global distribution models for whale sharks · global distribution models for whale sharks...

367
Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira (M.Sc. Modelling the Marine Environment) A thesis submitted to the Australia in fulfilment of the requirements for the degree of Doctor of Philosophy February 2013

Upload: others

Post on 13-Jun-2020

8 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

Global Distribution Models for Whale Sharks

(Assessing Occurrence Trends of Highly Migratory Marine Species)

Ana Micaela Martins Sequeira

(M.Sc. Modelling the Marine Environment)

A thesis submitted to the

Australia

in fulfilment of the requirements for the degree of

Doctor of Philosophy

February 2013

Page 2: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

iii

(Quando eu era pequenina...)

- A minha mãe ensinou-me como explicar sempre o que eu queria.

- O meu pai ensinou-me que não podemos ter sempre o que queremos no momento em que queremos.

Completar o meu doutoramento teria sido impossível se eu não tivesse aprendido estas duas lições.

Esta tese é dedicada aos meus pais.

Translation:

(When I was a little girl...)

- My mum taught me how to always explain what I wanted.

- My father taught me that we cannot always have what we want, when we want it.

Completing my Ph.D. would never have been possible without these two lessons

I dedicate my Ph.D. to my parents.

Page 3: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

iv

Page 4: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

v

TABLE OF CONTENTS

TABLE OF FIGURES ................................................................................................................ x

TABLE OF SUPPLEMENTARY FIGURES ............................................................................. xiii

APPENDIX B ........................................................................................ xiii

APPENDIX C ........................................................................................ xiv

APPENDIX D ........................................................................................ xiv

TABLE OF TABLES ................................................................................................................. xv

TABLE OF SUPPLEMENTARY TABLES .............................................................................. xvii

APPENDIX A ....................................................................................... xvii

APPENDIX B ....................................................................................... xvii

APPENDIX C .......................................................................................xviii

APPENDIX D .......................................................................................xviii

SUMMARY........ ...................................................................................................................... xix

STATEMENT OF ORIGINALITY ............................................................................................. xxi

ACKNOWLEDGMENTS........................................................................................................ xxiii

PREFACE.......... .................................................................................................................. xxvii

CHAPTER I. GENERAL INTRODUCTION ............................................................................. 30

Background............................................................................................30

Whale shark biology and ecology ..........................................................37

Outline of the thesis ...............................................................................44

CHAPTER II. Inferred global connectivity of whale shark populations .................................... 47

Statement of authorship ....................................................................................... 49

Page 5: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

vi

Inferred global connectivity of whale shark populations ....................................... 51

Abstract ................................................................................................. 51

Introduction ............................................................................................ 52

Knowledge base .................................................................................... 56

Global occurrences ............................................................................... 59

Tracking studies .................................................................................... 62

The big picture: formulating global hypotheses ..................................... 70

Foreseeing change ................................................................................ 77

Conclusions ........................................................................................... 78

CHAPTER III. Ocean-scale prediction of whale shark distribution .......................................... 82

Statement of authorship ....................................................................................... 83

Ocean-scale prediction of whale shark distribution .............................................. 85

Abstract ................................................................................................. 85

Introduction ............................................................................................ 86

Methods................................................................................................. 89

Results .................................................................................................. 96

Discussion ........................................................................................... 106

CHAPTER IV. Spatial and temporal predictions of inter-decadal trends in Indian Ocean whale

sharks .......................................................................................................... 114

Statement of authorship ..................................................................................... 115

Spatial and temporal predictions of inter-decadal trends in Indian Ocean whale

sharks .......................................................................................................... 117

Page 6: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

vii

Abstract ...............................................................................................117

Introduction ..........................................................................................118

Methods ...............................................................................................120

Results .................................................................................................128

Discussion ...........................................................................................134

CHAPTER V. Inter-ocean asynchrony in whale shark occurrence ....................................... 138

Statement of authorship ..................................................................................... 139

Inter-ocean asynchrony in whale shark occurrence ........................................... 141

Abstract ...............................................................................................141

Introduction ..........................................................................................142

Methods ...............................................................................................144

Results .................................................................................................150

Discussion ...........................................................................................158

CHAPTER VI. Predicting current and future global distributions of whale sharks ................. 165

Statement of authorship ..................................................................................... 167

Predicting current and future global distributions of whale sharks ..................... 169

Abstract ...............................................................................................169

Introduction ..........................................................................................170

Methods ...............................................................................................174

Results .................................................................................................184

Discussion ...........................................................................................190

CHAPTER VII. GENERAL DISCUSSION ............................................................................. 196

Page 7: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

viii

Overview of my research ..................................................................... 196

A worldwide perspective ...................................................................... 199

Better management ............................................................................. 201

Future improvements........................................................................... 203

Future directions .................................................................................. 209

AFTERWORD...... ................................................................................................................. 216

APPENDIX A – Inferred global connectivity of whale shark populations .............................. 217

