology is lacking, but it would help to illuminate the cause

behind the patterns found in our data. Alternatively, the differences in population

structure may stem from subtle, species-specific disparities in habitat usage that impact genetic structure.

Population structure in the Hawaiian Archipelago and Johnston Atoll

Both species show high haplotype diversity and a low nucleotide diversity in the

16S gene. Except for the H. whitmaei Laysan and Nihoa populations, haplotype diversity was similar at all localities in each species. Nucleotide diversity was higher in H. atra, but there were no other apparent patterns between localities in either species; there were also no apparent patterns between species.

Our mtDNA examination of H. atra and H. whitmaei revealed contrasting patterns

of population partitioning. Regional population partitioning was not detected in H.

whitmaei with AMOVA, whereas significant regional partitioning was detected in H.

atra. Pairwise population comparisons also reflected this pattern. All pairwise H.

whitmaei comparisons had low non-significant values except for the isolation of Nihoa.

Genetic similarity in H. whitmaei is characterized by three haplotypes common

45 throughout the sampled range, the most common of which was not detected in the low number of Nihoa samples. The absence of this haplotype is likely the reason why Nihoa stands out in these comparisons.

Two interesting patterns are revealed by the H. atra population partitioning.

Excluding Laysan Island, there are no significant pairwise differences between any other islands in the NWHI (spanning nearly 2000 km), suggesting that the NWHI, excluding

Laysan, comprises a single, large population. In contrast, there is significant structuring within the MHI (roughly 600 km), and between the NWHI and the MHI. This finding suggests that factors beyond merely geographic distance influence population partitioning in H. atra. All of the population structure in H. atra is driven by the three localities of

Oahu, Kauai, and Laysan. Johnston Atoll, the nearest neighboring land mass, roughly 860 km south of French Frigate Shoals, is genetically distinct from most of the MHI and

Laysan, and genetically similar to all of the NWHI except Laysan.

It has been suggested that Johnston Atoll acts as a stepping-stone into the

Hawaiian Islands (Maragos & Jokiel, 1986). Kobayashi and Polovina (2006) and

Kobayashi (2006) used computer simulations to predict two possible larval transport corridors from Johnston Atoll to the Hawaiian Archipelago: one corridor stretching from

Johnston to French Frigate Shoals in the NWHI, and one from Johnston to O‘ahu in the

MHI. Our data support the predicted larval transport corridor between Johnston Atoll and

French Frigate Shoals, but not the corridor predicted between Johnston Atoll and Kaua‘i.

Migration between areas in the Hawaiian Archipelago is heavily one-sided with migration from the MHI into the NWHI dominating. The effectively one-way migration rates into the NWHI and Johnston Atoll coupled with the strong genetic similarity

46

between Johnston Atoll and the NWHI suggest Johnston Atoll is an isolated outpost of

the Northwest Hawaiian Islands, providing support for a vortex model (Jokiel &

Martinelli 1992) rather than the stepping stone entry into Hawai‘i (Maragos & Jokiel

1986) for H. atra.

CONCLUSIONS

Many echinoderm species are the focus of artisanal or commercial fishing efforts, and managing these fisheries requires an understanding of dispersal pathways and population connectivity within a spatial management network. The Hawaiian Archipelago lies at the periphery of the tropical Central Pacific and is the most isolated island chain in the world; the question remains as to why some species maintain connectivity and species cohesion between the Hawaiian Islands and the rest of the Pacific, why some species diverge and become Hawaiian endemics, and why other species with similar inferred dispersal ability fail to colonize the Hawaiian Archipelago.

Considerable evidence indicates that it is indefensible to make predictions of connectivity based solely on proxies such as ecological or phylogenetic similarity, pelagic larval duration, or species range sizes (Bird et al., 2007; Lester et al., 2007;

Bradbury et al., 2008; Shanks, 2009; Weersing & Toonen, 2009). The fine-scale structuring of populations in H. atra and H. whitmaei suggest that place-based management approaches, as exemplified by ecosystem based management, are ideal for responding to the complex relationships between genetically distinct populations.

Holothuria atra and Holothuria whitmaei must be managed on a local scale; migration

47 between archipelagos, and often between islands, does not occur in ecologically relevant time frames.

REFERENCES

Bandelt HJ, Forster P, Sykes BC, Richards MB (1995) Mitochondrial portraits of human populations using median networks. Genetics. 141:743-753.

Bandelt HJ, Foster P, Röhl A (1999) Median-joining networks for inferring intraspecific phylogenies. Molecular Biology and Evolution. 16(1):37-48.

Barber PH, Palumbi SR, Erdmann MV, Moosa MK (2002) Sharp genetic breaks among populations of Haptosquilla pulchella (Stomatopoda) indicate limits larval transport: patterns, causes, and consequences. Molecular Ecology. 659-674.

Baums IB, Paris CB, Cherubin LM (2006) A bio-oceanographic filter to larval dispersal in a reef-building coral. Limnology and Oceanography. 51:1969–1981.

Beerli P (2006) Comparison of Bayesian and maximum likelihood inference of population genetic parameters. Bioinformatics. 22(3):341-345.

Benjamini Y, Krieger AM, Yekutieli D (2006) Adaptive linear step-up procedures that control the false discovery rate. Biometrika. 93(3):491-507.

Bird CE, Holland BS, Bowen BW, Toonen RJ (2007) Contrasting phylogeography in three endemic Hawaiian limpets (Cellana spp.) with similar life histories. Molecular

Ecology. 16(15):3173-3186.

48

Bonham K, Held EE (1963) Ecological observations on the sea cucumbers Holothuria

atra and H. leucospilota at Rongelap Atoll, Marshall Islands. Pacific Science. 17:305-

314.

Bradbury IR, Bentzen P (2007) Non-linear genetic isolation by distance: implications for

dispersal estimation in anadromous and marine fish populations. Marine Ecology

Progress Series 340:245-257.

Bradbury IR, Laurel B, Snelgrove PVR, Bentzen P, Campana SE (2008) Global patterns

in marine dispersal estimates: the influence of geography, taxonomic category and life

history. Proceedings of the Royal Society London B 275:1803-1810.

Briggs JC (1974) Marine Zoogeography. McGraw Hill Publishers, New York, NY.

Burton R, Feldman M (1981) Population genetics of coastal and estuarine invertebrates: are behavioral differences among species reflected in population structure? Estuaries.

4:239.

Byrne M, Rowe F, Uthicke S (2010) Molecular , phylogeny and evolution in the family Stichopodidae (Aspidochirotida: Holothuroidea) based on COI and 16S mitochondrial DNA. Molecular Phylogenetics and Evolution. 56: 1068-1081.

Chao, A., Jost, L., Chiang, S. C., Jiang, Y.-H. and Chazdon, R. (2008). A Two-stage probabilistic approach to multiple-community similarity indices. Biometrics, 64, 1178-

1186.

Clark AM Rowe FEW (1971) Monograph of shallow water Indo-West Pacific echinoderms. Trustees of the British Museum (Natural History), London.

49

Conand C (1994) Asexual reproduction by fission in Holothuria atra: variability of some

parameters in populations from the tropical Indo-Pacific. Oceanologica acta. 19(4):209-

216.

Craig MT, Eble JA, Bowen BW, Robertson DR (2007) High genetic connectivity across the Indian and Pacific Oceans in the reef fish Myripristes berndti (Holocentridae). Marine

Ecology Progress Series 334:245-254.

Dawson MN (2001) Phylogeography in coastal marine animals: solution from California?

Journal of Biogeography. 28:723-736.

Dawson MN, Hamner WM (2008) A biophysical perspective on dispersal and the

geography of evolution in marine and terrestrial systems. Journal of the Royal Society

Interface. 5:135-150.

Drummond AJ, Nicholls GK, Rodrigo AG, Solomon W (2002) Estimating mutation

parameters, population history and genealogy simultaneously from temporally spaced

sequence data. Genetics. 161:1307-1320.

Drummond AJ (2003) BEAST Version 1.5.4. Available from

http://code.google.com/p/beast-mcmc/downloads/list.

Dupanloup I, Schneider, Excoffier L (2002) A simulated annealing approach to define the

genetic structure of populations. Molecular Ecology. 11(12):2571-2581.

Ebert TA (1978) Growth and size of the tropical sea cucumber Holothuria (Halodeima)

atra Jager at Enewetak Atoll, Marshall Islands. Pacific Science 32:183-191.

50

Ebert TA (1983) Recruitment in Echinoderms. In: Echinoderm Studies 1. Jangoux M,

Lawrence JM (Eds.).

Eble JA, Toonen RJ, Bowen BW (2009) Endemism and dispersal: comparative

phylogeography of three surgeonfishes across the Hawaiian Archipelago. Marine

Biology. 156(4):689-698.

Eldredge LG, Evenhuis NL (2003) Hawai‘i’s biodiversity: a detailed assessment of the

numbers of species in the Hawaiian Islands. Bishop Museum Occasional Papers 76:1-28.

Excoffier L, Laval G, Schneider S (2005) Arlequin version 3.0: an integrated software package for population genetics data analysis. Evolutionary Bioinformatics Online. 1:47-

50.

Fogarty MJ, Botsford LW (2007) Population connectivity and spatial management of marine fisheries. Oceanography. 20(3):112-123.

Fu Y-X (1997) Statistical tests of neutrality of mutations against population growth, hitchhiking, and background selection. Genetics. 147:915-925.

Gaither MR, Toonen RJ, Robertson DR, Planes S, Bowen BW (2010) Genetic evaluation of marine biogeographical barriers: perspectives from two widespread Indo-Pacific snappers (Lutjanus kasmira and Lutjanus fulvus). Journal of Biogeography. 37(1):133-

147.

Gilman SE (2006) Life at the edge: an experimental study of poleward range boundary.

Oecologia. 148:270-279.

51

Grigg RW (1983) Community structure, succession and development of coral reefs in

Hawai‘i. Marine Ecology Progress Series 11:1-14.

Grosberg RK, Cunningham CW (2001) Genetic Structure in the Sea: From Populations to

Communites. In: Marine Community Ecology. Bertness MD, Gaines SD, Hay ME (Eds.).

Sinauer Associates, Inc., Sunderland, MA. 61-84.

Hall TA (1999) BioEdit: a user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucleic Acids Symposium Series. 41:95-98.

Hedgecock D, Barber PH, Edmands S (2007) Genetic approaches to measuring connectivity. Oceanography. 20(3):70-79.

Hellberg ME (2009) Gene flow and isolation among populations of marine animals.

Annual Review of Ecology, Evolution and Systematics. 40:291-310.

Hourigan TF, Reese ES (1987) Mid-ocean isolation and evolution of Hawaiian reef fishes. Trends in Ecology and Evolution. 2:187-191.

Hutchison DW, Templeton A (1999) Correlation of pairwise genetic and geographic distance measures: inferring the relative influences of gene flow and drift on the distribution of genetic variability. Evolution. 53:1898-1914.

Jokiel PL (1987) Ecology, biogeography and evolution of corals in Hawai;i. Trends in

Ecology & Evolution. 2:179-182.

Jokiel PL, Martinelli FJ (1992) The Vortex Model of Coral Reef Biogeography. Journal of Biogeography, 19(4):449-458.

52

Jost L (2008) Gst and its relatives do not measure differentiation. Molecular Ecology.

17(18)4015-4026.

Kappel CV, Martone RG, Duffy JE (2006) "Ecosystem-based management." In:

Encyclopedia of Earth. Cleveland CJ (Ed.). Washington, D.C.: Environmental

Information Coalition, National Council for Science and the Environment.

Kay EA (1980) Little worlds of the Pacific. An essay on Pacific Basin Biogeography.

Harold L. Lyon Arboretum, University of Hawai‘i, Lecture No. 9:1-40.

Kay EA (1984) Patterns of speciation in the Indo-West Pacific. In: Radovsky FJ, Raven

PH, Sohmer SH (Eds.). Biogeography of the Tropical Pacific. Assoc. of Systematics

Collections and B.P. Bishop Museum. 15-31.

Kay EA, Palumbi SR (1987) Endemism and evolution in Hawaiian marine invertebrates.

Trends in Ecology & Evolution. 2(7):183-186.

Kinlan BP, Gaines SD (2003) Propagule dispersal in marine and terrestrial environments: a community perspective. Ecology. 84:2007-2020.

Kinlan BP, Gaines SD , Lester SE (2005) Propagule dispersal and the scales of marine community process. Diversity & Distribution. 11:139–148

Kirkendale LA and Meyer CP (2004) Phylogeography of the Patelloida profunda group

(Gastropoda: Lottidae): diversification in a dispersal-driven marine system. Molecular

Ecology. 13:2749-2762.

53

Knowlton N, Keller B (1986) Larvae which fall short of their potential: highly localized

recruitment in an alpheid shrimp with extended larval development. Bulletin of Marine

Science. 39:213-223.

Kobayashi DR (2006) Colonization of the Hawaiian Archipelago via Johnston Atoll: A characterization of oceanographic transport corridors for pelagic larvae using computer simulation. Coral Reefs. 25:407–417.

Kobayashi DR, Polovina JJ (2006) Simulated seasonal and interannual variability in larval transport and oceanography in the Northwestern Hawaiian Islands using satellite remotely sensed data and computer modeling. Atoll Research Bulletin. 543:365–390.

Laxminarayana A (2005) Induced spawning and larval rearing of the sea cucumbers,

Bohadschia marmorata and Holothuria atra in Mauritius. Fisheries Research. 48-52.

Lee J, Byrne M, Uthicke S (2008) The influence of population density on fission and growth of Holothuria atra in natural mesocosms. Journal of Experimental Marine

Biology and Ecology. 365: 126-135

Lessios HA, Garrido M, Kessing BD (2001) Demographic history of Diadema antillarum, a keystone herbivore on Caribbean reefs. Proceedings of the Royal Society.

268:1-7.

Lessios HA, Kane J, Robertson DR (2003) Phylogeography of the pantropical sea urchin

Tripneustes: contrasting patterns of population structure between oceans. Evolution.

57(9):2026-2036.

54

Lester SE, Ruttenberg BI, Gaines SD, Kinlan BP (2007) The relationship between

dispersal ability and geographic range size. Ecology Letters. 10:745-758.

Levin L (2006) Recent progress in understanding larval dispersal: new directions and

digressions. Integrative and Comparative Biology. 46(3):282-297.

Maragos JE, Jokiel PL (1986) Reef corals of Johnston Atoll: one of the world's most isolated reefs. Coral Reefs. 4(3):141–150

Manni F, Guerard E, Heyer E (2004) Geographic patterns of (genetic, morphologic, linguistic) variation: how barriers can be detected by “Monmonier’s algorithm”. Human

Biology. 76(2):173-190.

McCleod K, Leslie H (eds.) (2009) Ecosystem-based management for the Oceans. Island

Press, Washington DC, USA. 392 pp.

Miller MA, Holder MT, Vos R, Midford PE, Liebowitz T, Chan L, Hoover P, Warnow T

(2010) The CIPRES Portals. http://www.phylo.org/sub_sections/portal

Myers R, Worm B (2003) Rapid worldwide depletion of predatory fish communities.

Nature. Vol. 423: 280-283.

Nei M, Li WM (1979) Mathematical model for studying genetic variation in terms of restriction endonucleases. Proceedings of the National Academy of Sciences of the USA.

11:291-306.

Paine RT (1969) A note on trophic complexity and community stability. American

Naturalist 103: 91-93.

55

Paulay G, Meyer C (2006) Dispersal and divergence across the greatest ocean region: Do

larvae matter? Integrative and Comparative Biology. 46(3):269-281.

Pauly D, Christensen V, Dalsgaard J, Froese R, Torres Jr. F (1998) Fishing Down Marine

Food Webs. Science. Vol. 279. 5352: 860 - 863

Pauly D, Watson R (2003) Counting the last fish. Scientific America. 289(1):42-7.

Posada D, Crandall KA (1998) Modeltest: testing the model of DNA substitution.

Bioinformatics. 14:817-818.

Power ME, Tilman D, Estes JA, Menge BA, Bond WJ, Mills LS, Daily G, Castilla JC,

Lubchenco J, Paine RT (1996) Challenges in the Quest for Keystones. BioScience. 46:

609-620.

Purcell SW (2010) Managing sea cucumber fisheries with an ecosystem approach. In:

FAO Fisheries and Aquaculture Technical Paper No. 520. Lovatelli A, Vasconcellos M,

Yimin Y (Eds.). Rome, FAO. 157 pp.

Rocha LA, Bowen BW (2008) Speciation in coral-reef fishes. Journal of Fish Biology.

72(5):1101-1121.

Ronquist F, Huelsenbeck JP (2003) MRBAYES 3: Bayesian phylogenetic inference under

mixed models. Bioinformatics. 19:1572-1574.

Ross PM, Hogg ID, Pilditch CA, Lundquist CJ (2009) Phylogeography of New

Zealand’s coastal benthos. New Zealand Journal of Marine and Freshwater Research. 43:

1009-1027.

56

Rozas J (2003) DnaSP, DNA polymorphism analyses by the coalescent and other

methods. Bioinformatics. 19(18):2496-2497.

Rozen S, Skaletsky H (2000) Primer3 on the WWW for general users and for biologist programmers. In: Krawetz S, Misener S (eds) Bioinformatics Methods and Protocols:

Methods in Molecular Biology. Humana Press, Totowa, NJ. pp 365-386.

Sala E, Boudouresque CF, Harmelin-Vivien M (1998) Fishing, Trophic Cascades, and the

Structure of Algal Assemblages: Evaluation of an Old but Untested Paradigm. Oikos. 82:

425-43.

Sanchirico JN, Wilen JE (2005) Optimal spatial management of renewable resources:

Matching policy scope to ecosystem scale. Journal of Environmental Econcomics &

Management. 50:23-46.

Schultz JK, Pyle RL, DeMartini E, Bowen BW (2007) Genetic connectivity among color morphs and Pacific archipelagos for the flame angelfish, Centropyga loriculus. Marine

Biology. 151:167-175.

Selkoe KA, Henzler C, Gaines SD (2009) Seascape genetics and the spatial ecology of

marine populations. Fish and Fisheries 9:363-377.

Severance E, Karl S. Contrasting population genetic structures of sympatric, mass-

spawning Caribbean corals. Marine Biology. 150:57-68.

Shanks A, Grantham B, Carr M (2003) Propagule dispersal distance and the size and

spacing of marine reserves. Ecological Applications. 13:S159-S169.

57

Shanks AL (2009) Pelagic Larval Duration and Dispersal Distance Revisited. Biological

Bulletin. 216:373-385.

Skillings DJ and Toonen RJ (2010) It’s just a flesh wound: non-lethal sampling for conservation genetics studies. Proceedings of the American Academy of Underwater

Sciences In: NW Pollock (ed). Diving for Science. Proceedings of the 29th American

Academy of Underwater Sciences Symposium. Dauphin Island, AL: AAUS; In press.

Slatkin M (1987) Gene flow and the geographic structure of natural populations. Science.

236: 787-792.

Slatkin M, Barton N (1989) A comparison of three indirect methods for estimating

average levels of gene flow. Evolution. 43:1349-1368.

Sloan, SA (1984) Echinoderm fisheries of the world: A review. In: Echinodermata:

Proceedings of the Fifth International Echinoderm Conference. Keegan BF, O’Connor

BDS (Eds.). A.A. Balkema, Rotterdam, Netherlands. 109-124.

Sotka EE, Wares JP, Barth JA, Grosberg RK, Palumbi SR (2004) Strong genetic clines and geographical variation in gene flow in the rocky intertidal barnacle Belanus glandula.

Molecular Ecology. 13:2143-2156.

Stamatakis A (2006) RAxML-VI-HPC: maximum likelihood-based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics. 22(21):2688-2690.

Strathmann RR (1993) Hypotheses on the origins of marine larvae. Annual Review of

Ecology and Systematics. 24:89-117.

58

Tajima F (1989) Statistical method for testing the neutral mutation hypothesis by DNA

polymorphism. Genetics. 125:585-593.

Tamura K, Nei M (1993) Estimation of the number of nucleotide substitutions in the control region of mitochondrial DNA in humans and chimpanzees. Molecular Biology and Evolution. 10:512-26.

Tavaré S (1986) Some probabilistic and statistical problems in the analysis of DNA sequences. Lectures on Mathematics in the Life Sciences. American Mathematical

Society. 17:57-86.

Thompson JD, Higgins DG, Gibson TJ (1994) ClustalW: improving the sensitivity of

progressive multiple sequence alignment through sequence weighting, position-specific

gap penalties and weight matrix choice. Nucleic Acids Research. 22:4673-4680.

Thorrold SR, Zacherl DC, Levin LA (2007) Population connectivity and larval dispersal using geochemical signatures in calcified structures. Oceanography. 20(3):80-89.

Thorson G (1950) Reproductive and larval ecology of marine bottom invertebrates.

Biological Review of the Cambridge Philosophical Society. 25:1-45.

Toonen, RJ, Andrews K, Baums I, Bird C, Concepcion G, Daly-Engel T, Eble J, Faucci

A, Gaither M, Iacchei M, Puritz J, Schultz J, Skillings D, Timmers M, and Bowen B

(2011) Defining boundaries for ecosystem-based management: A multispecies case study of marine connectivity across the Hawaiian Archipelago. Journal of Marine Biology.

ID#460173, 13 pp.

59

Treml EA, Halpin PN, Urban DL, Pratson LF (2008) Modeling population connectivity

by ocean currents, a graph-theoretic approach for marine conservation. Landscape

Ecology. 23:19-36.

Uthicke S (1999) Sediment bioturbation and impact of feeding activity of Holothuria

(Halodeima) atra and Stichopus chloronotus, two sediment feeding holothurians, at

Lizard Island, Great Barrier Reef. Bulletin of Marine Science 64:129-141.

Uthicke S, Schaffelke B, Byrne M (2009) A boom-bust phylum? Ecological and

evolutionary consequences of density variations in echonoderms. Ecological

Monographs. 79:3-24.

Weersing KA, Toonen RJ (2009) Population genetics, larval dispersal, and demographic

connectivity in marine systems. Marine Ecology Progress Series 393:1-12.

White C, Selkoe KA, Watson J, Siegel DA, Zacherl DC, Toonen RJ (2010) Ocean

currents help explain population genetic structure. Proceedings of the Royal Society B:

Biological Sciences. 277: 1685-1694.

Worm B, Barbier EB, Beaumont N, Duffy JE, Folke C, Halpern BS, Jackson JBC, Lotze

HK, Micheli F, Palumbi SR, Sala E, Selkoe KA, Stachowicz JJ, Watson R (2006) Impacts

of Biodiversity Loss on Ocean Ecosystem Services. Science 314: 787 - 790.

Ziegler AC (2002) Hawaiian natural history, ecology & evolution. University of Hawai‘i

Press, Honolulu HI, USA. 477 p.

60

TABLES

Holothuria atra

Region Site N H H u π ± SD h ± SD A E Tajima's D Fu's Fs Hilo 9 4 0 0.0041 ± 0.0030 0.81 ± 0.08 5.3 -0.27 0.08 Main Kona 21 10 2 0.0078 ± 0.0046 0.87 ± 0.06 7.5 -0.57 -2.21 Hawaiian Oahu 24 7 2 0.0052 ± 0.0033 0.79 ± 0.05 4.7 -0.88 -0.58 Islands Kauai 30 8 3 0.0033 ± 0.0023 0.75± 0.06 4.0 -1.21 -2.61 Niihau 5 4 1 0.0071 ± 0.0052 0.90 ± 0.16 10.0 -0.75 -0.33 French Frigate 28 10 2 0.0082 ± 0.0048 0.88 ± 0.04 8.3 -0.14 -1.12 Gardner 2 1 0 N/A N/A N/A N/A N/A Northwest Laysan 12 8 0 0.0064 ± 0.0041 0.89 ± 0.08 9.1 -1.06 -2.91 Hawaiian Pearl & Hermes 37 10 2 0.0086 ± 0.0049 0.79 ± 0.05 4.8 -1.26 -0.23 Islands Midway 35 14 7 0.0084 ± 0.0048 0.84 ± 0.05 6.2 -0.64 -3.61 Kure 23 8 2 0.0096 ± 0.0055 0.85 ± 0.04 6.9 -0.92 0.47 Johnston 26 7 1 0.0131 ± 0.0073 0.81 ± 0.05 5.3 -0.28 2.96 Overall 252 32 22 0.0087 ± 0.0004 0.88 ± 0.01 8.3 -1.34 -2.49

Holothuria whitmaei

Region Site N H H u π ± SD h ± SD A E Tajima's D Fu's Fs Main Kona 27 7 0 0.0055 ± 0.0034 0.80 ± 0.04 5.0 -0.19 -0.01 Hawaiian Kauai 33 7 1 0.0056 ± 0.0035 0.77 ± 0.05 4.3 0.11 0.15 Nihoa 10 2 0 0.0036 ± 0.0030 0.53 ± 0.09 2.1 1.83 3.34 French Frigate 25 10 2 0.0067 ± 0.0040 0.82 ± 0.06 5.6 -0.82 -2.20 Gardner 1 1 0 N/A N/A N/A N/A N/A Northwest Lisianski 2 1 0 N/A N/A N/A N/A N/A Hawaiian Laysan 21 6 1 0.0044 ± 0.0029 0.66 ± 0.10 2.9 -0.50 -0.01 Islands Pearl & Hermes 57 13 2 0.0057 ± 0.0034 0.86 ± 0.02 7.1 -2.68 1.30 Midway 27 13 6 0.0251 ± 0.0131 0.90 ± 0.04 10.0 -1.19 -4.59 Kure 39 10 2 0.0048 ± 0.0030 0.85 ± 0.03 6.7 -0.10 -1.91 Johnston 31 9 1 0.0078 ± 0.0045 0.85 ± 0.04 6.7 0.27 -0.18 Overall 273 27 15 0.0067 ± 0.0013 0.83 ± 0.01 5.9 -2.63 -9.37

Table 3.1 N sample size, H total number of haplotypes, H u number of unique haplotypes at

site, π nucleotide diversity, h haplotype diversity, A E effective number of alleles in COI. Bolded test values are significant at <0.05

61

Holothuria atra Morisita Dissimilarity 0.439 +/- 0.042 Region ______Northwest Hawaiian Islands______Main Hawaiian Islands__ Kure Midway Pearl & Laysan French Johnston Site Kauai Oahu Kona Atoll Atoll Hermes Island Frigate Atoll Kure 0.004 -0.013 0.124 0.000 0.278 0.193 0.009 -0.013 Northwest Midway 0.028 -0.012 0.173 -0.012 0.335 0.263 0.016 0.018 Hawaiian Pearl & Hermes 0.008 -0.004 0.194 -0.014 0.341 0.258 0.026 -0.005 Islands Laysan 0.066 0.082 0.119 0.145 0.000 0.018 0.049 0.156 French Frigate -0.005 -0.004 -0.006 0.035 0.305 0.220 -0.006 0.015 Main Kauai 0.144 0.186 0.211 -0.025 0.126 0.050 0.203 0.309 Hawaiian Oahu 0.082 0.161 0.166 0.025 0.091 0.050 0.127 0.231 Islands Kona 0.018 -0.014 0.003 0.019 -0.016 0.109 0.097 0.046 Johnston -0.004 0.019 -0.010 0.110 -0.002 0.204 0.163 0.027

Holothuria whitmaei Morisita Dissimilarity 0.190 +/- 0.059 Region ______Northwest Hawaiian Islands______Main Hawaiian _ Kure Midway Pearl & Laysan French Johnston Site Nihoa Kauai Kona Atoll Atoll Hermes Island Frigate Atoll Kure 0.009 0.000 0.045 -0.002 -0.010 -0.009 -0.009 0.017 Midway -0.004 0.013 0.013 -0.011 -0.035 -0.002 -0.007 -0.005 Northwest Pearl & Hermes 0.000 -0.008 0.030 0.008 0.022 -0.010 -0.018 -0.003 Hawaiian Laysan 0.059 0.052 0.054 0.058 0.160 0.021 0.005 0.027 Islands French Frigate 0.004 -0.012 0.006 0.073 -0.029 -0.021 -0.010 -0.005 Nihoa 0.094 0.126 0.123 0.295 0.140 0.006 0.022 0.023 Main Kauai 0.011 -0.006 0.002 0.026 -0.016 0.186 -0.022 -0.005 Hawaiian Kona -0.006 -0.015 -0.011 0.020 -0.013 0.128 -0.019 -0.022 Johnston 0.029 -0.003 0.006 0.057 -0.004 0.212 -0.004 -0.004

Table 3.2 Pairwise comparisons by site. Fst values are contained in the lower left half of each table and Φst values are in the upper right half of each table. Bolded values signify significant differences after correction using the procedure outlined in Benjamini 2008. Shaded cells signify significant differences between sites in both tests.

62

Holothuria whitmaei Holothuria - Atoll 0.167 0.000 0.038 0.194 0.000 0.000 0.000 0.708 Johnston Johnston - Kona 0.000 0.000 0.000 0.058 0.000 0.000 0.352 0.144 ------Oahu 0.515 0.782 0.048 0.000 0.010 0.072 0.172 0.000 Kauai 0.495 0.499 0.901 __Main Hawaiian__Main Islands__ - - - - 0.303 0.451 0.422 0.647 0.403 Nihoa - 0.024 0.000 0.035 0.230 0.000 0.000 0.512 0.629 French French Frigate values in are the lower left half whitmaei of the table H. and Holothuria atra Holothuria - 0 0.209 0.203 0.205 0.286 0.122 0.138 0.694 Island Laysan - 0.003 0.000 0.000 0.008 0.000 0.674 0.888 0.744 Pearl & Hermes - Atoll 0.000 0.000 0.555 0.000 0.000 0.093 0.880 0.846 Midway - ______Northwest Hawaiian Islands______Kure Atoll 0.160 0.035 0.483 0.111 0.000 0.666 0.410 -0.030 comparisons by species site. comparisons and H. atra est_chao Site Kure Kona Kauai Oahu Nihoa Laysan Midway Johnston French Frigate Pearl & Hermes sites. Bolded values signify significant differences using outlined the procedure correction after in Benjamini 2008. Table D 3.3 Pairwise Region Islands values in are the half upper right of the table. Cells were withspecies "-" indicate done a comparisons no that for between those Hawaiian Northwest

63

Table 3.4 Analyses of molecular variance (AMOVA) by region using haplotype distance (ΦST) and non-

distance (FST) measures. Source of % of Species Measure Regions Φ statistics P-Values Variation Variation AG 5.8 F = 0.058 <0.01 MHI; NWHI; CT Holothuria atra F ST Johnston Atoll AP(G) 3.22 F SC = 0.034 <0.01 WP 90.99 AG 12.65 Φ = 0.127 <0.01 MHI; NWHI; CT Holothuria atra Φ ST Johnston Atoll AP(G) 3.84 Φ SC = 0.044 <0.01 WP 83.51 AG -0.77 F = -0.008 0.50 MHI; NWHI; CT Holothuria whitmaei F ST Johnston Atoll AP(G) 2.48 F SC = 0.025 0.01 WP 98.3 AG -0.87 Φ = 0.373 0.87 MHI; NWHI; CT Holothuria whitmaei Φ ST Johnston Atoll AP(G) 0.67 Φ SC = 0.032 0.24 WP 100.2 AG = Among Groups, AP(G) = Among populations within Groups, WP = Within Populations

64

Table 3.5 Pairwise population migration

rate estimates (N e Μ )based on a Bayesian MCMC simulation. The value of M calculated by Migrate was multiplied by the θ, as calculated by Migrate, of the destination population to estimate migration. The estimates of migration are seperated by direction; the columns are source populations and the the rows are sink populations. MHI NWHI Johnston MHI - 2.59875 1.073025 NWHI 37.65388 - 79.09125 Johnston 0.3135 3.5416 -

65

FIGURES

The Hawaiian Archipelago Papah ānaumoku

ākea Marine National Monument

Midway Pearl and Hermes

Laysan Gardner Main Hawaiian Islands

French Frigate Shoals Kauai Oahu Maui

Hawaii

Figure 3.1 Map of the Hawaiian Archipelago

66

Figure 3.2 Haplotype network for Holothuria atra. Each circle represents a unique haplotype connected by a line to those that differ by one base pair. Nodes on lines indicate a missing haplotype and numbers represent multiple missing haplotypes. Each haplotype is color-coded by site and circle size is proportional to frequency. The smallest circles represent one occurrence of a haplotype.

67

Figure 3.3 Haplotype network for Holothuria whitmaei. Each circle represents a unique haplotype connected by a line to those that differ by one base pair. Nodes on lines indicate a missing haplotype and numbers represent multiple missing haplotypes. Each haplotype is color-coded by site and circle size is proportional to frequency. The smallest circles represent one occurrence of a haplotype.

68

CHAPTER FOUR

Gateways to Hawai‘i – Genetic population structure of the tropical sea cucumber

Holothuria atra

Published as:

Skillings DJ, Bird CE and Toonen RJ (2011) Gateways to Hawai‘i – genetic population

structure of the tropical sea cucumber Holothuria atra. Journal of Marine Biology.

Article ID 783030:1-16.

69

ABSTRACT

Holothuria atra is one of the most common and widest ranging tropical, coral reef sea cucumbers in the world, and here we examine population genetic structure based on mitochondrial COI to aid in determining the appropriate scale for coral reef management.

Based on SAMOVA, AMOVA and BARRIER analyses, we show that despite its large range, H. atra has hierarchical, fine-scale population structure driven primarily by between-archipelago barriers, but with significant differences between sites within an archipelago as well. Migrate analyses along with haplotype networks and patterns of haplotype diversity suggest that Hawai‘i and Kingman reef are important centers of the genetic diversity in the region rather than an evolutionary dead-end for migrants from the

Indo-Pacific. Finally we show that for H. atra Kingman Reef is the most likely stepping stone between Hawai‘i and the rest of the Pacific, not Japan or Johnston Atoll as previously presumed. Based on our data, Johnston Atoll can instead be seen as an outpost of the Northwestern Hawaiian Islands rather than a gateway to the Hawaiian

Archipelago.

INTRODUCTION

Echinoderms play a major role in structuring many marine ecosystems, and many

are described as "keystone species" because of their profound influence on benthic

community structure (e.g., Paine 1969, Power et al. 1996, Lessios et al. 2001, reviewed

by Uthicke et al. 2009). In addition to their important ecosystem functions, many

echinoderm species are also the focus of artisanal or commercial fishing efforts,

particularly the sea urchins and sea cucumbers (Sloan 1984, Sala et al. 1998, Purcell

70

2010). The influence of echinoderm harvest on a wide range of other commercial fisheries, such as abalone, lobster, kelp and kelp-associated fin fish, has long stimulated

discussions of multispecies approaches to managing their exploitation (e.g., Sloan 1984,

reviewed by Purcell 2010). Delineation of the appropriate spatial scales for management

zones within a spatial management network requires a detailed understanding of dispersal

pathways and population connectivity (reviewed by Hedgecock et al. 2007; Thorrold et

al. 2007; Fogarty & Botsford 2007).

Understanding connectivity in the sea is complicated by the fact that many marine organisms share a biphasic life cycle typified by an adult form that is relatively sedentary and a larval form that can potentially disperse across large expanses of open ocean

(Thorson, 1950; Strathmann, 1993; Kinlan & Gaines, 2003; Kinlan et al., 2005; Paulay &

Meyer, 2006). For example, in the sea urchin genus Tripneustes some well-known biogeographic barriers, such the Isthmus of Panama or the long stretch of deep water in the western Atlantic, are important barriers to dispersal whereas others, such as the

Eastern Pacific Barrier show no evidence for limiting dispersal (Lessios et al. 2003).

However, the geographic limits of such dispersal are uncertain because it is virtually impossible with current technology to directly track these microscopic juveniles during the pelagic phase (reviewed by Levin 2006) making indirect methods of quantifying larval dispersal particularly attractive (reviewed by Grosberg & Cunningham 2001;

Hedgecock et al. 2007; Selkoe et al. 2009; Hellberg 2009). Proxies for dispersal, such as pelagic larval duration (PLD) and geographic range have generally been used as rules of thumb in the absence of a detailed understanding of connectivity for most marine species.

Unfortunately, intuitive expectations of larval dispersal potential as a function of PLD

71

and range size are not upheld in recent meta-analyses of the existing literature (Lester et al. 2007; Bradbury & Bentzen 2007; Bradbury et al. 2008; Weersing & Toonen 2009;

Shanks 2009; Ross et al. 2009; Riginos et al. in press). Realized dispersal distance is typically less than potential dispersal distance because of the presence of biophysical or biogeographical barriers (Burton & Feldman, 1981; Knowlton & Keller, 1986; Shanks et al., 2003; Severance & Karl, 2006; Dawson & Hamner, 2008). Barriers that limit dispersal between marine populations include obvious geographical features such as land masses, i.e. the Isthmus of Panama (Bermingham & Lessios, 1993), but also more subtle factors such as currents and oceanographic regimes (Dawson, 2001; Barber et al., 2002;

Sotka et al., 2004; Baums et al., 2006; Treml et al., 2008). The correlation between geographic distance and the probability of larval exchange among sites is low in many marine systems (e.g., White et al. 2010), and thus quantitative estimates of connectivity are an important prerequisite for delineating the appropriate scale over which marine populations ought to be managed.

The Hawaiian Archipelago lies at the periphery of the tropical Central Pacific and is the most isolated island chain in the world, making it biogeographically partitioned from the rest of the Pacific Islands (reviewed by Ziegler 2002). This isolation results in one of the highest proportions of endemism in the world (e.g., Briggs 1974; Kay 1980;

Grigg 1983; reviewed by Ziegler 2002; Eldredge & Evenhuis, 2003). Though there are many examples of pan-pacific coral reef organisms in Hawai‘i, the isolation of the

Hawaiian Archipelago is thought to limit larval exchange sufficiently that colonization is rare (Jokiel & Martinelli 1992). For example, Kay (1984) estimated that Western Pacific marine species successfully colonize the Hawaiian Archipelago about once every 13,000

72

years. Unlike the terrestrial fauna, however, the Hawaiian marine fauna contains a large

proportion of endemics that are differentiated but not diversified from their Indo-West

Pacific roots (Kay & Palumbi 1987; Hourigan & Reese 1987; Jokiel 1987; Ziegler 2002).

Johnston Atoll is believed to be a stepping-stone into Hawai‘i, and simulations of larval

dispersal suggest that larvae from Johnston atoll can reach French Frigate Shoals or

Kaua‘i along two separate larval corridors (Kobayashi 2006; Kobayashi & Polovina

2006).