APPENDIX B – Ocean-scale prediction of whale shark distribution ..................................... 219

Receiving operator characteristic curve / area under the cure (ROC/AUC) ....... 227

Boyce index comparison between MaxEnt and GLM/GLMM models ................ 228

BIOMOD, GBM and BRT models for the Indian Ocean ..................................... 229

BIOMOD .............................................................................................. 229

BRT / GBM .......................................................................................... 233

APPENDIX C – Spatial and temporal predictions of decadal trends in Indian Ocean whale

sharks .......................................................................................................... 236

Model predictors (Step 1) .................................................................... 236

Prevalence test (down-weighting pseudo-absences) .......................... 242

APPENDIX D – Predicting current and future global distributions of whale sharks ............... 243

BIOMOD – Global model ..................................................................... 244

ADDENDUM – Thesis amendments ..................................................................................... 247

Outcome of thesis examination .......................................................................... 247

Letter from Examiner 1 ........................................................................ 247

Page 8: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

ix

Letter from Examiner 2 ........................................................................251

Reply to examiners ............................................................................................ 255

GENERAL COMMENTS: .....................................................................255

SUMMARY ..........................................................................................257

CHAPTER I ..........................................................................................258

CHAPTER II .........................................................................................267

CHAPTER III ........................................................................................272

CHAPTER IV .......................................................................................281

CHAPTER V ........................................................................................283

CHAPTER VI .......................................................................................301

CHAPTER VII ......................................................................................317

REFERENCES. .................................................................................................................... 325

Page 9: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

x

TABLE OF FIGURES

Figure 1: Diagram showing the components and main decision steps of species

distribution modelling. Adapted from Franklin (2010). .................................... 33

Figure 2: Whale shark (Rhincodon typus) (Smith, 1828) observed at Ningaloo Reef,

Western Australia. ........................................................................................ 38

Figure 3: Whale shark distribution range and sighting reports known as of 2009. ...... 43

Figure 4: Graphical illustration of the number of representative whale shark studies

(since 1997). .................................................................................................. 58

Figure 5: Compilation of worldwide whale shark occurrences. .................................... 60

Figure 6: Global sequence of whale shark monthly appearance timings (3 years

presented) in different locations worldwide. ................................................... 63

Figure 7: Global overview of published whale shark tracks. ........................................ 69

Figure 8: Whale shark migration patterns.................................................................... 76

Figure 9: Indian Ocean Tuna Commission (IOTC) datasets collected from 1991 to

2007 on tuna purse-seine fisheries in the Indian Ocean. ............................... 91

Figure 10: Seasonal variation in aggregated whale shark (Rhincodon typus, Smith

1828) sightings per unit effort (SPUE, where effort corresponds to the number

of fishing days) in the area of the Indian Ocean sampled by the Indian Ocean

Tuna Commission (IOTC) from 1991 to 2007. ............................................... 97

Figure 11: Habitat suitability of whale sharks (Rhincodon typus, Smith 1828) in the

Indian Ocean during autumn........................................................................ 100

Figure 12: Seasonal ensemble result for whale sharks (Rhincodon typus, Smith 1828)

habitat suitability. ......................................................................................... 107

Page 10: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

xi

Figure 13: Temporal variation in whale sharks sighted by the tuna purse-seine

fisheries and comparison with climatic indices. ............................................122

Figure 14: Example of variation in fishing effort (days – top row) and number of whale

sharks sighted (bottom row). ........................................................................123

Figure 15: Results of the fishing effort analysis across the western Indian Ocean

where the purse-seine fisheries operated in more than one year.................124

Figure 16: Partial effect of time on whale shark presence. ........................................132

Figure 17: Partial effects of NINO4 and Time (after accounting for climatic

contributions) in the probability of whale shark presence. ............................133

Figure 18: Whale shark (Rhincodon typus, Smith 1828) datasets recorded by tuna

purse-seiners ...............................................................................................146

Figure 19: Predicted habitat suitability of whale sharks (Rhincodon typus, Smith 1828)

in the Atlantic (left) and Pacific (right) Oceans during the months of April to

June. ............................................................................................................151

Figure 20: Yearly variation in (a) number of whale sharks sighted, (b) relative whale

shark occurrences, and (c) days spent fishing. ............................................154

Figure 21: Partial effect of time on the probability of whale shark presence during the

months of April to June. ...............................................................................156

Figure 22: Whale shark sightings per unit effort during the months of April to June in

the Atlantic (top row; 1981 - 2010) and Pacific (bottom row; 2000 - 2010). ..157

Figure 23: Fisheries data derived from tuna purse-seine fishery logbooks in the

Atlantic, Indian and western Pacific Oceans. ...............................................179

Figure 24: Boxplots for major environmental predictors within the area covered by the

fisheries in each ocean. Upper and lower limits indicate the maximum and

minimum values in each dataset. .................................................................185

Page 11: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

xii

Figure 25: Predicted whale shark habitat suitability within the area covered by tuna

purse-seine fisheries. ................................................................................... 186

Figure 26: Global predictions of current seasonal habitat suitability for whale sharks.