The lollyfish, Holothuria atra, is one of the most common shallow-water tropical sea cucumbers in the Indo-Pacific, spanning from Madagascar to French Polynesia (Clark and Rowe 1971; Conand 1994). H. atra performs vital ecosystem services on coral reefs for which there is an active fishery in many regions of the Pacific (Bonham & Held 1963;

Uthicke 1999; Purcell 2010). Echinoderms are described as a boom-bust phylum in which populations go through marked natural population cycles (Uthicke et al. 2009), an attribute that can compound problems in a harvested population, but may hasten repopulation in previously impacted areas. As such, there is a call for ecosystem-based management of sea cucumber harvests (Purcell 2010). Furthermore, the boom-bust nature of echinoderms has important implications for connectivity in evolutionary time-frames where biological attributes can drive population structure to a greater extent than oceanographic processes as hypothesized in the Tripneustes sea urchins (Lessios et al.

2003) Together these characteristics make H. atra an ideal organism to examine levels of connectivity and historical population dynamics to inform management and to test hypotheses about Hawai‘i’s connection with other archipelagos in the Central Pacific.

Here, we assess the inferred range of dispersal for H. atra in Hawai‘i and the Central

73

Pacific by investigating its mitochondrial genetic population structure in an attempt to

delineate the appropriate scales for management.

METHODS

Sampling, PCR, and Sequencing

Holothuria atra was sampled from five archipelagos (Hawaiian Islands, Line

Islands, Marshal Islands, Bonin Islands, and Ryukyu Islands) at a total of 19 sites (Figure

4.1). Sampling in the Northwest Hawaiian Islands and the Line Islands took place on research cruises aboard the NOAA R.V. Hi`ialakai. All other samples were collected on shore dives or while snorkeling. Sampling took place between spring 2006 and fall 2009.

Samples were obtained non-lethally through muscle-tissue biopsy and preserved in either

95% ethanol or DMSO salt buffer, and archived at the Hawai‘i Institute of Marine

Biology at room temperature. Skillings and Toonen (2010) contains an extended discussion of sampling and preservation protocol. No asexual morphs -- distinguished by transverse scarring, smaller body size, and their location in lagoonal habitats -- were found during sampling expeditions and no reports are known indicating the presence of the asexual stage of H. atra in the sampled locations. The asexual morph of H. atra appears to be located only in the Southern and West Pacific (e.g., Ebert 1983; Conand

1994; Lee et al., 2008).

Total genomic DNA was extracted using DNeasy™ Blood and Tissue Kits

(QIAGEN) following the manufacturer’s instructions. Polymerase chain reaction (PCR) was used to amplify a 423 base pair fragment of the mitochondrial cytochrome c oxidase subunit I gene (COI) using custom primers created with Primer3 (Rozen & Skaletsky,

74

2000) targeting Holothuria spp. : GenHol2L (5’- AACCAAATGGTTCTTGCTTACC -3’) and GenHol2R (5’- TTCTGATTAATCCCACCATCC -3’). PCR was performed using 15

µL reactions containing 1 µL of diluted DNA extract (one part template DNA to 199 parts nanopure water), 1 µL each of 0.2 µM forward and reverse primers, 0.6 µL of 0.5 µM

BSA, 7.5 µL of Bioline (Bioline) Biomix Red diluted as per manufacturer’s instructions, and 3.9 µL of nanopure water. PCR was done on Bio-Rad Icycler™ thermocyclers (Bio-

Rad Laboratories) with an initial denaturation at 95°C for 7 min followed by 35 cycles of a denaturing step at 95°C for 1 min, annealing at 50°C for 1 min, extension at 72°C for 1 min. A final extension at 72°C was held for 7 min before refrigeration. PCR product (8

µL) was treated with 0.7 µL of Exonuclease I combined with 0.7 µL of calf intestinal alkaline phosphatase (Exo-CIAP) and incubated at 37°C for 30 minutes and with a final inactivation step at 85°C for 10 minutes. The treated PCR product was sequenced using an ABI Prism automatic sequencer at the Hawai‘i Institute of Marine Biology’s EPSCoR sequencing facility. All samples were sequenced in the forward direction; uncertain sequences and all unique haplotypes were also sequenced in the reverse direction for confirmation. Sequences were compiled and trimmed using Sequencher 4.8 and aligned using ClustalW implemented in Bioedit 7.0.5 (Thompson et al., 1994; Hall, 1999).

Data Analysis

A statistical parsimony network of mitochondrial haplotypes was constructed by creating a reduced median network that was then used to make a median joining network; both procedures implemented in Network 4.516 (www.fluxus-engineering.com; Bandelt et al., 1995; Bandelt et al., 1999). The network was drawn using Network Publisher

1.1.0.7 (www.fluxus-engineering.com).

75

Nei’s average pairwise genetic difference (π) (Nei & Li, 1979) and haplotype

diversity (h) were calculated in DnaSP 4.1 (Rozas, 2003). The effective number of alleles

was calculated by hand following Jost (2008). Tajima’s D (Tajima, 1989), and Fu’s FS

(Fu, 1997) were calculated in ARLEQUIN v.3.1 with 10,000 simulations to establish

confidence intervals.

To assess levels of genetic differentiation between sites we calculated pairwise

ΦST values using Arlequin 3.1 (Excoffier et al., 2005) and pairwise Dest_chao values by

hand following Jost (Jost 2008). ΦST is a fixation index incorporating genetic distance

that ranges from 0-1, where a zero indicates identical haplotypic composition and a one

signifies alternate fixation of alleles and a complete lack of gene flow. Dest_chao is an index of genetic differentiation which does not account for genetic distance among haplotypes, but also ranges from 0-1 (note that both ΦST and Dest_chao can be slightly

negative due to bias correction for sampling error). In the case of Dest_chao, a zero also

indicates identical haplotypic composition, but unlike ΦST, a one simply indicates that no haplotypes are shared between the populations. The primary difference in interpretation is that in the absence of gene flow ΦST values can be significantly less than one, while

this is not the case for Dest_chao, which is argued to be an advantage of this latter statistic

(Jost, 2008). To correct the critical p value for statistical significance in pairwise

comparisons, the family-wise false discovery rate (FDR) correction of (Benjamini, 2006)

was implemented. Analysis of molecular variance (AMOVA) was used for hierarchical

analysis of the partitioning of COI diversity among sites within archipelagic regions and

among archipelagic regions using Arlequin 3.1. SAMOVA 1.0 was used to identify

groups of samples that maximize the proportion of total genetic variance due to

76

differences between regions (Dupanloup et al., 2002). The most important genetic

barriers were ranked using BARRIER 2.2 (Manni et al., 2004). BARRIER uses

Monmonier’s maximum-difference algorithm to compare a matrix of difference values,

such as pairwise ΦST values, with a matrix of geographic distances in order to identify the strongest barriers within the matrix. We compared barriers created using each ΦST and

Dest_chao distance matrix. AMOVAs were performed using groupings determined by

SAMOVA and BARRIER for hypothesis testing to compare genetic groupings to the

archipelagic groupings. The pairwise ΦST and AMOVA analyses were conducted using a

distance matrix with 50,000 permutations and the Tamura-Nei mutational model (Tamura

& Nei, 1993) with gamma = 0.0164. The mutational model HKY+G was selected using

AIC in Modeltest 3.7; the model hierarchy was used to select the closest available model when the best-fit model could not be implemented by the chosen program, as in the case of ARLEQUIN (Posada & Crandall, 1998). Regardless, the inferences are robust to the

mutational model and our conclusions are not altered regardless of which model is

chosen (data not shown).

MrBayes 3.1 was used to construct a Bayesian estimation of a phylogeny

containing all H. atra haplotypes from this study along with all Holothuria and

Actinopyga COI haplotypes available from GenBank as of February 2010 (Ronquist &

Huelsenbeck, 2003). Two independent runs with identical conditions were completed and

averaged. A general time reversible (GTR) simple nucleotide model with a gamma-

shaped rate variation of 0.0164 was used; Markov chain length = 4 x 3,000,000 sampled

every 100 generations with a 10% burn-in. The GTR nucleotide model was chosen as it is

the most general and neutral nucleotide model available in MrBayes 3.1 and corresponds

77

most closely to the Tamura-Nei model (Tavaré, 1986). The sea cucumber

agassizi was set as the outgroup. Program defaults were used for all other settings.

MrBayes was used to summarize all of the trees produced into a single consensus tree.

RAxML 7.0 (Stamatakis, 2006) implemented through CIPRES Web Portal v.1.15

(Miller et al., 2010) was used to construct the highest scoring maximum likelihood-based estimation of a phylogeny containing all haplotypes used in the Bayesian analysis and run

10,000 bootstrap simulations to assess branch support. A GTR nucleotide model that uses four discrete gamma rates set by the program was used for the analysis, program defaults were used for all other settings.

Bayesian coalescent-based calculations of migration rate among regions (NeΜ )

and the region mutation parameter (θ) were conducted using MIGRATE 3.1.3 (Beerli,

2006). Three independent runs of a Bayesian MCMC search strategy were completed and

averaged by MIGRATE. A nucleotide model with a transition-to-transversion ratio of

6.1584:1 and three regions of substitution rates with a gamma-shaped rate variation of

0.016 was used; Markov chain length = 1,000,000 sampled every 20 generations with a

10% burn-in. Program defaults were used for all other settings. The transition-to-

transversion ratio was calculated using Modeltest 3.7. Two replicate MIGRATE analyses

were run using different population groupings. Preliminary analyses that split the data by

sampling location returned flat posterior probabilities, presumably from having too many

parameters to estimate. The software’s author advocates using the minimal number of

sensible regions in order to reach convergence (Peter Beerli, pers. comm.) The first

analysis used regions separated along the most important breaks identified by the

program BARRIER. Archipelagos were used as regions for the second analysis with the

78

Hawaiian Archipelago further divided between the main Hawaiian Islands and the

Northwestern Hawaiian Islands plus Johnston Atoll and Kingman Reef separated from

the Line Islands; this grouping was chosen based on BARRIER and pairwise analyses.

Values for the migration rate among regions (Μ) and region mutation parameters (θ) were

taken from the highest peaks in the posterior probability distribution curves. The

posterior probability distributions were examined to determine the validity of each

estimated parameter.

RESULTS

A total of 385 individuals, 55 haplotypes, and 37 private haplotypes are sampled

in this study (Table 4.1). Of the 18 haplotypes shared across sampling sites, 10 are found

in multiple archipelagos (Figure 4.2). However, no haplotypes are shared between the

most distant regions: Hawai‘i and Japan. Because many population genetic estimates are

relatively insensitive to weak selection (Slatkin & Barton, 1989), loci which do not show

significant deviations from neutral expectations should provide reliable inferences about

population structure (Hutchison & Templeton, 1999). None of the site-by-site Tajima’s D values were significant, and only Laysan deviated from expectation using Fu’s Fs; thus,

there is no evidence to indicate that non-neutral processes are responsible for the pattern of COI haplotype diversity presented here.

To ensure there was no misidentification of the samples included here, we performed a phylogenetic reconstruction of our samples with those available in GenBank.

We confirmed that the samples included here are monophyletic and there are no

79

reciprocally monophyletic groups among the H. atra haplotypes included in our

population genetic analyses (Appendix 4.1).

Haplotype diversity, as a function of longitude, increases from west to east across

the Pacific Ocean (Table 1, R2=0.80, p<0.05). Japanese sites exhibit the lowest haplotype

diversity (h=0.51-0.66) and effective number of haplotypes (HE= 2.0-2.9); the centrally

located Line Islands (h=0.65-0.90, HE=2.9-9.7) and Marshall Islands (h=0.75-0.76,

HE=4.0-4.1) exhibit mid to high levels of diversity; and the Hawaiian sites exhibit the

highest diversity (h=0.75-0.90 , HE=4-10 ) Nucleotide diversity does not appear to be correlated with haplotype diversity because nucleotide diversity is lower in the Hawaiian

Archipelago than in all other locations except for Okinawa (Table 4.1). This pattern can be visualized in the haplotype network, where sites exhibiting high nucleotide diversity harbor disparate haplotypes separated by a relatively large number of mutations (Figure

4.2).

Four AMOVAs were run on the H. atra COI haplotype data (Table 4.2). In each

AMOVA a different method was employed to group the population samples into regions.

Under the first geographic hypothesis, sites were grouped into five regions by archipelago, with Johnston Atoll included with Hawai‘i. In order to assess the subdivision of the Hawaiian Archipelago into the NWHI and the MHI, this grouping hypothesis was compared to a second geographic hypothesis where population samples were grouped into six regions: Hawaiian sites were divided into the MHI and the NWHI

+ Johnston and four regions comprised of the four remaining archipelagos. Both grouping hypotheses (geographic hypothesis one and two) exhibited similar partitioning of variation among-groups (32.1% vs. 30.9%) and among-populations-within-groups (8.9%

80

vs 5.5 %). In both cases there was stronger partitioning among the groups of samples

(ΦCT=0.32, ΦCT =0.31; P<0.0001) than among the simples nested within the groupings

(ΦSC=0.13, ΦSC=0.07; P<0.0001).

The two geographic hypotheses were compared to six-region groupings identified

by SAMOVA and BARRIER; BARRIER selected the same dominant barriers using both

the ΦST and Dest_chao distance matrices (Table 4.2, Figure 4.1). Six regions were chosen for

a direct comparison to the archipelagic geographic hypothesis that included Hawai‘i

divided into two main regions. These groupings partitioned variance similarly to the

geographic hypotheses with the SAMOVA grouping minimizing among-population

within-group variance with more among-group variance explained (Table 4.2). The

grouping of population samples using BARRIER and SAMOVA had slightly greater

levels of genetic differentiation among groups of samples (ΦCT=0.33, ΦCT =0.37;

P<0.0001), and lower levels of differentiation among samples nested within groups

(ΦSC=0.07, ΦSC=0.03; P<0.0001). Overall, the four AMOVAs exhibited similar levels of

partitioning of variance and all tests were significant (P<0.0001).

There is a strong pattern of restricted gene flow between sites among the

population samples of H. atra. Pairwise comparisons for both ΦST and Dest_chao reveal

significant differences between sites located in different archipelagos in almost all cases

where sample sizes are 10 or greater (Table 4.3). The exceptions are between the Line

Islands and the Marshall Islands where one of six pairwise ΦST comparisons are

statistically significant; between the Line Islands and the Bonin Islands, one of three ΦST

comparisons are statistically significant; and between the Hawaiian Islands and Kingman

Reef where only four of twelve pairwise ΦST comparisons were statistically significant.

81

Some significant differences were also detected among samples within archipelagos. In

the Main Hawaiian Islands (MHI), O‘ahu and Kaua‘i are significantly different than the

Kona sample from the Big Island of Hawai‘i. Despite small sample sizes, Niihau is also

partitioned from the adjacent island of Kaua‘i as well as O‘ahu, but not the Big Island.

Within the Northwestern Hawaiian Islands (NWHI), only Laysan is significantly

partitioned from the other sampling sites, including Johnston Atoll. Overall, 42% of

pairwise comparisons between the MHI and the NWHI + Johnston were significant,

compared to 30% of the comparisons within the MHI and 19% of the comparisons within

the NWHI + Johnston. The samples from both the Line Islands (LI) and the Marshall

Islands (MI) were significantly partitioned within their respective archipelagos when

using Dest_chao, there was not significant partitioning between MI sites when using ΦST.

The results from the MIGRATE runs show similar patterns of gene flow between

regions (Table 4.4, Figure 4.1). Effective migration rates (NeΜ) between regions are low.

There is less than one migrant per generation, the rule-of-thumb number below which population cohesion starts to break down, between most regions (Slatkin, 1987). The exceptions include the one-way migration from the Main Hawaiian Islands into the

Northwest Hawaiian Islands and Johnston Atoll and the one-way migration from

Kingman Reef to the Main Hawaiian Islands (Table 4.4). The high effective migration rate from the NWHI to the MHI in the first analysis splits almost evenly between

Kingman Reef and the MHI when these two regions are separated in the second analysis

(Table 4.4). Overall, higher effective migration rates are observed leaving the Hawaiian regions then going into them (Table 4.4, Figure 4.1). Though effective migration rates are a product of migration and effective population size, the effective migration rates larger

82 than 1 migrant per generation are driven primarily by migration and not effective population size (Appendices 4.1 and 4.2). This pattern is indicative of recent migration rather than ancestral polymorphisms and high effective population sizes. Posterior probability distributions for all values were in the form of unimodal curves. A full description of Ne and Μ values for both analyses can be found in Appendices 4.2 and 4.3.

DISCUSSION

In this survey of population genetic structure we elucidate patterns of connectivity throughout the north-central range of the sea cucumber Holothuria atra with a focus on the Hawaiian Archipelago. The Hawaiian Archipelago is highly isolated, and also contains one of the highest proportions of endemism in the world (e.g., Briggs 1974; Kay

1980; Grigg 1983; reviewed by Ziegler 2002; Eldredge & Evenhuis, 2003). Though there are many pan-pacific marine organisms in Hawai‘i, the isolation of the archipelago is thought to limit larval exchange such that colonization is rare but sufficient to maintain species cohesion among these taxa. The Hawaiian marine fauna contains a large proportion of endemics that are differentiated but not diversified from its Indo-West

Pacific roots (Kay & Palumbi 1987; Hourigan & Reese 1987; Jokiel 1987; Ziegler 2002).

In this scenario Hawai‘i is seen primarily as a dead-end, an isolated land-mass that does not contribute in a significant way to the overall diversity of the tropical pacific. Counter to the island biogeography hypotheses of Hawaiian diversity, Jokiel and Martinelli (1992) proposed the Vortex model of speciation, wherein the stunning biodiversity of the Coral

Triangle is a result of centrifugal accumulation of species from the peripheral habitats around the Pacific. Though these two models primarily make predictions about speciation-level processes and do not speak directly to gene-flow within a species, they

83

do make opposite claims about the dominant direction of gene-flow and dispersal. H. atra

has a broad species range, extending from the Western Indian Ocean to the Eastern

Pacific Ocean, which suggests the capacity for long-distance dispersal; however,

populations showed significant population structuring within archipelagos, sometimes

across very short oceanic distances. Even so, hierarchical genetic population structure in

H. atra gives insight into the phylogeography of the north-central tropical Pacific. Our

data test between the divergent hypotheses of whether periphery archipelagos act as a

source of genetic diversity in the Pacific and the likely colonization routes, into and out

of, the extremely isolated Hawaiian Archipelago.

Biogeography and range size

If a large species range is a consequence of high dispersal potential, then H. atra should have little pronounced population structure, especially across small scales

(Thorson 1950; Gilman 2006; Paulay & Meyer 2006). Indeed, this is the case for many species in the central West Pacific (Lessios et al., 2003; Craig et al., 2007; Schultz et al.,

2007; Eble et al., 2010; Gaither et al., 2010). Despite a species range which stretches from the Western Red Sea to the eastern Central Pacific in which H. atra is found in almost all shallow tropical habitats, we did not find support for extensive dispersal. The majority of sites from which we sampled H. atra were genetically distinct, with some

sites less than 75 km apart being among the most distinct in our study (Table 4.3). These

contrasting patterns highlight the dangers of making predictions about population

connectivity and diversity based solely on the location and size of a species’ range.

84

The larval life history of H. atra is not known exactly, but they require at least 18-25 days

to reach competency to settle, and are capable of traversing long oceanic distances with

sufficient frequency to maintain species cohesion across a very broad geographic range

(Laxminarayana, 2005). The obvious question becomes: why then is population

subdivision found on such small geographic scales (e.g., Kingman Reef and Palmyra

Atoll are only 67 km apart)? Counter to intuition, the geographic distance among sites is

a poor predictor of the ease with which larvae can disperse among locations; the

“oceanographic distance” experienced by larvae between sites is uncorrelated with

geographic separation between them (Baums et al. 2006; White et al. 2010). Likewise, recent meta-analyses indicate the relationship between the length of pelagic larval development and dispersal ability is not as tight as has been generally assumed (Bradbury et al. 2008; Weersing & Toonen 2009; Shanks 2009; Ross et al. 2009; Riginos et al. in press). Finally, a broad meta-analysis by Lester et al. (2007) indicates the intuitive relationship between range size and larval dispersal potential are poorly correlated overall, but can play an important role in some taxa. Toonen et al. in this issue also how a number of breaks in the Hawaiian Archipelago that are shared by several species and are unexplained solely by appeal to one metric such as range size or larval dispersal potential. Although the mechanism of isolation across small scales remains unknown, our data clearly indicate that H. atra is not one of those species for which range size predicts relative dispersal ability.

Population structure in the Hawaiian Archipelago and Johnston Atoll

Our mtDNA examination of Holothuria atra reveals significant genetic population structure across the surveyed portion of the range. There are two interesting

85 patterns to this structure. Excluding Laysan Island, there are no significant pairwise differences between any other islands in the NWHI (spanning nearly 2000 km), suggesting that the NWHI, excluding Laysan, comprises a single, large population. In contrast, there is significant structuring within the MHI (roughly 600 km), and between the NWHI and the MHI. This finding suggests that factors beyond merely geographic distance influence population partitioning.

Johnston Atoll, the nearest neighboring land mass, roughly 860 km south of

French Frigate Shoals, is genetically distinct from most of the MHI and Laysan, and genetically similar to all of the NWHI except Laysan. It has been suggested that Johnston

Atoll acts as a stepping-stone into the Hawaiian Islands (Maragos & Jokiel, 1986).

Kobayashi (Kobayashi & Polovina, 2006; Kobayashi, 2006) used computer simulations to predict two larval transport corridors from Johnston Atoll to the Hawaiian Archipelago: one corridor stretching from Johnston to French Frigate Shoals in the NWHI, and one from Johnston to O‘ahu in the MHI. Our data support the predicted larval transport corridor between Johnston Atoll and French Frigate Shoals, but not the corridor predicted between Kaua‘i and Johnston. Additionally, based on our data, Kingman Reef may also be an important stepping stone into and out of Hawai‘i. The BARRIER analysis shows the division between the NWHI, including Johnston Atoll, and the MHI to be the strongest barrier to gene flow within the Archipelago (Figure 4.1). Migration across this barrier is heavily one-sided where migration from the MHI into the NWHI dominates.

The effectively one-way migration rates into the NWHI and Johnston Atoll coupled with the strong genetic similarity between Johnston Atoll and the NWHI suggest Johnston

Atoll is an isolated outpost of the Northwest Hawaiian Islands, providing support for a

86 vortex model (Jokiel & Martinelli 1992) rather than the stepping stone entry into Hawai‘i

(Maragos & Jokiel 1986) for H. atra. These data indicate that Johnston Atoll exchanges migrants with Hawai‘i far more often than its nearest neighbors to the south and the same can be said for Kingman Reef. This result is particularly surprising because in the case of

Kingman Reef, H. atra sampled there show greater similarity to populations in Hawai‘i

(roughly 1700 km southwest of Honolulu) than they do to those sampled at Palmyra

Atoll, only 67 km away.

Phylogeographic relationships between Archipelagos

Counter to conventional wisdom that Hawai‘i is a passive recipient of rare dispersal from the diverse Pacific, the weight of available evidence, including pairwise

ΦST values, mtDNA phylogeny, BARRIER divisions, and clustering within the haplotype network, provides substantial evidence for the opposite pattern in H. atra; Johnston Atoll is an outpost of Hawaiian diversity, and Kingman Reef acts as the primary stepping stone between the Hawaiian Archipelago and the rest of the Pacific. As far as we are aware, this is the first time empirical evidence has been provided for such a pathway. Higher haplotypic diversity in Hawai‘i and the Line Islands relative to the other archipelagos supports a scenario in which population sizes are far greater, or Hawai‘i and/or Kingman

Reef are the ancestral population in the region. Likewise, the dominant haplotypes found in the Japanese Archipelagos are relatively distantly related and appear derived (Figure

4.2), suggesting the western portion of the surveyed range was colonized in at least two separate events (or one of them has gone extinct in Hawai‘i and Kingman), one of which did not make it all the way to Okinawa.

87

Excluding the dense sampling within the Hawaiian Archipelago, virtually all

pairwise comparisons between sites are significantly different from each other (Table

4.3). The few comparisons that were not significant between sites have relatively high

pairwise values but low sample sizes, a likely statistical limitation also noted by Bird et

al. (2007). Hierarchical population structuring was detected with AMOVA using either

Dest or ΦST values; sampling sites within archipelagos are significantly different from

each other, but are more similar within than between archipelagos. Four of the five most

substantial restrictions to gene flow uncovered in this study (and the top ranked by

BARRIER) were those between archipelagos; the one exception to this trend being

Kingman Reef which is included with the Hawaiian rather than the Line Islands (Figure

4.1). The AMOVA run using the regions selected by BARRIER was only minimally

different than the AMOVA run using regions divided by archipelagos; grouping Kingman

Reef with the MHI, as per BARRIER, did explain 1.5% more of the overall variance.

In addition to the distinct archipelagic groupings in the haplotype network (Figure

4.2), several other patterns are noteworthy here. First, the NWHI and Johnston Atoll haplotypes are clustered together and interspersed, whereas the MHI haplotypes are clustered together. Also, the Japanese haplotypes occur in two divergent areas of the network. The Line Island and Marshall Island haplotypes are inter-dispersed throughout the network, suggesting that these island groups are either mixing or transition zones.

Nearly every locality haplotype (those found in only one sampling location), branch off in a starburst pattern from the major haplotypes found primarily in the same archipelago.

This pattern is an indication that regional populations have been separated long enough for new haplotypes to arise, and that these new haplotypes are not being spread to other

88

archipelagos by long-distance dispersal. Uniformly low migration rates between

archipelagos estimated with MIGRATE support this isolation scenario.

CONCLUSIONS

Many echinoderm species are the focus of artisanal or commercial fishing efforts,

and managing these fisheries requires a detailed understanding of dispersal pathways and

population connectivity within a spatial management network. The Hawaiian Archipelago lies at the periphery of the tropical Central Pacific and is the most isolated island chain in the world; the question remains as to why some species maintain connectivity and species cohesion between the Hawaiian Islands and the rest of the Pacific, why some species diverge and become Hawaiian endemics, and why other species with similar inferred dispersal ability fail to colonize the Hawaiian Archipelago at all.

The genetic diversity of COI in H. atra across the studied portion of the range presents a complex pattern, but it is not inscrutable. Based on AMOVA, SAMOVA and

BARRIER analyses it can be seen that population structuring is hierarchical; there are significant differences between sites but the primary degree of population structure is archipelago by archipelago. Our analyses taken together suggest that the Hawaiian

Archipelago and Kingman reef are ancestral populations in the region with migration moving out of these periphery archipelagos toward a less diverse central Pacific rather than the reverse. This pattern is inconsistent with the hypothesis that Hawai‘i is a dead-

end for rare migrants from the Indo-Pacific. Instead, the weight of the evidence shows

that these peripheral populations are not sinks, but important centers for the generation of

genetic diversity feeding back towards the West Pacific. Specifically for H. atra, our data

89

suggest that the pathway between Hawai‘i and the rest of the Pacific is primarily out

through Kingman Reef and the Line Islands and not in through Japan, the Marshall

Islands or the closest neighbor to the Hawaiian Archipelago, Johnston Atoll. We show that, at least for H. atra, Johnston Atoll is in fact an outpost of the Northwestern

Hawaiian Islands and not a primary gateway for colonization of the Archipelago.

Considerable evidence is accumulating that it is indefensible to make predictions of connectivity based solely on proxies such as ecological or phylogenetic similarity, pelagic larval duration, or species range sizes (Bird et al., 2007; Lester et al., 2007;

Bradbury et al., 2008; Shanks, 2009; Weersing & Toonen, 2009). The fine-scale structuring of populations in H. atra suggest that place-based management approaches, as exemplified by ecosystem based management, are ideal for responding to the complex relationships between genetically distinct populations. Holothuria atra must be managed on a local scale; migration between archipelagos, and often between islands, does not occur in ecologically relevant time frames.

ACKNOWLEDGEMENTS

We thank the Papahānaumokuākea Marine National Monument, US Fish and

Wildlife Services, and Hawai‘i Division of Aquatic Resources (DAR) for coordinating research activities and permitting, and the National Oceanic and Atmospheric

Administration (NOAA) research vessel Hi‘ialakai and her crew for years of outstanding service and support. Special thanks go to B. Bowen, the members of the ToBo Lab, UH

Dive Program, NMFS, PIFSC, CRED, M. Skillings, K. Boyle, J. Claisse, D. Wagner, P.

Aldrich, M. Iacchei, J. Puritz, J. Eble, I. Baums, M. Timmers, N. Yasuda, R. Kosaki, S.

90

Karl, C. Meyer, S. Godwin, M. Stat, X. Pochon, H. Kawelo, T. Daly-Engel, M. Craig, L.

Rocha, M. Gaither, G. Conception, Y. Papastamatiou, M. Crepeau, Z. Szabo, J. Salerno

and the HIMB NSF-EPSCoR Core Genetics Facility. We also thank the anonymous

reviewers who put in the extra time to help strengthen the quality of this work.

This work was funded in part by grants from the National Science Foundation

(DEB#99-75287, OCE#04-54873, OCE#06-23678, OCE#09-29031), National Marine

Sanctuaries NWHICRER-HIMB partnership (MOA-2005-008-6882), National Marine

Fisheries Service, NOAA's Coral Reef Conservation Program, and the Hawai‘i Coral

Reef Initiative. This is contribution #1421 from the Hawai‘i Institute of Marine Biology, and SOEST 8049.

91

REFERENCES

Bandelt HJ, Forster P, Sykes BC, Richards MB (1995) Mitochondrial portraits of human populations using median networks. Genetics. 141:743-753.

Bandelt HJ, Foster P, Röhl A (1999) Median-joining networks for inferring intraspecific phylogenies. Molecular Biology and Evolution. 16(1):37-48.

Barber PH, Palumbi SR, Erdmann MV, Moosa MK (2002) Sharp genetic breaks among populations of Haptosquilla pulchella (Stomatopoda) indicate limits larval transport: patterns, causes, and consequences. Molecular Ecology. 659-674.

Baums IB, Paris CB, Cherubin LM (2006) A bio-oceanographic filter to larval dispersal in a reef-building coral. Limnology and Oceanography. 51:1969–1981.

Beerli P (2006) Comparison of Bayesian and maximum likelihood inference of population genetic parameters. Bioinformatics. 22(3):341-345.

Benjamini Y, Krieger AM, Yekutieli D (2006) Adaptive linear step-up procedures that control the false discovery rate. Biometrika. 93(3):491-507.

Bird CE, Holland BS, Bowen BW, Toonen RJ (2007) Contrasting phylogeography in three endemic Hawaiian limpets (Cellana spp.) with similar life histories. Molecular

Ecology. 16(15):3173-3186.

Bonham K, Held EE (1963) Ecological observations on the sea cucumbers Holothuria atra and H. leucospilota at Rongelap Atoll, Marshall Islands. Pacific Science. 17:305-

314.

92

Bradbury IR, Bentzen P (2007) Non-linear genetic isolation by distance: implications for dispersal estimation in anadromous and marine fish populations. Marine Ecology

Progress Series. 340:245-257.

Bradbury IR, Laurel B, Snelgrove PVR, Bentzen P, Campana SE (2008) Global patterns in marine dispersal estimates: the influence of geography, taxonomic category and life history. Proceedings of the Royal Society London B. 275:1803-1810.

Briggs JC (1974) Marine Zoogeography. McGraw Hill Publishers, New York, NY.

Burton R, Feldman M (1981) Population genetics of coastal and estuarine invertebrates: are behavioral differences among species reflected in population structure? Estuaries.

4:239.

Byrne M, Rowe F, Uthicke S (2010) Molecular taxonomy, phylogeny and evolution in the family Stichopodidae (Aspidochirotida: Holothuroidea) based on COI and 16S mitochondrial DNA. Molecular Phylogenetics and Evolution. 56: 1068-1081.

Clark AM Rowe FEW (1971) Monograph of shallow water Indo-West Pacific echinoderms. Trustees of the British Museum (Natural History), London.

Conand C (1994) Asexual reproduction by fission in Holothuria atra: variability of some parameters in populations from the tropical Indo-Pacific. Oceanologica acta. 19(4):209-

216.

93

Craig MT, Eble JA, Bowen BW, Robertson DR (2007) High genetic connectivity across

the Indian and Pacific Oceans in the reef fish Myripristes berndti (Holocentridae). Marine

Ecology Progress Series. 334:245-254.

Dawson MN (2001) Phylogeography in coastal marine animals: solution from California?

Journal of Biogeography. 28:723-736.

Dawson MN, Hamner WM (2008) A biophysical perspective on dispersal and the

geography of evolution in marine and terrestrial systems. Journal of the Royal Society

Interface. 5:135-150.

Drummond AJ, Nicholls GK, Rodrigo AG, Solomon W (2002) Estimating mutation

parameters, population history and genealogy simultaneously from temporally spaced

sequence data. Genetics. 161:1307-1320.

Drummond AJ (2003) BEAST Version 1.5.4. Available from

http://code.google.com/p/beast-mcmc/downloads/list.

Dupanloup I, Schneider, Excoffier L (2002) A simulated annealing approach to define the

genetic structure of populations. Molecular Ecology. 11(12):2571-2581.

Ebert TA (1978) Growth and size of the tropical sea cucumber Holothuria (Halodeima)

atra Jager at Enewetak Atoll, Marshall Islands. Pacific Science 32:183-191.

Ebert TA (1983) Recruitment in Echinoderms. In: Echinoderm Studies 1. Jangoux M,

Lawrence JM (Eds.).

94

Eble JA, Toonen RJ, Bowen BW (2009) Endemism and dispersal: comparative

phylogeography of three surgeonfishes across the Hawaiian Archipelago. Marine

Biology. 156(4):689-698.

Eldredge LG, Evenhuis NL (2003) Hawai‘i’s biodiversity: a detailed assessment of the

numbers of species in the Hawaiian Islands. Bishop Museum Occasional Papers. 76:1-28.

Excoffier L, Laval G, Schneider S (2005) Arlequin version 3.0: an integrated software package for population genetics data analysis. Evolutionary Bioinformatics Online. 1:47-

50.

Fogarty MJ, Botsford LW (2007) Population connectivity and spatial management of marine fisheries. Oceanography. 20(3):112-123.

Fu Y-X (1997) Statistical tests of neutrality of mutations against population growth, hitchhiking, and background selection. Genetics. 147:915-925.

Gaither MR, Toonen RJ, Robertson DR, Planes S, Bowen BW(2010) Genetic evaluation of marine biogeographical barriers: perspectives from two widespread Indo-Pacific snappers (Lutjanus kasmira and Lutjanus fulvus). Journal of Biogeography. 37(1):133-

147.

Gilman SE (2006) Life at the edge: an experimental study of poleward range boundary.

Oecologia. 148:270-279.

Grigg RW (1983) Community structure, succession and development of coral reefs in

Hawai‘i. Marine Ecology Progress Series. 11:1-14.

95

Grosberg RK, Cunningham CW (2001) Genetic Structure in the Sea: From Populations to

Communites. In: Marine Community Ecology. Bertness MD, Gaines SD, Hay ME (Eds.).

Sinauer Associates, Inc., Sunderland, MA. 61-84.

Hall TA (1999) BioEdit: a user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucleic Acids Symposium Series. 41:95-98.

Hedgecock D, Barber PH, Edmands S (2007) Genetic approaches to measuring connectivity. Oceanography. 20(3):70-79.

Hellberg ME (2009) Gene flow and isolation among populations of marine animals.

Annual Review of Ecology, Evolution and Systematics. 40:291-310.

Hourigan TF, Reese ES (1987) Mid-ocean isolation and evolution of Hawaiian reef fishes. Trends in Ecology and Evolution. 2:187-191.

Hutchison DW, Templeton A (1999) Correlation of pairwise genetic and geographic distance measures: inferring the relative influences of gene flow and drift on the distribution of genetic variability. Evolution. 53:1898-1914.

Jokiel PL (1987) Ecology, biogeography and evolution of corals in Hawai;i. Trends in

Ecology and Evolution. 2:179-182.

Jokiel PL, Martinelli FJ (1992) The Vortex Model of Coral Reef Biogeography. Journal

of Biogeography, 19(4):449-458.

Jost L (2008) Gst and its relatives do not measure differentiation. Molecular Ecology.

17(18)4015-4026.

96

Kappel CV, Martone RG, Duffy JE (2006) "Ecosystem-based management." In:

Encyclopedia of Earth. Cleveland CJ (Ed.). Washington, D.C.: Environmental

Information Coalition, National Council for Science and the Environment.

Kay EA (1980) Little worlds of the Pacific. An essay on Pacific Basin Biogeography.

Harold L. Lyon Arboretum, University of Hawai‘i, Lecture No. 9:1-40.

Kay EA (1984) Patterns of speciation in the Indo-West Pacific. In: Radovsky FJ, Raven

PH, Sohmer SH (Eds.). Biogeography of the Tropical Pacific. Assoc. of Systematics

Collections and B.P. Bishop Museum. 15-31.

Kay EA, Palumbi SR (1987) Endemism and evolution in Hawaiian marine invertebrates.

Trends in Ecology and Evolution. 2(7):183-186.

Kinlan BP, Gaines SD (2003) Propagule dispersal in marine and terrestrial environments: a community perspective. Ecology. 84:2007-2020.

Kinlan BP, Gaines SD , Lester SE (2005) Propagule dispersal and the scales of marine community process. Diversity and Distribution. 11:139–148

Kirkendale LA and Meyer CP (2004) Phylogeography of the Patelloida profunda group

(Gastropoda: Lottidae): diversification in a dispersal-driven marine system. Molecular

Ecology. 13:2749-2762.

Knowlton N, Keller B (1986) Larvae which fall short of their potential: highly localized recruitment in an alpheid shrimp with extended larval development. Bulletin of Marine

Science. 39:213-223.

97

Kobayashi DR (2006) Colonization of the Hawaiian Archipelago via Johnston Atoll: A characterization of oceanographic transport corridors for pelagic larvae using computer simulation. Coral Reefs. 25:407–417.

Kobayashi DR, Polovina JJ (2006) Simulated seasonal and interannual variability in larval transport and oceanography in the Northwestern Hawaiian Islands using satellite remotely sensed data and computer modeling. Atoll Research Bulletin. 543:365–390.

Laxminarayana A (2005) Induced spawning and larval rearing of the sea cucumbers,

Bohadschia marmorata and Holothuria atra in Mauritius. Fisheries Research. 48-52.