.................................................................................................................... 188

Figure 27: Predicted shift in global of whale sharks habitat suitability for 2070 under a

no-climate-policy reference (no greenhouse gas emissions) scenario. ....... 189

Figure 28: Map of world organisations managing fisheries. ...................................... 210

Figure 29: Conditional modes of the random year intercept term for models including

time as fixed effect (quadratic term; left), and without the fixed temporal trend

(right). .......................................................................................................... 293

Page 12: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

xiii

TABLE OF SUPPLEMENTARY FIGURES

APPENDIX B

Figure B1: Bathymetry of the Indian Ocean as per the General Bathymetric Chart of

the Oceans (1 minute grid ~ 1.8 km). ...........................................................222

Figure B2: Seasonal chlorophyll a concentration (Chl a) obtained from averaged of

weekly SeaWiFS satellite composites at a 9 km spatial resolution from 1997

to the end of 2007. .......................................................................................222

Figure B3: Seasonal sea surface temperature (SST) obtained from averaged weekly

MODIS-Aqua satellite composites at a 9-km spatial resolution from 2002 to

the end of 2007. ...........................................................................................223

Figure B4: Probabilities (top panel) and resulting pseudo-absences locations (bottom

panel) generated by three different techniques. ...........................................223

Figure B5: Moran’s I plots showing the reduction in spatial autocorrelation in the GLM

residuals. ......................................................................................................224

Figure B6: Whale shark (Rhincodon typus, Smith 1828) habitat suitability for each

season as predicted by MaxEnt when using the full background with only

linear and quadratic features. .......................................................................225

Figure B7: Seasonal habitat suitability of whale sharks (Rhincodon typus, Smith 1828)

in the Indian Ocean. .....................................................................................226

Figure B8: Whale shark autumn distribution in the Indian Ocean as a result of the

ensemble of models in BIOMOD. .................................................................232

Figure B9: 3D interpretation of the two top ranked pairwise interactions found with

BRT. .............................................................................................................234

Page 13: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

xiv

Figure B10: Map with probability of whale shark occurrence in the Indian Ocean

obtained by BRT. ......................................................................................... 235

APPENDIX C

Figure C1: Purse-seine fishing effort (in days) during autumn for each year from 1991

to 2007. ........................................................................................................ 238

Figure C2: Habitat suitability of whale sharks (Rhincodon typus Smith 1828) in the

Indian Ocean during autumn........................................................................ 239

Figure C3: Correlation between climatic predictor variables. .................................... 240

Figure C4: Partial effect of the Indian Ocean Dipole (IOD) on the probability of whale

shark sightings. ............................................................................................ 241

Figure C5: Partial effect of time on whale shark presence, using a presence:pseudo-

absence ratio of 0.5 (prevalence). ............................................................... 242

APPENDIX D

Figure D1: Global sea surface temperature averages from April to June. ................. 243

Figure D2: Comparison of GLMM and BIOMOD ensemble prediction for the area

covered by the tuna fisheries (mean probability). ........................................ 246

Page 14: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

xv

TABLE OF TABLES

Table 1: List of published whale shark tracks from the Indian Ocean. .........................64

Table 2: List of published whale shark tracks from the Pacific Ocean and Indo-Pacific

area. ...............................................................................................................66

Table 3: List of published whale shark tracks from the Atlantic Ocean. .......................67

Table 4: Summary of generalised linear models relating probability of whale shark

(Rhincodon typus, Smith 1828) occurrence to ocean properties. ...................99

Table 5: Comparison of results for the MaxEnt and generalised linear (GLM) and

mixed effects (GLMM) models. ....................................................................103

Table 6: Summary of the MaxEnt models relating probability of whale shark

(Rhincodon typus, Smith 1828) occurrence to individual ocean properties. .104

Table 7: Summary of the generalised linear mixed-effects models relating probability of

whale shark occurrence to spatial and temporal predictors. ........................130

Table 8: Summary of spatial generalised linear models (step 1) relating probability of

whale shark (Rhincodon typus, Smith 1828) occurrence to ocean properties.

.....................................................................................................................152

Table 9: Summary of the temporal generalised linear mixed-effects models (step 2).