Lee J, Byrne M, Uthicke S (2008) The influence of population density on fission and growth of Holothuria atra in natural mesocosms. Journal of Experimental Marine

Biology and Ecology. 365: 126-135

Lessios HA, Garrido M, Kessing BD (2001) Demographic history of Diadema antillarum, a keystone herbivore on Caribbean reefs. Proceedings of the Royal Society.

268:1-7.

Lessios HA, Kane J, Robertson DR (2003) Phylogeography of the pantropical sea urchin

Tripneustes: contrasting patterns of population structure between oceans. Evolution.

57(9):2026-2036.

Lester SE, Ruttenberg BI, Gaines SD, Kinlan BP (2007) The relationship between dispersal ability and geographic range size. Ecology Letters. 10:745-758.

Levin L (2006) Recent progress in understanding larval dispersal: new directions and digressions. Integrative and Comparative Biology. 46(3):282-297.

98

Maragos JE, Jokiel PL (1986) Reef corals of Johnston Atoll: one of the world's most

isolated reefs. Coral Reefs. 4(3):141–150

Manni F, Guerard E, Heyer E (2004) Geographic patterns of (genetic, morphologic,

linguistic) variation: how barriers can be detected by “Monmonier’s algorithm”. Human

Biology. 76(2):173-190.

McCleod K, Leslie H (eds.) (2009) Ecosystem-based management for the Oceans. Island

Press, Washington DC, USA. 392 pp.

Miller MA, Holder MT, Vos R, Midford PE, Liebowitz T, Chan L, Hoover P, Warnow T

(2010) The CIPRES Portals. http://www.phylo.org/sub_sections/portal

Myers R, Worm B (2003) Rapid worldwide depletion of predatory fish communities.

Nature. Vol. 423: 280-283.

Nei M, Li WM (1979) Mathematical model for studying genetic variation in terms of

restriction endonucleases. Proceedings of the National Academy of Sciences of the USA.

11:291-306.

Paine RT (1969) A note on trophic complexity and community stability. American

Naturalist. 103: 91-93.

Paulay G, Meyer C (2006) Dispersal and divergence across the greatest ocean region: Do

larvae matter? Integrative and Comparative Biology. 46(3):269-281.

Pauly D, Christensen V, Dalsgaard J, Froese R, Torres Jr. F (1998) Fishing Down Marine

Food Webs. Science. Vol. 279. 5352: 860 - 863

99

Pauly D, Watson R (2003) Counting the last fish. Scientific American. 289(1):42-7.

Posada D, Crandall KA (1998) Modeltest: testing the model of DNA substitution.

Bioinformatics.14:817-818.

Power ME, Tilman D, Estes JA, Menge BA, Bond WJ, Mills LS, Daily G, Castilla JC,

Lubchenco J, Paine RT (1996) Challenges in the Quest for Keystones. BioScience. 46:

609-620.

Purcell SW (2010) Managing sea cucumber fisheries with an ecosystem approach. In:

FAO Fisheries and Aquaculture Technical Paper No. 520. Lovatelli A, Vasconcellos M,

Yimin Y (Eds.). Rome, FAO. 157pp.

Rocha LA, Bowen BW (2008) Speciation in coral-reef fishes. Journal of Fish Biology.

72(5):1101-1121.

Ronquist F, Huelsenbeck JP (2003) MRBAYES 3: Bayesian phylogenetic inference under mixed models. Bioinformatics. 19:1572-1574.

Ross PM, Hogg ID, Pilditch CA, Lundquist CJ (2009) Phylogeography of New

Zealand’s coastal benthos. New Zealand Journal. Marine and Freshwater Research. 43:

1009-1027.

Rozas J (2003) DnaSP, DNA polymorphism analyses by the coalescent and other methods. Bioinformatics. 19(18):2496-2497.

100

Rozen S, Skaletsky H (2000) Primer3 on the WWW for general users and for biologist programmers. In: Krawetz S, Misener S (eds) Bioinformatics Methods and Protocols:

Methods in Molecular Biology. Humana Press, Totowa, NJ. pp 365-386.

Sala E, Boudouresque CF, Harmelin-Vivien M (1998) Fishing, Trophic Cascades, and the

Structure of Algal Assemblages: Evaluation of an Old but Untested Paradigm. Oikos. 82:

425-43.

Sanchirico JN, Wilen JE (2005) Optimal spatial management of renewable resources:

Matching policy scope to ecosystem scale. Journal of Environmental Economics &

Management 50:23-46.

Schultz JK, Pyle RL, DeMartini E, Bowen BW (2007) Genetic connectivity among color morphs and Pacific archipelagos for the flame angelfish, Centropyga loriculus. Marine

Biology. 151:167-175.

Selkoe KA, Henzler C, Gaines SD (2009) Seascape genetics and the spatial ecology of marine populations. Fish and Fisheries 9:363-377.

Severance E, Karl S. Contrasting population genetic structures of sympatric, mass- spawning Caribbean corals. Marine Biology. 150:57-68.

Shanks A, Grantham B, Carr M (2003) Propagule dispersal distance and the size and spacing of marine reserves. Ecological Applications. 13:S159-S169.

Shanks AL (2009) Pelagic Larval Duration and Dispersal Distance Revisited. Biological

Bulletin. 216:373-385.

101

Skillings, D.J. & R.J. Toonen (2010) It’s just a flesh wound: non-lethal sampling for

conservation genetics studies. Proceedings of the American Academy of Underwater

Sciences In: NW Pollock (ed). Diving for Science. Proceedings of the 29th American

Academy of Underwater Sciences Symposium. Dauphin Island, AL: AAUS; In press.

Slatkin M (1987) Gene flow and the geographic structure of natural populations. Science.

236: 787-792.

Slatkin M, Barton N (1989) A comparison of three indirect methods for estimating

average levels of gene flow. Evolution. 43:1349-1368.

Sloan, SA (1984) Echinoderm fisheries of the world: A review. In: Echinodermata:

Proceedings of the Fifth International Echinoderm Conference. Keegan BF, O’Connor

BDS (Eds.). A.A. Balkema, Rotterdam, Netherlands. 109-124.

Sotka EE, Wares JP, Barth JA, Grosberg RK, Palumbi SR (2004) Strong genetic clines and geographical variation in gene flow in the rocky intertidal barnacle Belanus glandula.

Molecular Ecology. 13:2143-2156.

Stamatakis A (2006) RAxML-VI-HPC: maximum likelihood-based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics. 22(21):2688-2690.

Strathmann RR (1993) Hypotheses on the origins of marine larvae. Annual Review of

Ecology and Systematics. 24:89-117.

Tajima F (1989) Statistical method for testing the neutral mutation hypothesis by DNA polymorphism. Genetics. 125:585-593.

102

Tamura K, Nei M (1993) Estimation of the number of nucleotide substitutions in the

control region of mitochondrial DNA in humans and chimpanzees. Molecular Biology

and Evolution. 10:512-26.

Tavaré S (1986) Some probabilistic and statistical problems in the analysis of DNA

sequences. Lectures on Mathematics in the Life Sciences. American Mathematical

Society. 17:57-86.

Thompson JD, Higgins DG, Gibson TJ (1994) ClustalW: improving the sensitivity of

progressive multiple sequence alignment through sequence weighting, position-specific

gap penalties and weight matrix choice. Nucleic Acids Research. 22:4673-4680.

Thorrold SR, Zacherl DC, Levin LA (2007) Population connectivity and larval dispersal

using geochemical signatures in calcified structures. Oceanography. 20(3):80-89.

Thorson G (1950) Reproductive and larval ecology of marine bottom invertebrates.

Biological Review of the Cambridge Philosophical Society. 25:1-45.

Treml EA, Halpin PN, Urban DL, Pratson LF (2008) Modeling population connectivity by ocean currents, a graph-theoretic approach for marine conservation. Landscape

Ecology. 23:19-36.

Uthicke S (1999) Sediment bioturbation and impact of feeding activity of Holothuria

(Halodeima) atra and Stichopus chloronotus, two sediment feeding holothurians, at

Lizard Island, Great Barrier Reef. Bulletin of Marine Science 64:129-141.

103

Uthicke S, Schaffelke B, Byrne M (2009) A boom-bust phylum? Ecological and

evolutionary consequences of density variations in echinoderms. Ecological

Monographs. 79:3-24.

Weersing KA, Toonen RJ (2009) Population genetics, larval dispersal, and demographic

connectivity in marine systems. Marine Ecology Progress Series. 393:1-12 .

White C, Selkoe KA, Watson J, Siegel DA, Zacherl DC, Toonen RJ (2010) Ocean

currents help explain population genetic structure. Proceedings of the Royal Society B:

Biological Sciences. 277: 1685-1694.

Worm B, Barbier EB, Beaumont N, Duffy JE, Folke C, Halpern BS, Jackson JBC, Lotze

HK, Micheli F, Palumbi SR, Sala E, Selkoe KA, Stachowicz JJ, Watson R (2006) Impacts

of Biodiversity Loss on Ocean Ecosystem Services. Science 314: 787 - 790.

Ziegler AC (2002) Hawaiian natural history, ecology & evolution. University of Hawai‘i

Press, Honolulu HI, USA. 477 p.

104

FIGURES

105

106

TABLES

107

108

109

110

APPENDICES

111

112

113

CHAPTER FIVE

Contrasting Phylogeography Patterns in Two Pacific Brittle Stars, Ophiocoma erinaceous

and Ophiocoma pica

114

INTRODUCTION

Coastal marine ecosystems are in global decline due to the increasing intensity and diversity of stresses in near-shore waters (Hutchings 2000; Jackson et al. 2001). In response to the threats to coral reefs (Wilkinson 2000; Jackson et al. 2001; Pandolfi et al.

2003; Hughes et al. 2003; Sale et al. 2005; McLeod et al. 2009), marine protected areas

(MPAs) have established worldwide to help mitigate stressors, strengthen resilience, and facilitate continued ecosystem viability (Lubchenco et al. 2003; Hughes et al. 2006;

Mumby et al. 2006). However, the efficacy of MPAs in mitigating the effects of stressors and establishing locally persistent communities is fundamentally linked to the scale of dispersal and the degree of connectivity among populations (Eckert 2003; Sanchirico &

Wilen 2005; Botsford et al. 2009). Delineation of the appropriate spatial scales for management zones within a spatial management network requires an understanding of dispersal pathways and population connectivity (reviewed by Hedgecock et al. 2007;

Thorrold et al. 2007; Fogarty & Botsford 2007). Population connectivity, or lack thereof, also contributes to population genetic differentiation and thus plays a key role in the evolution processes that drive local adaptation and speciation (Bradbury et al. 2008;

Walter et al. 2009).

Most coral reef organisms live in spatially patchy environments where populations are connected primarily by pelagic larval dispersal via ocean currents

(Thorson, 1950; Sale 1980; Strathmann, 1993; Kinlan & Gaines, 2003; Kinlan et al.,

2005; Lessios & Robertson 2006; Paulay & Meyer, 2006; Dawson & Hamner 2008). The duration of the pelagic larval stage ranges from a few minutes to as long as a year, providing the opportunity for long-distance dispersal across large expanses of open

115 ocean. Some larvae also have the ability, and perhaps the propensity, to return to their natal region following residency in the pelagic environment (Joens et al. 2005; Almany et al. 2007). Behavioral traits may minimize long-distance dispersal, possibly assuring that larvae will be near suitable habitat (Fisher and Bellwood 2003). Thus, local conditions and behavioral processes have the potential to contribute substantially to the demography and genetic connectivity of reef populations.

Despite recent advances, there are currently insufficient empirical data to make generalizations on expected patterns of connectivity among coral reefs. Proxies for dispersal, such as pelagic larval duration (PLD) and geographic range have generally been used as rules of thumb in the absence of a detailed understanding of connectivity for most marine species. However, intuitive expectations of larval dispersal potential as a function of PLD are not upheld in detailed meta-analyses of the existing literature (Lester et al. 2007; Bradbury & Bentzen 2007; Bradbury et al. 2008; Weersing & Toonen 2009;

Shanks 2009; Ross et al. 2009). Realized dispersal distance is typically less than potential dispersal distance because of the presence of biophysical or biogeographical barriers (Burton & Feldman, 1981; Knowlton & Keller, 1986; Shanks et al., 2003;

Severance & Karl, 2006; Dawson & Hamner, 2008). The correlation between geographic distance and the probability of larval exchange among sites is low in many marine systems (e.g., White et al. 2010), and thus quantitative estimates of connectivity are an important prerequisite for determining the size and spacing of reserves that will maximize their effectiveness (Fogarty and Botsford 2007; Jones et al . 2007).

MPA design often centers on simply protecting a target fraction of key habitats

(reviewed by Sale et al. 2005), rather than the effects of MPAs on metapopulation

116 dynamics (Kritzer and Sale 2006). It has been commonly assumed that the communities naturally occurring in the protected habitats will persist in them, and contribute to unprotected habitats outside MPAs. Modeling studies have indicated how larval dispersal and the spatial configuration of MPAs interact to promote population persistence

(Crowder et al. 2000; Botsford et al. 2001; Kaplan et al. 2006). However, while these interactions are clear in modeling results, efforts to apply conclusions from these models to assess and design effective MPAs are hindered by uncertainty about larval dispersal

(Stockhausen et al. 2000; Botsford et al. 2001). In discontinuous habitat, such as the

Hawaiian Archipelago, planktonic larval dispersal among islands might serve to counteract local recruitment that could otherwise lead to local inbreeding and population subdivision. At larger scales however (i.e. among islands in the central Pacific), larval exchange among populations may be limited by vast geographical distances’ leading to regional isolation.

The Hawaiian archipelago stretches more than 2500 km in length, and consists of two regions: the main Hawaiian Islands (MHI) which are populated, high volcanic islands; and the Northwestern Hawaiian Islands (NWHI) which are an uninhabited string of tiny islands, atolls, shoals, and banks. The oldest island in the chain, Kure Atoll, lies to the north-west and has been dated a c. 25 million years (Myr) old. The chain becomes progressively younger, and the exposed land masses greater, to the south-east, terminating with the volcanically active island of Hawai‘i (see Neall and Trewick, 2008, for a review of Hawaiian palaeogeology). The archipelago is one of the most isolated on the planet, the nearest continent (North America) lying some 4100 km to the east, and the

117 nearest emergent land at Johnston and Palmyra atolls lying c. 800 km and c. 1500 km, respectively, from the center of the chain.

Due to their isolation, the roughly 4,500 square miles of coral reefs found throughout the NWHI are among the healthiest and most extensive remaining in the world (Pandolfi et al. 2003). The reefs of the NWHI represent almost undamaged coral reef ecosystems with abundant and large apex predators and an extremely high proportion of endemic species across many taxa (DeMartini and Friedlander 2004; Friedlander and

DeMartini 2002), and few human impacts compared to the heavily populated MHI

(Selkoe et al. 2009).

The Central and South Pacific, including Hawai‘i, have been the focus of several recent studies examining phylogeography of marine organisms including fishes and invertebrates (Rivera et al. 2004; Bird et al. 2007; Craig et al. 2007, 2010; Schultz et al.

2007; Ramon et al. 2008; Eble et al. 2009, 2011; DiBatista et al. 2011; Skillings et al.

2011; Timmers et al. 2011). Taken together, these studies show that genetic connectivity over both evolutionary and ecological time-scales is variable, and that this region is affected by several processes that interact to influence phylogeographic patterns across taxa. Variability among species appears to be the rule rather than the exception, and has led to a call for explicit multi-species comparisons of connectivity among many species and across all trophic levels to broadly define the boundaries for management and determine shared borders to exchange among ecosystems (Toonen et al. 2011).

The Hawaiian Archipelago lies at the periphery of the tropical Central Pacific and is the most isolated island chain in the world, making it biogeographically partitioned

118

from the rest of the Pacific Islands (reviewed by Ziegler 2002). This isolation results in

one of the highest proportions of endemism in the world (e.g., Briggs 1974; Kay 1980;

Grigg 1983; reviewed by Ziegler 2002; Eldredge & Evenhuis, 2003). Though there are

many examples of pan-pacific coral reef organisms in Hawai‘i, the isolation of the

Hawaiian Archipelago is thought to limit larval exchange sufficiently that colonization is

rare (Jokiel & Martinelli 1992; Randall 1998, Mundy 2005). For example, Kay (1984)

estimated that Western Pacific marine species successfully colonize the Hawaiian

Archipelago about once every 13,000 years. Community similarities indicate the

possibility of ongoing connections between the geographically remote Hawaiian

Archipelago and the broader Indo-Pacific, but the rate of connectivity is difficult to assess. The broad distribution of many species similarly limits the identification of dispersal corridors and the directionality of colonization (Hourigan & Reese 1987, Jokiel

1987, Craig et al. 2010). Nevertheless, Hawai‘i’s marine fauna is widely regarded as a biogeographic ‘dead end’, with diversity flowing into but not out of the islands (Hourigan

& Reese 1987, Jokiel 1987, Kay & Palumbi 1987, Randall 1998, Briggs 1999). This assumption is beginning to be challenged as evidence accumulates that genetic diversity originating in Hawai‘i is flowing back out to the rest of the Pacific (Eble et al. 2011;

Skillings et al. 2011). Johnston Atoll has been proposed as a link between Hawai‘i and other Central Pacific communities (Kobayashi 2006; Kobayashi & Poloyina 2006), though faunal affinities and recent genetic studies indicate that Johnston is more likely an outpost of Hawai‘i rather than a link to the broader Pacific (Hourigan & Reese 1987,

Skillings et al. 2011, Timmers et al. 2011).

119

Ophiocoma erinaceous and Ophiocoma pica are two common shallow-water

brittle stars associated with coral reefs throughout the tropical Pacific. Echinoderms are

described as a boom-bust phylum in which populations go through marked natural

population cycles (Uthicke et al. 2009).The boom-bust nature of echinoderms has important implications for connectivity in evolutionary time-frames where biological attributes can drive population structure to a greater extent than oceanographic processes as hypothesized in the Tripneustes sea urchins (Lessios et al. 2003). O. erinaceous and O.

pica have been found in overlapping habitats. Both species are found in coral rubble and

within densely branching coral heads, especially dead, crustose coralline algae encrusted

Pocillopora meandrina; though O. pica is more commonly found in coral heads and O.

erinaceous is more commonly associated with loose coral rubble (D.J.S. personal

observation). Based on habitat affiliation, we expect these two species to have concordant

population genetic structure. Here, we examine levels of genetic connectivity in O.

erinaceous and O. pica in in order to test hypotheses about genetic connectivity within

the Hawaiian Archipelago. We also examine Hawai‘i’s connection with other

archipelagos in the Central Pacific in an attempt to understand historical population

dynamics, dispersal, and colonization in order to help delineate the appropriate scales for

coral reef management.

METHODS

Sampling, PCR, and Sequencing

Ophiocoma erinaceous and Ophiocoma pica were sampled from five regions

(Hawaiian Islands, Johnston Atoll, Line Islands, French Polynesia, and Saudi Arabia) at a

120

total of 20 sites from 19 islands, though both species were not found at every site (Table

5.1). Sampling in the Northwest Hawaiian Islands, Johnston Atoll, and the Line Islands took place on research cruises aboard the NOAA R.V. Hi`ialakai. All other samples were collected on shore dives or while snorkeling. Sampling took place between Spring 2006 and Fall 2011. Samples were obtained non-lethally -- through the removal and collection of arms which can be later regenerated -- preserved in either 95% ethanol or DMSO salt buffer, and archived at the Hawai‘i Institute of Marine Biology at room temperature.