.....................................................................................................................155

Table 10: Estimated weight of evidence for each temporal predictor used in the

generalised linear mixed-effects models: Hsuit - spatial predictor derived from

the spatial distribution models, effort - temporal variation in fishing effort and a

linear and quadratic term for time (years). ...................................................155

Table 11: Number of whale shark sightings in each ocean per decade as recorded by

tuna purse-seine fisheries. ...........................................................................177

Page 15: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

xvi

Table 12: Summary of generalised linear mixed-effects models relating probability of

whale shark occurrence to ocean properties in the Atlantic, Indian and

western Pacific during the season of peak occurrences for each ocean. .... 181

Page 16: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

xvii

TABLE OF SUPPLEMENTARY TABLES

APPENDIX A

Table A1: Summary of representative whale shark studies (since 1996) included in

Figure 4. .......................................................................................................217

APPENDIX B

Table B1: Summary of the MaxEnt models relating probability of whale shark

(Rhincodon typus, Smith 1828) occurrence to individual ocean properties. .219

Table B2: Comparison of the Area Under the Curve (AUC) and Kappa results for the

two modelling approaches used in chapter III: MAxEnt and GLM/GLMM for

each of the pseudo-absence dataset generated method (P/A) and for each

season: Autumn (Aut), Winter (Win), Spring (Spr) and Summer (Sum). ......227

Table B3: Boyce’s index for the three pseudo-absence generation techniques (P/A) for

each modelling technique (MaxEnt and GLM) for autumn. Standard deviation

was always < 0.001. .....................................................................................228

Table B4: BIOMOD modeling summary.....................................................................229

Table B5: Best model selection in BIOMOD ..............................................................230

Table B6: Ranking of variable importance for each model considered within the

BIOMOD platform. .......................................................................................231

Table B7: Analysis of pairwise interactions between variables using BRT ................233

Page 17: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

xviii

APPENDIX C

Table C1: Treatment given to each explanatory variable (standardization) used to

obtained the spatial predictor (SpatialP) (Sequeira et al., 2012) for the first

step of our temporal models; SST – sea surface temperature; Chl a –

chlorophyll a. ............................................................................................... 237

APPENDIX D

Table D1: BIOMOD modelling summary for the runs with the global dataset. ........... 244

Table D2: Model selection in BIOMOD...................................................................... 245

Table D3: Ranking of variable importance for the global model considered within the

BIOMOD platform. ....................................................................................... 245

Page 18: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

xix

SUMMARY

The processes driving distribution and abundance patterns of highly migratory marine

species, such as filter-feeding sharks, remain largely unexplained. The whale shark

(Rhincodon typus Smith 1828) is a filter-feeding chondrichthyan that can reach > 18 m

in total length, making it the largest extant fish species. Its geographic range has been

defined within all tropical and warm temperate waters around the globe. However,

even though mitochondrial and microsatellite DNA studies have revealed low genetic

differentiation among the three major ocean basins, most studies of the species are

focussed on the scale of single aggregations. Our understanding of the species’

ecology is therefore based on only a small proportion of its life stages, such that we

cannot yet adequately explain its biology and movement patterns (Chapter I). I present

a worldwide conceptual model of possible whale shark migration routes, while

suggesting a novel perspective for quantifying the species‘ behaviour and ecology.

This model can be used to trim the hypotheses related to whale shark movements and

aggregation timings, thereby isolating possible mating and breeding areas that are

currently unknown (Chapter II). In the next chapter, I quantify the seasonal suitable

habitat availability in the Indian Ocean (ocean basin-scale study) by applying

generalised linear, spatial mixed-effects and maximum entropy models to produce

maps of whale shark habitat suitability (Chapter III). I then assess the inter-annual

variation in known whale shark occurrences to unearth temporal trends in a large area

of the Indian Ocean. The results from the Indian Ocean suggest both temporal and

spatial variability in the whale sharks occurrence (Chapter IV). Therefore, I applied the

same analysis to the Atlantic and Pacific Oceans using similar broad-scale datasets.

While the results for the Pacific Ocean were inconclusive with respect to temporal

Page 19: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

xx

trends, in the Atlantic Ocean I found preliminary evidence for a cyclic regularity in

whale shark occurrence (Chapter V). In Chapter VI, I build a model to predict global

whale shark habitat suitability for the present, as well as within a climate change

scenario for 2070. Finally, Chapter VII provides a general discussion of the work

developed within this thesis and presents ideas for future research.

Page 20: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

xxi

STATEMENT OF ORIGINALITY

This work contains no material which has been accepted for the award of any other

degree or diploma in any university or other tertiary institution and, to the best of my

knowledge and belief, contains no material previously published or written by another

person, except where due reference has been made in the text.

I give consent to this copy of my thesis when deposited in the University

Library, being made available for loan and photocopying, subject to the provisions of

the Copyright Act 1968.

The author acknowledges that copyright of published works contained within

this thesis (as listed below*) resides with the copyright holder(s) of those works.

I also give permission for the digital version of my thesis to be made available

on the web, via the University’s digital research repository, the Library catalogue, and

also through web search engines, unless permission has been granted by the

University to restrict access for a period of time.

Ana Micaela Martins Sequeira

Adelaide 27 February 2013

* SEQUEIRA, A., MELLIN, C., ROWAT, D., MEEKAN, M. G. & BRADSHAW, C. J. A. (2012). Ocean-

scale prediction of whale shark distribution. Diversity and Distributions 18, 504-518.

*SEQUERIA, A., MELLIN, C., MEEKAN, M.G., SIMS, D.W., BRADSHAW, C.J.A. (2013). Inferred global

connectivity of whale shark Rhincodon typus populations. Journal of Fish Biology, 82, 367-389.

Page 21: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

xxii

Page 22: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

xxiii

ACKNOWLEDGMENTS

Thanks to the Fundação para a Ciência e Tecnologia (Portuguese Foundation for Science and Technology) and European Social Funds (Ph.D. scholarship SFRH/BD/47465/2008), and to the University of Adelaide for funding my Ph.D. project. Also, thanks to the Australian Institute of Marince Science, Flying Sharks, Russel and Baudinette Travel Scholarship, and Society for Conservation Biology for additional funding / support.

I am thankful to my supervisors for the opportunity to undertake this project and for their support:

- Corey Bradshaw, who (in first place) believed in me and in my capacity to undertake this project, and fully supported me throughout (including overcoming my portuguenglish writing) always in a positive and encouraging way

- Mark Meekan, who made sure I had enough time to think, re-think and think again, and then finally be certain, that I wanted to do this Ph.D., and only then allowed me the opportunity to live unique “whale shark moments”

- Camille Mellin, who joined in as my Ph.D. supervisor after my arrival to Adelaide, and turned out to be my research “pillar” always with supportive and friendly advice.

Thanks to all the fishery-vessel owners (pole and line and purse-seiners) for access to log-book data used throughout my Ph.D. and to the Indian Ocean Tuna Commission and Institut de Recherche pour le Développement (France), the Azores Fisheries Observer Programme and the University of the Azores (Portugal), the Inter-American Tropical Tuna Commission and the Secretariat of the Pacific Community – in particular, A. Fonteneau, M. Herrera, R. Pianet, P. Afonso, M. A. Gaspar, M. Hall and P. Williams for extraction of the data. Also to D. Rowat for initial assistance with access to data.

Thanks to everybody involved in the whale shark field trips (2009 – 2012) for their help, partaking of knowledge and sharing of wonderful “whale shark moments”. Special thanks to: Iain Field, Michelle Thums, Kim Brooks and Terry Maxwell.

Page 23: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

xxiv

Thanks to the staff members from all the institutions involved, who in one way or another contributed and helped me throughout my Ph.D. In particular: Michelle Picard (for friendly support), Janelle Palmer, John Jennings, Louise Gillis, Maria Albanese and Lucy Zuzolo (from the University of Adelaide); Shona Pettigrew, Sharon Haswell, Susan Berry, Gheslaine Holmes and Mary Bell (from the Australian Institute of Marine Sciences), and Nuno Lima (from Fundação para a Ciência e Tecnologia).

Thanks to the Global Ecology Lab group for help and support. A special thanks to: Steven Delean (for helpful assistance with statistics), Salvador Herrando-Pérez (my mate in the lab), Lochran Trail, Nerissa Haby and Stephen Gregory.

While in Adelaide, my friends (Nadja Klafke, Suzanne McAllister, Kym Abrams, Esther Breed, Maria Marklund and Julian Felton), the King family (Gordon, Adrienne, Nicholas, Karen, Michael, Simone and Kyra – now part of my family as well) and the South Coast Wing Chun Kung Fu crew, kept the experience “human” ;-) and provided friendly and happy moments.

From the other side of the planet (where my reality makes sense...) there is not enough space to thank all: my family and beloved friends!

But I cannot avert mentioning here:

- My parents: Lurdes and Artur - who take care of me so well (even from afar)

- My sisters: Sónia and Suzete - their believe in me gives me strength

- My friends: Sofia, Célia and Inês - they keep me alive instead of just living

Thanks for being here, there... every time everywhere!

João Nuno (R.I.P.) thanks for dreaming the Australian dream with me - it became REAL!

Thanks for everything ♥Gary♥ (especially, for holding my hand).

Page 24: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

xxv

As requested by the major funding organisation, here I include their logos:

Ph.D. scholarship SFRH/BD/47465/2008

Page 25: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

xxvii

PREFACE

““WWhheenn yyoouu wwaanntt ssoommeetthhiinngg,, tthhee wwhhoollee uunniivveerrssee ccoonnssppiirreess ttoo

hheellpp yyoouu rreeaalliissee yyoouurr ddeessiirree”” ((PPaauulloo CCooeellhhoo))

My Ph.D. has a story! One which I cannot refrain from mentioning here - as I feel it

belongs to this thesis as much as all the text, figures and tables.