Skillings and Toonen (2010) contains an extended discussion of sampling and preservation protocols.

Total genomic DNA was extracted using DNeasy™ Blood and Tissue Kits

(QIAGEN) or E.Z.N.A. Tissue DNA Kits (OMEGA Bio-tek) following the manufacturer’s instructions. Polymerase chain reaction (PCR) was used to amplify a fragment of the mitochondrial 16 S ribosomal DNA gene (16S). Primers 16SarL (5’ –

CGCCTGTTTATCAAAAACAT-3’) and 16SbrH (5’ –

CCGGTCTGAACTCAGATCACGT-3’) (Palumbi et al. 1991) were used to amplify a

~486 base pair fragment from O. erinaceous and a ~488 base pair fragment from O. pica.

PCR was performed using 20 µL reactions containing 6 µL of diluted DNA extract (one part template DNA to 199 parts nanopure water), 1 µL each of 5 µM forward and reverse primers, 0.5 µL of 0.5 µM BSA, 0.5 µL of 50 µM MgCL, 10 µL of Bioline (Bioline)

Biomix Red diluted as per manufacturer’s instructions, and 0.4 µL of nanopure water.

PCR was done on Bio-Rad Icycler™ thermocyclers (Bio-Rad Laboratories) with an initial denaturation at 95°C for 7 min followed by 40 cycles of a denaturing step at 95°C for 1 min, annealing at 51°C for 1 min, extension at 72°C for 1 min. A final extension at

121

72°C was held for 7 min before refrigeration. PCR products (8 µL) were treated with 0.7

µL of Exonuclease I combined with 0.7 µL of calf intestinal alkaline phosphatase (Exo-

CIAP) and incubated at 37°C for 30 minutes, with a final inactivation step at 85°C for 10

minutes. The treated PCR products were sequenced using an ABI Prism 3730 automatic

sequencer (Applied Biosystems, Foster City, CA, USA) at the Hawai‘i Institute of Marine

Biology’s EPSCoR sequencing facility. All samples were sequenced in the forward

direction; uncertain sequences and all unique haplotypes were also sequenced in the

reverse direction for confirmation. Sequences were compiled and trimmed using

Geneious Pro 5.4.6 (Biomatters, LTD, Auckland, NZ) and aligned using MAFFT 6.814b

(Katoh et al. 2002) implemented in Geneious.

DATA ANALYSIS

A statistical parsimony network of mtDNA haplotypes was constructed by creating a reduced median network that was then used to make a median joining network; both procedures implemented in Network 4.516 (www.fluxus-engineering.com; Bandelt et al., 1995; Bandelt et al., 1999). The network was drawn using Network Publisher

1.1.0.7 (www.fluxus-engineering.com) and edited using Adobe Illustrator.

Nei’s average pairwise genetic difference (π) (Nei & Li, 1979), haplotype diversity (h), Tajima’s D (Tajima, 1989), and Fu’s FS (Fu, 1997) were calculated in

DnaSP 4.1 (Rozas, 2003). The effective number of alleles (1/(1-h)) was calculated by hand following Jost (2008).

To assess genetic differentiation between sites we calculated pairwise FST and ΦST

values using Arlequin 3.1 (Excoffier et al., 2005). Pairwise Dest_chao values (Jost 2008)

122

were calculated using SPADE (Chao and Shen 2010). To correct the critical P value for

statistical significance in pairwise comparisons, the family-wise false discovery rate

(FDR) correction of Benjamini (2006) was implemented. Analysis of molecular variance

(AMOVA) was used for hierarchical analysis of the partitioning of 16S diversity among sites within archipelagic regions and among archipelagic regions using Arlequin. The

pairwise ΦST and AMOVA analyses were conducted using a distance matrix with 50,000

permutations and the Tamura-Nei mutational model (Tamura & Nei, 1993) with gamma =

0.096. The mutational model HKY+G was selected using AIC in Modeltest 3.7 (Posada

& Crandall, 1998); the model hierarchy was used to select the closest available model

when the best-fit model could not be implemented by the chosen program, as in the case

of Arlequin. Inferences are robust to the mutational model and our conclusions are not

altered regardless of which model is chosen.

Bayesian coalescent-based calculations of migration rate among regions (NeΜ )

and the region mutation parameter (θ) were obtained using IMa2 (Hey, 2010). A MCMC

chain with a length of 2,000,000 sampled every 20 generations with a 10% burn-in was used to estimate parameters. Region groupings were chosen based on AMOVA, median-

joining network, and pair-wise comparison results. Values for the migration rate among

regions (Μ) and region mutation parameters (θ) were taken from the highest peaks in the

posterior probability distribution curves. The posterior probability distributions were

examined to determine the validity of each estimated parameter.

123

RESULTS

Ophiocoma erinaceous

In 400 individuals we observed 127 haplotypes, 75 of which were only observed in one locality. Due to low sample size, Kaula Rock (N = 1), Gardner Pinnacles (N = 1), and Saudi Arabia (N = 3) were excluded from all population analyses. The number of individuals (N), number of haplotypes (H), number of unique haplotypes at site (Hu), nucleotide diversity (π), haplotype diversity (h), and effective number of alleles (AE) are reported in the upper half of Table 1. Overall nucleotide diversity was moderate (π =

0.0629 ± 0.0026) while the corresponding haplotype diversity was high (h = 0.93 ± 0.01).

Across all sites nucleotide diversity varied by an order of magnitude, from π = 0.0040 at

Palmyra Atoll to π = 0.0826 at Pearl & Hermes Reef. Haplotype diversity ranged from h

= 0.42 at Midway Atoll to h = 0.95 at Johnston Atoll. There were no haplotypes shared across all sites.

The Main Hawaiian Islands exhibited the highest haplotype diversity (h = 0.91) and effective number of haplotypes (AE = 12.6) along with Johnston Atoll (h = 0.95, AE =

20.0). Palmyra (h = 0.85, AE = 6.7), French Polynesia (h = 0.83, AE = 5.9) and the

Northwest Hawaiian Islands (h = 0.79, AE = 5.6) were all characterized by lower overall haplotype diversity and effective number of haplotypes. Nucleotide diversity does not appear to be correlated with haplotype diversity. The median joining network is characterized by five common haplotypes in three groups separated 46 and 65 mutational steps (Figure 1). One group consists primarily of haplotypes found in Palmyra Atoll,

French Polynesia, and Saudi Arabia (N = 3); a second group consist primarily of

124

haplotype localized to the Hawaiian Archipelago. The haplotypes found at Johnston Atoll are split between these two groups separated by 46 steps. The third group is separated by

65 steps and localized to the three most Northwestern Hawaiian Islands with the exception of one brittle star from Nihoa Island.

Four AMOVAs were run on the O. erinaceous 16S haplotype data (Table 5.2).

Sampling sites were grouped by Archipelago. 1) Previous studies have shown significant

population differentiation between islands in the Northwest Hawaiian Islands and the

Main Hawaiian Islands in multiple species (Skillings et al. 2011; Toonen et al. 2011). 2)

The Hawaiian Archipelago into these two general regions, as well as pulling out Johnston

Atoll. 3 + 4) The two regional AMOVAs were repeated differing only in the measure

used to determine genetic diversity; FST, or strict haplotype counts, and ΦST, which incorporates genetic distance between haplotypes. Both measures displayed a similar partitioning of among-population-within-region variance, but the among-group partitioning of variance was 4-7 times higher using ΦST. Overall, all tests were

significant (P < 0.05), indicating significant genetic partitioning between regions. The

AMOVA that explained the most among-group variance used the MHI, the three most

northwestern NWHI, the lower NWHI, Johnston Atoll, Palmyra Atoll, and French

Polynesia as the groupings.

There is a strong pattern of restricted gene flow between sites among the

population samples of O. erinaceous, overall Morisita nearly unbiased dissimilarity

between sites was 0.595. Pairwise comparisons for both ΦST and Dest_chao reveal

significant differences, the primary difference being more significant population structure

revealed by ΦST (Table 5.3). There were significant differences between sites located in

125

different archipelagos, the one exception being between French Polynesia and Palmyra

Atoll in the Line Islands south of Hawai‘i. The most northwestern Hawaiian Islands,

Kure Atoll, Midway Atoll, and Pearl & Hermes Atoll, were significantly different from

almost every other locality. Oahu and Hilo (Hawai‘i Island) were also significantly

different than most of the other sites, especially with ΦST. All other sites within the

Hawaiian Island chain were not significantly partitioned.

Ophiocoma pica

In 467 specimens we observed 112 haplotypes, 83 of which were only observed in

one locality. Due to low sample size, Oahu (n=2) and Gardner Pinnacles (n=1) were

excluded from population analyses. The number of individuals (N), number of

haplotypes (H), number of unique haplotypes at site (Hu), nucleotide diversity (π),

haplotype diversity (h), and effective number of alleles (AE) can be found in the lower half of Table 1. Overall nucleotide diversity was low (π = 0.0082 ± 0.0007) while the corresponding haplotype diversity was high, but lower than O. erinaceous (h = 0.85 ±

0.01). Across all sites nucleotide diversity varied by an order of magnitude, from π =

0.0035 at Hilo on the Big Island of Hawai‘i to π = 0.0114 at French Polynesia. Haplotype diversity ranged from h = 0.70 at Laysan Island to h = 0.93 at Maui. There were no haplotypes shared across all sites.

The highest haplotype diversity and effective number of haplotypes were found in

French Polynesia (h=0.92, AE=12.5) and Palmyra (h=0.91, AE=11.1). Haplotype diversity

and effective number of haplotypes were not substantially lower in the MHI (h=0.87,

AE=8.6) but continued to drop off in the NWHI (h=0.82, AE=6.6) and Johnston Atoll

126

(h=0.79, AE=4.8). The median joining network is characterized by five common haplotypes in two groups separated by 12 mutational steps (Fig. 5.2). Haplotypes appear

to be distributed fairly evenly between sites.

Six AMOVA were run on the O. pica 16S haplotype data, three regional

groupings tested using FST and ΦST (Table 5.2). 1) Sampling sites were first grouped by

Archipelago, 2) the Hawaiian Archipelago split into the NWHI and the MHI, and 3)

Johnston Atoll grouped with the Hawaiian Archipelago. Calculating variability using ΦST

consistently accounted for a greater proportion of among-group variance. Grouping

Hawai‘i and Johnston Atoll together explained the most among-group variance (43.33%) but at P-value of 0.08. Pulling Johnston Atoll out of Hawai‘i explained 36.29% of the

among-group variance at a P-value of 0.03.

There is less pronounced population structuring between sites in O. pica

compared to O. erinaceous, overall nearly unbiased dissimilarity between sites was

0.200. Pairwise comparisons for both ΦST and Dest_chao reveal significant differences

(Table 5.3). Both measures revealed significant population differences between Palmyra

Atoll and French Polynesia and between those two sites and all other sites; although not

every pair-wise comparison was significant, there was a significant difference between

Palmyra Atoll, French Polynesia and all other sites in at least one of the measures. Within

the Hawaiian Archipelago the only significant structure measured using Dest_chao were

comparisons between Laysan and other sites, and between Kona and other sites. There

were no clear patterns measured by ΦST; Maro Reef was significantly different from three

other sites within the Hawaiian Islands, as was Hilo.

127

DISCUSSION

Ophiocoma pica and O. erinaceous are two widely distributed brittle stars with

similar life histories. In this survey of population genetic structure we elucidate patterns

of connectivity in three central Pacific archipelagos with a focus on the Hawaiian

Archipelago. We were interested in illuminating aspects of marine phylogeography in

these two species in order to test the assumption that a single species can be used as a proxy for a group of closely related species to estimate dispersal among marine communities. Better estimates of marine dispersal will contribute to the management of marine communities.

The Hawaiian Archipelago is highly isolated, and also contains one of the highest proportions of endemism in the world (e.g., Briggs 1974; Kay 1980; Grigg 1983; reviewed by Ziegler 2002; Eldredge & Evenhuis, 2003). Though there are many pan- pacific marine organisms in Hawai‘i, the isolation of the archipelago is thought to limit larval exchange such that colonization is rare but sufficient to maintain species cohesion among these taxa. The Hawaiian marine fauna contains a large proportion of endemics that are differentiated but not diversified from its Indo-West Pacific roots (Kay &

Palumbi 1987; Hourigan & Reese 1987; Jokiel 1987; Ziegler 2002). In this scenario

Hawai‘i is seen primarily as a dead-end, an isolated land-mass that does not contribute in a significant way to the overall diversity of the tropical Pacific. An alternative to the island biogeography hypotheses of Hawaiian diversity is the Vortex model of speciation

(Jokiel and Martinelli, 1992), wherein the stunning biodiversity of the Coral Triangle is a

result of centrifugal accumulation of species from the peripheral habitats around the

Pacific. Though these two models primarily make predictions about speciation-level

128

processes and do not speak directly to gene-flow within a species, they do make opposite

claims about the dominant direction of gene-flow and dispersal.

Despite having similar ecological niches and sympatric distributions across the

samples portion of the range, the congeners O. erinaceous and O. pica have responded

differently to the Pacific archipelagic environments. There is a strong disparity in

phylogeographic structure within the Hawaiian Islands. A similar phylogeographic

pattern was found between Archipelagos, but the mutational distances between

haplotypes differed greatly. A similar pattern was found between the congeneric sea

cucumbers Holothuria atra and Holothuria whitmaei within the Hawaiian Archipelago

(Chapter 3).

Phylogeography and Population Expansion

The median-joining networks reconstructed two distinct haplotypic patterns. Both

networks have very common haplotypes with multiple star-like radiations. The haplotypes in the O. pica network were all closely linked, except for relatively rare grouping separated by 12 mutational steps. The haplotypes were also distributed throughout the sampled range with no strong geographic patterns, though there were frequency differences between regional sites. The O. erinaceous network was substantially more structured. Three major groups, separated by 46 and 65 mutational steps, were recovered. One group was primarily localized to French Polynesia, Palmyra, and Johnston Atoll; though, haplotypes from this group were found at low frequencies in the MHI. The second group, 46 mutational steps away, was primarily localized to the

MHI, Johnston Atoll, and the first few NWHI. The individuals sampled from Johnston

129

Atoll contained almost equal proportions of haplotypes from groups 1 and 2. The third

group, separated from the second by 65 mutational steps, was found almost exclusively in

the three most northwestern Hawaiian Islands; only one sampled individual from Nihoa was recovered with a haplotype from this group.

The star-like patterns in both median-joining networks indicate non-equilibrium dynamics (Grosberg & Cunningham 2001; Crandall et al. 2008). This is supported by significantly negative regional Fu’s FS and Tajima’s D values that could indicate a

demographic expansion, selective sweep, or background selection. It is possible that

positive selection could have produced an advantageous variant which then swept to

fixation (Gillespie 2001). Such an event would diminish haplotype diversity, leaving a

single central haplotype rather than high diversity and multiple simultaneous star

expansions, as seen in both species (Crandall et al. 2008). It seems unlikely that two taxa

would experience contemporaneous selective sweeps across the sampled range.

A perhaps more parsimonious explanation of the nonequilibrium patterns would

be population expansion after the colonization of an archipelago that has not yet reached

equilibrium. Given that regional pairwise structure indicates relatively little secondary

contact, it seems likely that both species would show signatures of population expansion

after colonization of a new archipelago; an event that might stochastically favor a few

haplotypes and form the characteristic star-shapes (Slatkin & Hudson 1991).

Population genetic structure between regions and sites

Significant regional population partitioning was detected in both species with

AMOVA, but regional patterns of differentiation were not significant. Pairwise

130 comparisons revealed gene-flow restrictions between French Polynesia, Palmyra Atoll, and the Hawaiian Archipelago plus Johnston Atoll. These were the only consistent pairwise gene-flow restrictions between O. pica sites. These restrictions were much stronger in O. erinaceous, the species with much finer population structuring.

There was almost no dissimilarity, measured by Jost’s D, between Johnston Atoll and the rest of the Hawaiian Archipelago in O. pica. The NWHI and Johnston Atoll appear to form one population. The most southeastern MHI were patchily dissimilar from the NWHI, with Kona being dissimilar from almost all NWHI sites and the Hilo location on the opposite side of the same island. Within the Hawaiian Islands three distinct groups were found in O. erinaceous. The three most northwestern NWHI (Kure Atoll, Midway

Atoll, Pearl & Hermes Atoll); the southeastern most NWHI and the MHI; and Johnston

Atoll. The differences in the scale of population differentiation suggest that factors beyond merely geographic distance influence population partitioning.

It has been suggested that Johnston Atoll acts as a stepping-stone into the

Hawaiian Islands (Maragos & Jokiel, 1986). Kobayashi (Kobayashi & Polovina, 2006;

Kobayashi, 2006) used computer simulations to predict two larval transport corridors from Johnston Atoll to the Hawaiian Archipelago: one corridor stretching from Johnston to French Frigate Shoals in the NWHI, and one from Johnston to O‘ahu in the MHI. Our data give more support to the predicted larval transport corridor between Johnston Atoll and French Frigate Shoals than the corridor predicted between Kaua‘i and Johnston.

The migration rates between regions that were calculated for O. pica indicate that

Hawai‘i is more isolated from the Palmyra and French Polynesia than those two regions

131

are from each other. A substantial number of migrants are exchanged between Palmyra

and French Polynesia every generation. There is one-way migration from Palmyra to the

Hawaiian Archipelago and Johnston Atoll, and one-way migration from the Hawaiian

Archipelago and Johnston Atoll to French Polynesia. Both of these migration rates were

substantially lower than between Palmyra and French Polynesia.

There was a substantial amount of migration from south to north in O. erinaceous

based on our admittedly low sampling. The dominant migration route was from French

Polynesia to Palmyra Atoll, a one-way conveyor belt from Palmyra Atoll and French

Polynesia to Johnston Atoll. From there migration dominated in the direction of the MHI up the chain to the northwest and feedback from the MHI back to Johnston Atoll. This pattern can also be seen the median-joining network. The effectively one-way migration rates into the NWHI and Johnston Atoll coupled with the strong genetic similarity between Johnston Atoll and the NWHI indicate Johnston Atoll is an isolated outpost of the Northwest Hawaiian Islands that occasionally acts as a stepping stone into Hawai‘i. It

appears that migrants can make it to Johnston Atoll from the north and south, but rarely

successfully disperse from Johnston Atoll.

CONCLUSIONS

The Hawaiian Archipelago lies at the periphery of the tropical Central Pacific and is the most isolated island chain in the world; the question remains as to why some species maintain connectivity and species cohesion between the Hawaiian Islands and the rest of the Pacific, why some species diverge and become Hawaiian endemics, and why

other species with similar inferred dispersal ability fail to colonize the Hawaiian

132

Archipelago at all. While these Hawaiian Ophiocoma exhibit different population genetic patterns, they nonetheless show significant structure in both species.

Considerable evidence is accumulating that it is indefensible to make predictions of connectivity based solely on proxies such as ecological or phylogenetic similarity, pelagic larval duration, or species range sizes (Bird et al., 2007; Lester et al., 2007;

Bradbury et al., 2008; Shanks, 2009; Weersing & Toonen, 2009; Skillings et al. 2011;

Toonen et al. 2011). The fine-scale structuring of populations in O. erinaceous and O. pica suggest that a species-specific approach is needed for responding to the complex relationships between genetically distinct populations. These species must be managed on a local scale; migration between archipelagos, and often between islands, does not occur in ecologically relevant time frames.