It took me several years to ‘swap’ from ‘pure' chemistry (my background) to

marine ecology, and ending up doing a Ph.D. on whale sharks. It was a winding road

which involved working in several different environments: marine chemistry,

oceanographic modelling, ecosystem modelling (linking hydrodynamics, to

biogeochemistry and to species physiology), and finally, marine biology while working

on connectivity of marine invasive species. During this process, my pursuit was

focused on doing something about the global lack of knowledge on one of the most

enigmatic (and impressive) species of the ocean. I wrote (and dropped) several drafts

of possible research projects before the one I sent to Mark and Corey - which came

back with positive feedback. After all my contacts, I finally got supervisors for my Ph.D.

project. The hunt for funding followed...

Four years passed since the first email I sent to Mark until I finally got a

scholarship which allowed me to start my Ph.D. in 2009. So no, the road was not easy.

But in hindsight, I would not have been able to complete this Ph.D. if it had gone any

other way. The experience I gained, the challenges involved in gathering new skills

and working in different environments, the mind frame I got into before starting this

Ph.D., and my perseverance were all vital to getting me here. They allowed the

Page 26: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

xxviii

compilation of a comprehensive study on one species while still keeping a broader

view on the wider applications of the study and tools developed. The ‘road’ would have

felt longer and windier if I was alone - the credit for my perseverance does not belong

just to me. I would like to formally thank all of those who supported my aspiration,

listened to me blabber all these years, encouraged me whenever I needed it, and in

the end, allowed me to believe that:

There is nothing more fulfilling than pursuing our dreams!

This thesis is the first outcome of my pursuit.

Note: This photo was taken during my first whale shark field trip in Ningaloo, Western Australia, when I wore for the first time the t-shirt I received as a present from my sister Sonia (after starting my Ph.D.). I include it here as an ode to all of those with whom I exchanged paths, and whom enriched me as a person.

Page 27: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

29

Global Distribution Models for Whale Sharks

(Assessing Occurrence Trends of Highly Migratory Marine

Species)

Page 28: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

30

CHAPTER I. GENERAL INTRODUCTION

Background

The oceans cover 70 % of the Earth’s surface, and hence play a vital role in the water

cycle and climate regulation (Chahine, 1992). Global changes to this system are now

affecting the hydrological cycle (Stocker & Raible, 2005), with major, if largely

unpredictable, consequences for ocean life (Behrenfeld et al., 2006). Despite the

plethora of ways in which humans exploit the marine environment, the marine realm is

largely under-studied and its complexities remain therefore mysterious (Richardson &

Poloczanska, 2008). This is mainly due to the physical and economic constraints of

collecting data in the marine realm (Richardson & Poloczanska, 2008); for example,

broad-scale biological data collection only became possible after the recent

developments on satellite-sensor technology that allowed detection of changes in

ocean colour associated with phytoplankton abundance (Sambrotto, 1999). However,

this technology applies only to the sea surface, and so most of the ocean is still largely

unsampled. The latter also applies to sampling marine species from higher trophic

levels, which still relies on expedition surveys. In spatial ecology (Tilman & Kareiva,

1997), the distribution of species can be interpreted as a match between their

ecological requirements and the environmental conditions available. Therefore, the

difficulties of sampling the marine environment (through spatial surveys) hamper

detection of the relevant spatial coverage to interpret these associations correctly.

Modelling approaches are used to assist the understanding of the relationships

between fundamental biological conditions (or processes) and species occurrence - as

I discuss below.

Page 29: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

31

The distribution of marine populations is also strongly associated with

movement, whether through migration (e.g., herring; Dickey-Collas et al., 2009)or

through larval dispersal (e.g., reef fishes; Jones et al., 2009). This leads many marine

ecology studies to embrace and test metapopulation concepts and hypotheses

(Hanski, 1998; Roughgarden et al., 1985), even if not specifically acknowledged.

Metapopulation dynamics are the processes associated with population turnover, i.e.,

extinction versus colonisation (based on Levin’s model from 1969), between patchily

distributed subpopulations with some degree of connectivity between them (Hanski &

Gilpin, 1991). In terrestrial systems, this theory has been widely applied to a range of

species including butterflies (Hanski, 1994), rabbits (Nielsen et al., 2008), rats (O’Brien

et al., 2008) and bears (Hellgren et al., 2005). Indeed, fragmentation of terrestrial

habitats makes the application of applied metapopulation theory highly relevant. In the

marine environment, metapopulation concepts have been applied mostly to

invertebrates and reef fishes (Sale & Kritzer, 2003); however, testing hypotheses at

broader spatial scales might not be suitable or even practical for many marine species

that are either highly migratory or widely dispersing (Kritzer & Sale, 2006).

Despite our limited understanding of oceans relative to the terrestrial realm,

global declines of commercially important fish stocks have been reported for years

(e.g., Pauly et al., 2005), and climate- and fishing-driven shifts in the distribution of

marine species are now being detected (Cheung et al., 2010; Dulvy et al., 2008;

Edwards & Richardson, 2004; Hiddink & ter Hofstede, 2008; Perry et al., 2005).

Together with evidence that marine biodiversity improves the ocean’s resilience to

disturbance (Folke et al., 2004), it is imperative that studies are implemented to

determine the underlying mechanisms driving these changes, as well as quantifying

the magnitude of their impacts and identifying possible solutions.

Page 30: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

32

To that end, more and more modelling tools are being developed and used to

understand how the marine environment functions and how it might change in the near

future. These range from hydrodynamic models that emulate ocean circulation or

ocean-atmosphere interactions (Bush & Philander, 1998; Lazure & Dumas, 2008;

Penven et al., 2001; Schmittner et al., 2002; Woods, 1985), to ecosystem models

constructing marine trophic web (Baretta et al., 1995; deYoung et al., 2004),

estimating coastal carrying capacity (Ferreira et al., 2008) and guiding sustainable

management (Sequeira et al., 2008). Another class of model – ‘species distribution

models’ (also known as ‘habitat suitability models’ or ‘resource selection functions’) –

that has been applied mainly in terrestrial systems (Lehmann et al., 2002) are now

increasingly being used in the marine environment (Elith & Leathwick, 2009; Robinson

et al., 2011). These models combine information on species occurrence with

concurrent environmental conditions to define a species’ ecological niche (Kearney &

Porter, 2009), which can then be used to predict the probability of occurrence where

no relevant biological information is available (Robertson et al., 2003). Species

distribution models are therefore irreplaceable tools to assist resource management

and conservation planning, such as in designing marine reserves (e.g., Beger &

Possingham, 2008), assessing the consequences of climate change to biodiversity

(Thuiller, 2007), or in determining species richness and abundance (e.g., Mellin et al.,

2010b).

Franklin (2010) schematised the main requirements of species distribution

models (Figure 1) – these include (i) the hypothesis to test (e.g., species occurrence

depending on specific environmental conditions), (ii) data on species occurrence and

assessment of their integrity, (iii) information on environmental variables within the

study area at a relevant resolution, (iv) choice of a modelling method (dependent on

Page 31: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

33

the previous components), (v) validation procedures and (vi) mapping tools (Elith &

Leathwick, 2009; Franklin, 2010).

Figure 1: Diagram showing the components and main decision steps of species distribution modelling. Adapted from Franklin (2010).

The data on species occurrence first needs to be assessed for its integrity which involves determining the data type (occurrences and absences data), if there is any bias associated with the sampling, and the probability of detection of the species within the time frame of the sampling scheme. Biogeographical environmental data assist identifying the species ecological niche, and is needed as a digital map covering an area (i.e., a digital layer). These data are also useful at the stage of mapping the predicted occurrences. With both species occurrence and environmental data, the modelling procedure can take place. It typically involves the choice of modelling methods, variables to include, response variable and decision on any transformations needed in the data. Any bias derived from sampling procedures should be included in the models (e.g., through the use of ‘offsets’ or weighting the variables according to the known bias). Modelling validation can take different forms and be both quantitative (assess model fit and accuracy) and qualitative (compare the response results with real data – species data). An important step when the model’s objective is to explain the influence of environmental predictors in the occurrence of a species is to assess spatial autocorrelation. Nearby locations

A NOTE:

This figure/table/image has been removed to comply with copyright regulations. It is included in the print copy of the thesis held by the University of Adelaide Library.

Page 32: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

34

tend to have similar properties, and are likely to influence each other or be influenced by the same processes in one same pattern. This can affect the precision of the coefficients estimated and bias the variable selection by the model. Spatial correlograms of the residuals can help detecting spatial autocorrelation and some strategies can be used to reduce it (e.g., choosing observations that are within a pre-determined distance of others). When building the prediction map, an important step is the definition of a threshold to separate presence locations from absences. This map can be visually compared with real data or expert knowledge. Usually this map is built for the sampled area, although in some instances it might be useful to project outside the sampled area – extrapolation in space and time. This step should be careful, as when extrapolating to outside the range of the values for which the model was fit, the predictions can be erroneous (this step is indicated with a dashed line at the bottom of the scheme).

Addressing each of these model components can be challenging (Araújo &

Guisan, 2006). For example, selecting the ‘right’ model depends on the nature of both

the response variable (occurrence of species, such as whether presences, absences

or both are available) and predictors (environmental variables) (Araújo & Guisan,

2006). Different model constructs are available depending on data types: box

classifiers (e.g., BIOCLIM), machine learning (e.g., MaxEnt) (Phillips et al., 2004) and

statistical/ regression methods (e.g., generalised linear or additive models) (Hastie &

Tibshirani, 1986; Nelder & Wedderburn, 1972). Different methods also result in

different performances (summarized in Franklin, 2010), but there is general agreement

that regression methods using presence and absence data tend to provide more

realistic projections (Brotons et al., 2004; Elith et al., 2006; Zaniewski et al., 2002).

With the increasing pressures on biodiversity from global change, species distribution

models have now been coupled to climate forecasting models to predict future shifts in

species occurrence and habitat changes (Araújo et al., 2005; Cheung et al., 2010;

Thuiller, 2004).