REFERENCES

Bandelt HJ, Forster P, Sykes BC, Richards MB (1995) Mitochondrial portraits of human populations using median networks. Genetics. 141:743-753.

Bandelt HJ, Foster P, Röhl A (1999) Median-joining networks for inferring intraspecific phylogenies. Molecular Biology and Evolution. 16(1):37-48.

Barber PH, Palumbi SR, Erdmann MV, Moosa MK (2002) Sharp genetic breaks among populations of Haptosquilla pulchella (Stomatopoda) indicate limits larval transport: patterns, causes, and consequences. Molecular Ecology. 659-674.

Baums IB, Paris CB, Cherubin LM (2006) A bio-oceanographic filter to larval dispersal in a reef-building coral. Limnology and Oceanography. 51:1969–1981.

133

Beerli P (2006) Comparison of Bayesian and maximum likelihood inference of population genetic parameters. Bioinformatics. 22(3):341-345.

Benjamini Y, Krieger AM, Yekutieli D (2006) Adaptive linear step-up procedures that control the false discovery rate. Biometrika. 93(3):491-507.

Bird CE, Holland BS, Bowen BW, Toonen RJ (2007) Contrasting phylogeography in three endemic Hawaiian limpets (Cellana spp.) with similar life histories. Molecular

Ecology. 16(15):3173-3186.

Bonham K, Held EE (1963) Ecological observations on the sea cucumbers Holothuria atra and H. leucospilota at Rongelap Atoll, Marshall Islands. Pacific Science. 17:305-

314.

Bradbury IR, Bentzen P (2007) Non-linear genetic isolation by distance: implications for dispersal estimation in anadromous and marine fish populations. Marine Ecology

Progress Series 340:245-257.

Bradbury IR, Laurel B, Snelgrove PVR, Bentzen P, Campana SE (2008) Global patterns in marine dispersal estimates: the influence of geography, taxonomic category and life history. Proceedings of the Royal Society of London B. 275:1803-1810.

Briggs JC (1974) Marine Zoogeography. McGraw Hill Publishers, New York, NY.

Burton R, Feldman M (1981) Population genetics of coastal and estuarine invertebrates: are behavioral differences among species reflected in population structure? Estuaries.

4:239.

134

Byrne M, Rowe F, Uthicke S (2010) Molecular taxonomy, phylogeny and evolution in

the family Stichopodidae (Aspidochirotida: Holothuroidea) based on COI and 16S

mitochondrial DNA. Molecular Phylogenetics and Evolution. 56:1068-1081.

Chao, A. and Shen, T.-J. (2010) Program SPADE (Species Prediction And Diversity

Estimation. http://chao.stat.nthu.edu.tw

Clark AM Rowe FEW (1971) Monograph of shallow water Indo-West Pacific

echinoderms. Trustees of the British Museum (Natural History), London.

Conand C (1994) Asexual reproduction by fission in Holothuria atra: variability of some

parameters in populations from the tropical Indo-Pacific. Oceanologica acta. 19(4):209-

216.

Craig MT, Eble JA, Bowen BW, Robertson DR (2007) High genetic connectivity across the Indian and Pacific Oceans in the reef fish Myripristes berndti (Holocentridae). Mar

Ecol Prog Ser. 334:245-254.

Crandall ED, Frey MA, Grosberg RK, Barber PH (2008) Contrasting demographic history and phylogeographical patterns in two Indo-Pacific gastropods. Molecular

Ecology. 17:611-626.

Dawson MN (2001) Phylogeography in coastal marine animals: solution from California?

Journal of Biogeography. 28:723-736.

135

Dawson MN, Hamner WM (2008) A biophysical perspective on dispersal and the geography of evolution in marine and terrestrial systems. Journal of the Royal Society

Interface. 5:135-150.

Drummond AJ, Nicholls GK, Rodrigo AG, Solomon W (2002) Estimating mutation parameters, population history and genealogy simultaneously from temporally spaced sequence data. Genetics. 161:1307-1320.

Drummond AJ (2003) BEAST Version 1.5.4. Available from http://code.google.com/p/beast-mcmc/downloads/list.

Drummond AJ, Ashton B, Buxton S, Cheung M, Cooper A, Duran C, Field M, Heled J,

Kearse M, Markowitz S, Moir R, Stones-Havas S, Sturrock S, Thierer T, Wilson A (2011)

Geneious v5.4.6

Dupanloup I, Schneider, Excoffier L (2002) A simulated annealing approach to define the genetic structure of populations. Molecular Ecology. 11(12):2571-2581.

Ebert TA (1978) Growth and size of the tropical sea cucumber Holothuria (Halodeima) atra Jager at Enewetak Atoll, Marshall Islands. Pacific Science 32:183-191.

Ebert TA (1983) Recruitment in Echinoderms. In: Echinoderm Studies 1. Jangoux M,

Lawrence JM (Eds.).

Eble JA, Toonen RJ, Bowen BW (2009) Endemism and dispersal: comparative phylogeography of three surgeonfishes across the Hawaiian Archipelago. Marine

Biology. 156(4):689-698.

136

Eldredge LG, Evenhuis NL (2003) Hawai‘i’s biodiversity: a detailed assessment of the

numbers of species in the Hawaiian Islands. Bishop Museum Occasional Papers. 76:1-28.

Excoffier L, Laval G, Schneider S (2005) Arlequin version 3.0: an integrated software package for population genetics data analysis. Evolutionary Bioinformatics Online. 1:47-

50.

Fogarty MJ, Botsford LW (2007) Population connectivity and spatial management of marine fisheries. Oceanography. 20(3):112-123.

Fu Y-X (1997) Statistical tests of neutrality of mutations against population growth, hitchhiking, and background selection. Genetics. 147:915-925.

Gaither MR, Toonen RJ, Robertson DR, Planes S, Bowen BW(2010) Genetic evaluation of marine biogeographical barriers: perspectives from two widespread Indo-Pacific snappers (Lutjanus kasmira and Lutjanus fulvus). Journal of Biogeography. 37(1):133-

147.

Gilman SE (2006) Life at the edge: an experimental study of poleward range boundary.

Oecologia. 148:270-279.

Grigg RW (1983) Community structure, succession and development of coral reefs in

Hawai‘i. Marine Ecology Progress Series. 11:1-14.

Grosberg RK, Cunningham CW (2001) Genetic Structure in the Sea: From Populations to

Communites. In: Marine Community Ecology. Bertness MD, Gaines SD, Hay ME (Eds.).

Sinauer Associates, Inc., Sunderland, MA. 61-84.

137

Hall TA (1999) BioEdit: a user-friendly biological sequence alignment editor and analysis

program for Windows 95/98/NT. Nucleic Acids Symposium Series. 41:95-98.

Hedgecock D, Barber PH, Edmands S (2007) Genetic approaches to measuring

connectivity. Oceanography. 20(3):70-79.

Hellberg ME (2009) Gene flow and isolation among populations of marine animals.

Annual Review of Ecology, Evolution and Systematics. 40:291-310.

Hey, J (2010) Isolation with Migration Models for More Than Two Populations.

Molecular Biology and Evolution. 27:905-920.

Hourigan TF, Reese ES (1987) Mid-ocean isolation and evolution of Hawaiian reef

fishes. Trends in Ecology and Evolution. 2:187-191.

Hutchison DW, Templeton A (1999) Correlation of pairwise genetic and geographic

distance measures: inferring the relative influences of gene flow and drift on the

distribution of genetic variability. Evolution. 53:1898-1914.

Jokiel PL (1987) Ecology, biogeography and evolution of corals in Hawai;i. Trends in

Ecology & Evolution. 2:179-182.

Jokiel PL, Martinelli FJ (1992) The Vortex Model of Coral Reef Biogeography. Journal

of Biogeography, 19(4):449-458.

Jost L (2008) Gst and its relatives do not measure differentiation. Molecular Ecology.

17(18):4015-4026.

138

Kappel CV, Martone RG, Duffy JE (2006) "Ecosystem-based management." In:

Encyclopedia of Earth. Cleveland CJ (Ed.). Washington, D.C.: Environmental

Information Coalition, National Council for Science and the Environment.

Katoh, Misawa, Kuma, Miyata (2002) MAFFT: a novel method for rapid multiple sequence alignment based on fast Fourier transform. Nucleic Acids Research. 30:3059-

3066

Kay EA (1980) Little worlds of the Pacific. An essay on Pacific Basin Biogeography.

Harold L. Lyon Arboretum, University of Hawai‘i, Lecture No. 9:1-40.

Kay EA (1984) Patterns of speciation in the Indo-West Pacific. In: Radovsky FJ, Raven

PH, Sohmer SH (Eds.). Biogeography of the Tropical Pacific. Association of Systematics

Collections and B.P. Bishop Museum. 15-31.

Kay EA, Palumbi SR (1987) Endemism and evolution in Hawaiian marine invertebrates.

Trends in Ecology & Evolution. 2(7):183-186.

Kinlan BP, Gaines SD (2003) Propagule dispersal in marine and terrestrial environments: a community perspective. Ecology. 84:2007-2020.

Kinlan BP, Gaines SD , Lester SE (2005) Propagule dispersal and the scales of marine community process. Diversity & Distribution. 11:139–148

Kirkendale LA and Meyer CP (2004) Phylogeography of the Patelloida profunda group

(Gastropoda: Lottidae): diversification in a dispersal-driven marine system. Molecular

Ecology. 13:2749-2762.

139

Knowlton N, Keller B (1986) Larvae which fall short of their potential: highly localized

recruitment in an alpheid shrimp with extended larval development. Bulletin of Marine

Science. 39:213-223.

Kobayashi DR (2006) Colonization of the Hawaiian Archipelago via Johnston Atoll: A characterization of oceanographic transport corridors for pelagic larvae using computer

simulation. Coral Reefs. 25:407–417.

Kobayashi DR, Polovina JJ (2006) Simulated seasonal and interannual variability in

larval transport and oceanography in the Northwestern Hawaiian Islands using satellite

remotely sensed data and computer modeling. Atoll Research Bulletin. 543:365–390.

Laxminarayana A (2005) Induced spawning and larval rearing of the sea cucumbers,

Bohadschia marmorata and Holothuria atra in Mauritius. Fisheries Research. 48-52.

Lee J, Byrne M, Uthicke S (2008) The influence of population density on fission and

growth of Holothuria atra in natural mesocosms. Journal of Experimental Marine

Biology and Ecology. 365: 126-135

Lessios HA, Garrido M, Kessing BD (2001) Demographic history of Diadema

antillarum, a keystone herbivore on Caribbean reefs. Proceedings of the Royal Society.

268:1-7.

Lessios HA, Kane J, Robertson DR (2003) Phylogeography of the pantropical sea urchin

Tripneustes: contrasting patterns of population structure between oceans. Evolution.

57(9):2026-2036.

140

Lester SE, Ruttenberg BI, Gaines SD, Kinlan BP (2007) The relationship between

dispersal ability and geographic range size. Ecology Letters. 10:745-758.

Levin L (2006) Recent progress in understanding larval dispersal: new directions and

digressions. Integrative and Comparative Biology. 46(3):282-297.

Maragos JE, Jokiel PL (1986) Reef corals of Johnston Atoll: one of the world's most

isolated reefs. Coral Reefs. 4(3):141–150

Manni F, Guerard E, Heyer E (2004) Geographic patterns of (genetic, morphologic,

linguistic) variation: how barriers can be detected by “Monmonier’s algorithm”. Human

Biology. 76(2):173-190.

McCleod K, Leslie H (eds.) (2009) Ecosystem-based management for the Oceans. Island

Press, Washington DC, USA. 392 pp.

Miller MA, Holder MT, Vos R, Midford PE, Liebowitz T, Chan L, Hoover P, Warnow T

(2010) The CIPRES Portals. http://www.phylo.org/sub_sections/portal

Myers R, Worm B (2003) Rapid worldwide depletion of predatory fish communities.

Nature. Vol. 423: 280-283.

Nei M, Li WM (1979) Mathematical model for studying genetic variation in terms of restriction endonucleases. Proceedings of the National Academy of Sciences of the USA.

11:291-306.

Paine RT (1969) A note on trophic complexity and community stability. Am. Nat. 103:

91-93.

141

Paulay G, Meyer C (2006) Dispersal and divergence across the greatest ocean region: Do

larvae matter? Integrative and Comparative Biology. 46(3):269-281.

Pauly D, Christensen V, Dalsgaard J, Froese R, Torres Jr. F (1998) Fishing Down Marine

Food Webs. Science. 279(5352):860 - 863

Pauly D, Watson R (2003) Counting the last fish. Scientific America. 289(1):42-7.

Posada D, Crandall KA (1998) Modeltest: testing the model of DNA substitution.

Bioinformatics. 14:817-818.

Power ME, Tilman D, Estes JA, Menge BA, Bond WJ, Mills LS, Daily G, Castilla JC,

Lubchenco J, Paine RT (1996) Challenges in the Quest for Keystones. BioScience. 46:

609-620.

Purcell SW (2010) Managing sea cucumber fisheries with an ecosystem approach. In:

FAO Fisheries and Aquaculture Technical Paper No. 520. Lovatelli A, Vasconcellos M,

Yimin Y (Eds.). Rome, FAO. 157pp.

Rocha LA, Bowen BW (2008) Speciation in coral-reef fishes. Journal of Fish Biology.

72(5):1101-1121.

Ronquist F, Huelsenbeck JP (2003) MRBAYES 3: Bayesian phylogenetic inference under

mixed models. Bioinformatics. 19:1572-1574.

Ross PM, Hogg ID, Pilditch CA, Lundquist CJ (2009) Phylogeography of New

Zealand’s coastal benthos. New Zealand Journal of Marine Freshwater Research. 43:

1009-1027.

142

Rozas J (2003) DnaSP, DNA polymorphism analyses by the coalescent and other

methods. Bioinformatics. 19(18):2496-2497.

Rozen S, Skaletsky H (2000) Primer3 on the WWW for general users and for biologist programmers. In: Krawetz S, Misener S (eds) Bioinformatics Methods and Protocols:

Methods in Molecular Biology. Humana Press, Totowa, NJ. pp 365-386.

Sala E, Boudouresque CF, Harmelin-Vivien M (1998) Fishing, Trophic Cascades, and the

Structure of Algal Assemblages: Evaluation of an Old but Untested Paradigm. Oikos. 82:

425-43.

Sanchirico JN, Wilen JE (2005) Optimal spatial management of renewable resources:

Matching policy scope to ecosystem scale. Journal of Environmental Economics &

Management. 50:23-46.

Schultz JK, Pyle RL, DeMartini E, Bowen BW (2007) Genetic connectivity among color morphs and Pacific archipelagos for the flame angelfish, Centropyga loriculus. Marine

Biology. 151:167-175.

Selkoe KA, Henzler C, Gaines SD (2009) Seascape genetics and the spatial ecology of

marine populations. Fish and Fisheries. 9:363-377.

Severance E, Karl S. Contrasting population genetic structures of sympatric, mass-

spawning Caribbean corals. Marine Biology. 150:57-68.

Shanks A, Grantham B, Carr M (2003) Propagule dispersal distance and the size and

spacing of marine reserves. Ecological Applications. 13:S159-S169.

143

Shanks AL (2009) Pelagic Larval Duration and Dispersal Distance Revisited. Biol. Bull.

216:373-385.

Skillings DJ and Toonen RJ (2010) It’s just a flesh wound: non-lethal sampling for

conservation genetics studies. Proceedings of the American Academy of Underwater

Sciences In: NW Pollock (ed). Diving for Science. Proceedings of the 29th American

Academy of Underwater Sciences Symposium. Dauphin Island, AL: AAUS; In press.

Slatkin M (1987) Gene flow and the geographic structure of natural populations. Science.

236: 787-792.

Slatkin M, Barton N (1989) A comparison of three indirect methods for estimating

average levels of gene flow. Evolution. 43:1349-1368.

Sloan, SA (1984) Echinoderm fisheries of the world: A review. In: Echinodermata:

Proceedings of the Fifth International Echinoderm Conference. Keegan BF, O’Connor

BDS (Eds.). A.A. Balkema, Rotterdam, Netherlands. 109-124.

Sotka EE, Wares JP, Barth JA, Grosberg RK, Palumbi SR (2004) Strong genetic clines and geographical variation in gene flow in the rocky intertidal barnacle Belanus glandula.

Molecular Ecology. 13:2143-2156.

Stamatakis A (2006) RAxML-VI-HPC: maximum likelihood-based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics. 22(21):2688-2690.

Strathmann RR (1993) Hypotheses on the origins of marine larvae. Annual Review of

Ecology and Systematics. 24:89-117.

144

Tajima F (1989) Statistical method for testing the neutral mutation hypothesis by DNA

polymorphism. Genetics. 125:585-593.

Tamura K, Nei M (1993) Estimation of the number of nucleotide substitutions in the control region of mitochondrial DNA in humans and chimpanzees. Molecular Biology

and Evolution. 10:512-26.

Tavaré S (1986) Some probabilistic and statistical problems in the analysis of DNA

sequences. Lectures on Mathematics in the Life Sciences. American Mathematical

Society. 17:57-86.

Thompson JD, Higgins DG, Gibson TJ (1994) ClustalW: improving the sensitivity of

progressive multiple sequence alignment through sequence weighting, position-specific

gap penalties and weight matrix choice. Nucleic Acids Research. 22:4673-4680.

Thorrold SR, Zacherl DC, Levin LA (2007) Population connectivity and larval dispersal using geochemical signatures in calcified structures. Oceanography. 20(3):80-89.

Thorson G (1950) Reproductive and larval ecology of marine bottom invertebrates. Biol

Rev Camb Philos Soc. 25:1-45.

Treml EA, Halpin PN, Urban DL, Pratson LF (2008) Modeling population connectivity by ocean currents, a graph-theoretic approach for marine conservation. Landscape

Ecology. 23:19-36.

Uthicke S (1999) Sediment bioturbation and impact of feeding activity of Holothuria

(Halodeima) atra and Stichopus chloronotus, two sediment feeding holothurians, at

Lizard Island, Great Barrier Reef. Bulletin of Marine Science 64:129-141.

145

Uthicke S, Schaffelke B, Byrne M (2009) A boom-bust phylum? Ecological and

evolutionary consequences of density variations in echonoderms. Ecological

Monographs. 79:3-24.

Weersing KA, Toonen RJ (2009) Population genetics, larval dispersal, and demographic

connectivity in marine systems. Mar. Ecol. Prog. Ser. 393:1-12 .

White C, Selkoe KA, Watson J, Siegel DA, Zacherl DC, Toonen RJ (2010) Ocean

currents help explain population genetic structure. Proc. Roy. Soc. B: Biol. Sci. 277:

1685-1694.

Worm B, Barbier EB, Beaumont N, Duffy JE, Folke C, Halpern BS, Jackson JBC, Lotze

HK, Micheli F, Palumbi SR, Sala E, Selkoe KA, Stachowicz JJ, Watson R (2006) Impacts

of Biodiversity Loss on Ocean Ecosystem Services. Science 314: 787 - 790.

Ziegler AC (2002) Hawaiian natural history, ecology & evolution. University of Hawai‘i

Press, Honolulu HI, USA. 477 p.

146

Ophiocoma erinaceous

Region Site N H H u π ± SD h ± SD A E Tajima's D Fu's Fs Hilo 26 15 7 0.0055 ± 0.0034 0.93 ± 0.03 14.3 -1.21 -8.84 Kona 17 11 5 0.0506 ± 0.0262 0.93 ± 0.04 14.3 1.33 2.68 Main Maui 40 21 10 0.0419 ± 0.0210 0.93 ± 0.02 14.3 0.97 1.03 Hawaiian Oahu 36 17 7 0.0063 ± 0.0038 0.93 ± 0.02 14.3 -1.16 -8.09 Islands Kauai 38 17 6 0.0393 ± 0.0198 0.82 ± 0.06 5.6 0.84 2.93 Kaula Rock 1 1 1 N/A N/A N/A N/A N/A Nihoa 12 8 2 0.0687 ± 0.0363 0.85 ± 0.10 6.7 -0.31 4.33 Gardner 1 1 0 N/A N/A N/A N/A N/A Northwest Maro Reef 42 19 7 0.0397 ± 0.0199 0.89 ± 0.03 9.1 1.05 2.33 Hawaiian Laysan 19 9 6 0.0451 ± 0.0232 0.81 ± 0.08 5.3 1.41 5.95 Islands Pearl & Hermes 6 5 3 0.0826 ± 0.0485 0.93 ± 0.12 14.3 1.13 3.17 Midway 9 3 1 0.0421 ± 0.0234 0.42 ± 0.19 1.7 -2.04 11.00 Kure 32 12 7 0.0762 ± 0.0379 0.82 ± 0.05 5.6 3.13 13.22 Johnston 49 25 13 0.0557 ± 0.0275 0.95 ± 0.02 20.0 2.68 1.92 Palmyra 31 17 13 0.0040 ± 0.0026 0.85 ± 0.06 6.7 -2.16 -14.26 French Polynesia 38 22 14 0.0086 ± 0.0049 0.83 ± 0.06 5.9 -2.65 -11.94 Saudi Arabia 3 3 3 0.0750 ± 0.0568 N/A N/A N/A N/A Overall 400 127 75 0.0629 ± 0.0026 0.93 ± 0.01 14.3 0.28 -0.89

Ophiocoma pica

Region Site N H H u π ± SD h ± SD A E Tajima's D Fu's Fs Hilo 23 10 5 0.0035 ± 0.0024 0.85 ± 0.05 6.7 -1.47 -4.86 Kona 25 13 6 0.0062 ± 0.0037 0.88 ± 0.05 8.3 -2.07 -5.13 Main Maui 10 7 1 0.0038 ± 0.0027 0.93 ± 0.06 14.3 -0.54 -3.54 Hawaiian Oahu 2 2 0 0.0307 ± 0.0317 N/A N/A N/A N/A Islands Kauai 26 13 4 0.0104 ± 0.0058 0.82 ± 0.07 5.6 -0.42 -2.20 Kaula Rock 12 6 1 0.0172 ± 0.0096 0.88 ± 0.06 8.3 1.16 2.54 Nihoa 10 8 2 0.0143 ± 0.0084 0.93 ± 0.08 14.3 -0.03 -1.12 French Frigate 45 14 6 0.0082 ± 0.0047 0.77 ± 0.05 4.3 -0.78 -1.99 Gardner 1 1 0 N/A N/A N/A N/A N/A Northwest Maro Reef 43 19 8 0.0036 ± 0.0024 0.85 ± 0.05 6.7 -2.16 -16.00 Hawaiian Lisianski 25 13 7 0.0081 ± 0.0047 0.86 ± 0.06 7.1 -1.66 -3.63 Islands Laysan 39 13 6 0.0053 ± 0.0032 0.70 ± 0.79 3.3 -1.92 -4.07 Pearl & Hermes 42 15 6 0.0055 ± 0.0033 0.78 ± 0.06 4.5 -1.92 -5.56 Midway 53 21 8 0.0074 ± 0.0042 0.85 ± 0.04 6.7 -1.37 -8.76 Kure 49 17 6 0.0046 ± 0.0029 0.83 ± 0.05 5.9 -2.23 -8.81 Johnston 15 8 1 0.0095 ± 0.0055 0.79 ± 0.10 4.8 -0.83 -0.39 Palmyra 34 16 11 0.0100 ± 0.0056 0.91 ± 0.03 11.1 -1.53 -3.70 French Polynesia 9 7 5 0.0114 ± 0.0069 0.92 ± 0.09 12.5 -1.55 -0.99 Overall 467 112 83 0.0082 ± 0.0007 0.85 ± 0.01 6.7 -2.09 -3.94

Table 5.1 N sample size, H total number of haplotypes, H u number of unique haplotypes at site, π

nucleotide diversity, h haplotype diversity, A E effective number of alleles in COI. Bolded test values are significant at <0.05

147

Table 5.2 Analyses of molecular variance (AMOVA) by region using haplotype distance (ΦST) and non-distance (FST) measures. Source of % of Species Measure Regions Φ statistics P-Values Variation Variation AG 3.90 F = 0.039 0.01 Ophiocoma MHI; NWHI; Johnston Atoll; CT F pica ST Palmyra, French Polynesia AP(G) 0.75 F SC = 0.008 0.07 WP 95.34 AG 21.10 Φ = 0.211 <0.01 Ophiocoma MHI; NWHI; Johnston Atoll; CT Φ pica ST Palmyra, French Polynesia AP(G) 1.24 Φ SC = 0.016 0.02 WP 77.66 AG 7.07 F = 0.071 0.01 Ophiocoma Hawaii; Johnston Atoll; CT F pica ST Palmyra, French Polynesia AP(G) 0.93 F SC = 0.010 0.03 WP 92.00 AG 36.29 Φ = 0.363 0.03 Ophiocoma Hawaii; Johnston Atoll; CT Φ pica ST Palmyra, French Polynesia AP(G) 0.85 Φ SC = 0.013 0.03 WP 62.86 AG 10.08 F = 0.101 <0.01 Ophiocoma Hawaii + Johnston Atoll; CT F pica ST Palmyra, French Polynesia AP(G) 0.72 F SC = 0.008 0.06 WP 89.20 AG 43.33 Φ = 0.433 0.08 Ophiocoma Hawaii + Johnston Atoll; CT Φ pica ST Palmyra, French Polynesia AP(G) 0.69 Φ SC = 0.012 0.05 WP 55.98 AG 3.48 F = 0.035 0.04 Ophiocoma MHI; NWHI; Johnston Atoll; CT F erinaceous ST Palmyra, French Polynesia AP(G) 3.98 F SC = 0.041 <0.01 WP 92.54 AG 31.83 Φ = 0.318 0.01 Ophiocoma MHI; NWHI; Johnston Atoll; CT Φ erinaceous ST Palmyra, French Polynesia AP(G) 19.03 Φ SC = 0.279 <0.01 WP 49.14 MHI; Kure+Midway+P&H; Lower AG 6.18 F = 0.062 <0.01 Ophiocoma CT F NWHI; Johnston Atoll; Palmyra, erinaceous ST AP(G) 1.53 F SC = 0.016 0.02 French Polynesia WP 92.28 MHI; Kure+Midway+P&H; Lower AG 46.78 Φ = 0.468 <0.01 Ophiocoma CT Φ NWHI; Johnston Atoll; Palmyra, erinaceous ST AP(G) 4.58 Φ SC = 0.086 <0.01 French Polynesia WP 48.65

148

0.670 0.802 0.733 0.756 0.583 0.497 0.666 0.824 0.676 0.765 0.759 0.702 0.793 0.618 French French Polynesia 0.077 0.712 0.904 0.690 0.057 0.870 0.714 0.702 0.713 0.928 0.690 0.689 0.460 0.929 0.081 0.118 0.166 0.138 0.107 0.138 0.140 0.134 0.125 0.117 0.142 0.090 -0.001 French French Palmyra Polynesia Atoll 0.062 0.027 0.001 0.024 0.106 0.713 0.921 0.714 0.132 0.888 0.738 0.718 0.729 0.951 0.705 0.714 0.476 0.956 0.064 -0.006 -0.008 -0.032 -0.049 -0.037 -0.030 -0.005 -0.024 Palmyra Johnston Johnston Hilo Atoll 0.008 0.021 0.060 0.006 0.027 0.012 0.020 0.004 0.040 0.035 0.013 0.037 0.357 0.654 0.090 0.500 0.071 0.073 0.336 0.314 0.052 0.061 0.047 0.157 0.120 -0.002 -0.004 Johnston Johnston Hilo Kona 0.019 0.004 0.005 0.006 0.021 0.009 0.049 0.036 0.007 0.028 0.101 0.401 0.893 0.102 0.770 0.129 0.231 0.129 0.224 0.137 0.249 0.110 0.122 0.054 -0.001 -0.004 -0.008 0.040 Maui Kona 0.003 0.000 0.012 0.020 0.088 0.075 0.023 0.012 0.031 0.295 0.681 0.064 0.477 0.280 0.092 0.104 -0.024 -0.052 -0.024 -0.049 -0.029 -0.020 -0.008 -0.001 -0.002 -0.002 ______Main Hawaiian ______Maui 0.017 0.047 0.050 0.001 0.040 0.039 0.045 Kauai 0.120 ≤ 0.028 ≤ 0.330 0.714 0.124 0.539 0.151 0.078 0.088 0.077 0.123 -0.022 -0.023 -0.021 -0.028 -0.019 -0.006 -0.004 -0.004 -0.052 -0.026 p values in are the half upper right table. of each Bolded values signify st 0.064 0.002 0.001 0.011 0.117 0.025 0.076 0.081 0.005 0.027 Oahu 0.069 0.439 0.902 0.141 0.273 0.149 0.797 0.251 0.111 0.124 0.180 Nihoa -0.001 -0.004 -0.032 -0.004 -0.013 -0.023 0.020 0.007 0.014 0.020 0.001 0.003 0.044 0.021 0.011 0.055 0.008 0.100 0.332 0.726 0.140 0.551 0.042 0.039 0.147 0.156 0.077 Kauai -0.022 -0.021 -0.028 -0.011 -0.010 -0.021 French French Frigate ______Main Hawaiian Islands______Reef 0.044 0.013 0.007 0.028 0.023 0.002 0.004 0.000 0.001 0.023 0.010 0.081 Maro 0.187 0.587 0.110 0.331 0.121 0.131 0.039 0.039 Nihoa -0.017 -0.022 -0.043 -0.001 -0.003 -0.011 -0.022 Reef 0.016 0.016 0.002 0.006 0.010 0.011 0.013 0.010 0.017 0.091 Maro 0.337 0.725 0.101 0.553 0.118 0.128 -0.027 -0.004 -0.014 -0.005 -0.009 -0.006 -0.004 -0.006 -0.023 -0.007 -0.021 Lisianski Family-wise false discovery rate corrected p-values: p-values: corrected rate discovery false Family-wise 0.016 0.037 0.029 0.018 0.014 0.016 0.008 0.014 0.027 0.005 0.166 0.301 0.700 0.202 0.502 0.167 0.047 0.179 0.082 0.086 0.060 -0.003 -0.007 -0.006 -0.008 -0.014 -0.022 Laysan Laysan 0.046 0.089 0.071 0.063 0.004 0.003 0.000 0.042 0.026 0.151 0.136 0.138 0.091 0.071 0.066 0.116 0.059 0.131 0.046 0.122 -0.022 -0.010 -0.007 -0.004 -0.004 -0.007 -0.030 Pearl & Pearl & Hermes Hermes values are contained in the lower left half of each table and Φ st 0.117 0.004 0.000 0.013 0.022 0.011 0.263 0.101 0.327 0.351 0.284 0.308 0.270 0.263 0.294 0.277 0.314 0.251 0.323 0.038 0.082 -0.005 -0.007 -0.001 -0.013 -0.007 -0.021 Midway Midway Kure Kure 0.016 0.036 0.006 0.005 0.024 0.027 0.011 0.122 ______Northwest Hawaiian Islands______0.114 0.072 0.071 0.059 0.062 0.059 0.056 0.051 0.162 0.067 0.174 0.095 -0.006 -0.006 -0.009 -0.008 -0.015 -0.004 -0.025 ______Northwest Hawaiian Islands______Morisita Dissimilarity 0.595 +/- 0.031 Morisita Dissimilarity 0.200 +/- 0.051 Site Hilo Site Hilo Kure Kure Maui Kona Maui Kona Oahu Maro Kauai Kauai Nihoa Nihoa Laysan Laysan Midway Midway Palmyra Palmyra Lisianski Johnston Maro Reef Maro Johnston Atoll French Frigate Pearl & Hermes Pearl & Hermes French PolynesiaFrench French PolynesiaFrench significant differences Shaded cells using outlined the procedure signify correction after in Benjamini significant differences 2008. between sites in tests. both Table 5.3 by site. Pairwise F comparisons Main Main Region Islands Islands Islands Islands Region Hawaiian Hawaiian Hawaiian Hawaiian Northwest Northwest Ophiocoma erinaceous Ophiocoma pica Ophiocoma

149

Table 5.4 Pairwise population migration rate estimates (N e Μ )based on a Bayesian MCMC simulation. The value of M calculated by IMa2 was multiplied by the θ, as calculated by IMa2, of the destination population to estimate migration. The estimates of migration are seperated by direction; the columns are source populations and the the rows are sink populations.

Fr. Polynesia, MHI, Johnston NWHI Ophiocoma erinaceous Palmyra Fr. Polynesia, Palmyra - 0.1375 0.0925 MHI, Johnston 66.19 - 6.848 NWHI 0.4125 15.77 - French Polynesia Johnston Atoll MHI NWHI Palmyra French Polynesia - 0.009375 1.614 0.009375 9.084 Johnston Atoll 63.72 - 251.1 3.692 117.1 MHI 0.2156 5.769 - 7.566 25.06 NWHI 0.01688 2.672 8.102 - 0.3544 Palmyra 285.5 16.57 0.8231 0.6281 -

Hawaiian Islands, Fr. Polynesia Palmyra Ophiocoma pica Johnston Fr. Polynesia - 4.857 14.99 Hawaiian Islands, Johnston 0.1825 - 6.253 Palmyra 19.24 0.2375 -

150

Figure 5.1 Haplotype network for Ophiocoma erinaceous. Each circle represents a unique haplotype connected by a line to those that differ by one base pair. Nodes on lines indicate a missing haplotype and numbers represent multiple missing haplotypes. Each haplotype is color-coded by site and circle size is proportional to frequency with the number of specimens indicated in circles with N > 2. The smallest circles represent one occurrence of a haplotype.

151

Figure 5.2 Haplotype network for Ophiocoma pica. Each circle represents a unique haplotype connected by a line to those that differ by one base pair. Nodes on lines indicate a missing haplotype and numbers represent multiple missing haplotypes. Each haplotype is color-coded by site and circle size is proportional to frequency with the number of specimens indicated in circles with N > 2. The smallest circles represent one occurrence of a haplotype.

152

CHAPTER SIX

Summary

153

SUMMARY

In my dissertation research, I found contrasting patterns of population genetic connectivity in four species of tropical coral reef echinoderms. One of my primary goals in this dissertation is to inform marine conservation management.

Many of my field sites are located within marine conservation areas, including the

Papahānaumokuākea Marine National Monument (PMNM), a chain of coral reef atolls and small basalt pinnacles, home to the most beautiful reefs in Hawai‘i. The size, remoteness, and near “pristine” condition of PMNM make it a unique and ideal natural laboratory for conservation and scientific studies. In service of decreasing the footprint of scientific studies where possible, I reviewed non-lethal sampling techniques for use in genetic studies of many species of marine invertebrates (Chapter 2).

The phylogeographic survey of the sea cucumbers Holothuria atra and H. whitmaei, revealed contrasting patterns of genetic differentiation within Hawai‘i. While

H. whitmaei showed no population structure across the Hawaiian Archipelago and

Johnston Atoll, H. atra was highly structured, with a prominent break between the

Northwestern Hawaiian Islands (NWHI) and the Main Hawaiian Islands (MHI) (Chapter

3). In testing hypotheses about connectivity between the Hawaiian Archipelago and

Johnston Atoll I revealed strong connectivity between Johnston Atoll and the NWHI, with migration primarily in the direction of the NWHI.

Because H. whitmaei has been effectively extirpated from most of its range, only

H. atra was available for a more extensive survey throughout a greater part of their range.

154

I examined the genetic population structure of H. atra across the central tropical Pacific

to show -- based on SAMOVA, AMOVA and BARRIER analyses -- that despite its large

range, H. atra has hierarchical, fine-scale population structure driven primarily by between-archipelago barriers, but with significant differences between sites within an archipelago (Chapter 4). Migrate analyses along with haplotype networks and patterns of haplotype diversity suggest that Hawai‘i and Kingman reef are important centers of the genetic diversity in the region rather than an evolutionary dead-end for migrants from the

Indo-Pacific. Finally I showed that for H. atra Kingman Reef is the most likely stepping stone between Hawai‘i and the rest of the Pacific, not Japan or Johnston Atoll as previously presumed. Johnston Atoll can instead be seen as an outpost of the

Northwestern Hawaiian Islands rather than a gateway to the Hawaiian Archipelago.

Finally, I compared population genetic patterns in two congeneric brittle stars,

Ophiocoma pica and O. erinaceous, across Hawai‘i and Central and Eastern Polynesia

(Chapter 5). Despite having similar ecological niches and sympatric distributions across the samples portion of the range, the congeners O. erinaceous and O. pica have responded differently to the Pacific archipelagic environments. There is a strong disparity in phylogeographic structure within the Hawaiian Islands. A similar phylogeographic pattern was found between Archipelagos, but the mutational distances between haplotypes differed greatly and median-joining networks reconstructed two distinct haplotypic patterns. The fine-scale structuring of populations in O. erinaceous and O. pica suggest that a species-specific approach is needed for responding to the complex relationships between genetically distinct populations.

155

SYNTHESIS

Echinoderms play a major role in structuring many marine ecosystems, and many are described as "keystone species" because of their profound influence on benthic community structure. This dissertation primarily focused on the Hawaiian Archipelago, which lies at the periphery of the tropical Central Pacific and is the most isolated island chain in the world. The central questions were: why some species maintain connectivity and species cohesion between the Hawaiian Islands and the rest of the Pacific, why some species diverge and become Hawaiian endemics, and why other species with similar inferred dispersal ability fail to colonize the Hawaiian Archipelago at all.

Considerable evidence is accumulating that it is indefensible to make predictions of connectivity based solely on proxies such as ecological or phylogenetic similarity, pelagic larval duration, or species range sizes. The fine-scale structuring of populations in

H. atra, H. whitmaei, O. erinaceous and O. pica suggest that a species-specific approach is needed for responding to the complex relationships between genetically distinct populations. These species must be managed on a local scale; migration between archipelagos, and often between islands, does not occur in ecologically relevant time frames.

Given the real-world constraints of limited time and money in marine ecosystem management it would be ideal if model species could stand in as proxies for a host of similar species. This dissertation shows that this ideal scenario is unlikely to be the case; similar life histories and close phylogenetic relationships do not appear to predict population connectivity. Generalizations based on a few representative taxa are unlikely

156 to offer much in terms of delineating boundaries for spatial management areas. Though a more inclusive multi-species approach is bound to cost more in terms of time and resources, it should ultimately payout as more informed, if complex, management.

157