Shifts in the distribution of marine species associated with global changes

(Hoegh-Guldberg & Bruno, 2010; Sumaila et al., 2011; Wernberg et al., 2011) have

now been described for a range of taxa, from primary producers such as plankton

(Beaugrand et al., 2009; Edwards, 2004; Southward et al., 1995), macroalgae

Page 33: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

35

(Wernberg et al., 2011), and filter feeders (Berge et al., 2005), to grazers (Rose, 2005;

Sabatés et al., 2006) and predators (e.g., cod; Beaugrand & Kirby, 2010; Dulvy et al.,

2008; Perry et al., 2005). Moreover, modelling forecasts show that similar shifts are to

be expected for other species (Cheung et al., 2010), along with possible extinctions

(Caldeira & Wickett, 2003; Thuiller et al., 2004a).

Assessment of climate-induced shifts on the distribution of marine predators

such as sharks has to date been hindered by our limited knowledge of their movement

patterns and their susceptibility to environmental shifts. Nevertheless, many shark

species are in decline, largely as a result of over-fishing (Baum et al., 2003; Field et

al., 2009; Musick et al., 2000; Schindler et al., 2002). In some countries, shark

landings have even surpassed those of most other commercially exploited species

(e.g., Sri Lanka) (Barker & Schluessel, 2005). Due to their slow life histories (delayed

maturity, long life spans, few offspring), sharks are particularly vulnerable to

exploitation (Baum et al., 2003; Field et al., 2009; García et al., 2007; Hutchings et al.,

2012; Musick et al., 2000; Schindler et al., 2002). Consequently, many recent

management strategies, increasingly reliant on good biological information and model

projections, have been devised to maximise the long-term persistence of shark

populations in the face of these pressures (Barker & Schluessel, 2005).

Appropriate management of shark populations is also a commercially lucrative

ambition because of the high revenue some shark species bring to the tourism

industry (Vianna et al., 2012). Sharks are often the subject of major tourism ventures,

such as shark watching and diving (Anderson & Waheed, 2001; Brunnschweiler, 2010;

Catlin & Jones, 2009), and some have argued that certain species are worth more

alive than dead – a recurrent source of profit instead of a one-off income (Topelko &

Dearden, 2009; Vianna et al., 2012). Moreover, sharks are also thought to play a key

Page 34: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

36

role in controlling the food web structure and the dynamics of marine ecosystems (Frid

et al., 2008; Myers et al., 2007; Scheffer et al., 2005). Despite the importance of this

taxon, the basic biology and ecology of most sharks are still poorly quantified. For

example, the economically important (tourism, fishing) whale shark – the largest fish

species – is still largely unstudied. There is some evidence that whale sharks can

travel across oceans, which raises the possibility of applying concepts and testing

hypotheses associated with metapopulation theory. Depending on true connectivity

(currently unquantified) of whale shark sub-populations worldwide, they might actually

be part of a unique global meta-population, such that localised perturbations might

have detrimental flow-on effects to other sub-populations in distant waters. Whale

shark distribution is also strongly associated with temperature; they are generally

found in warm and temperate seas (Last & Stevens, 2009) and spend most of their

time at the surface (Gunn et al., 1999; Sleeman et al., 2010a; Sleeman et al., 2010b;

Wilson et al., 2006) within a narrow range of temperatures (Sequeira et al., 2012). This

limited temperature range suggests that their time in deeper waters is limited by

thermoregulatory constraints (Thums et al., 2012) or minimises the time spent in deep,

oxygen-poor water (Graham et al., 2006). Changes in water temperature associated

with climate change (Hoegh-Guldberg & Bruno, 2010), are thus likely to change whale

shark distribution or abundance as they re-equilibrate to new temperature regimes and

adapt to changing prey distribution (Gutierrez et al., 2008).

Being filter feeders, whale sharks are also likely to be affected by the

expected climate-driven changes in ocean productivity and food-web dynamics

(Hoegh-Guldberg & Bruno, 2010). Plankton distribution is expected to shift and its

abundance will likely change in response to the thermal properties of changing oceans

(Reygondeau & Beaugrand, 2011). This logically suggest that such changes will

Page 35: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

37

propagate through to higher trophic levels (Kirby & Beaugrand, 2009). This led Hayes

et al. (2005) to refer to plankton monitoring programmes as “sentinels” to identify

changes in marine ecosystems. Whale sharks have the capacity to travel large

distances, so altered plankton distribution will repeatedly affect whale shark migration

patterns (Chin et al., 2010). Moreover, whale sharks aggregations are usually

associated with specific productivity events, such as coral spawn in Ningaloo, Western

Australia (Taylor, 1996) that promotes an increase in zooplankton production.

Changes in the timing of these events, as well as other impacts of climate change on

coral reefs (e.g., habitat destruction within whale shark feeding areas), might therefore

also impact whale shark occurrence (Chin et al., 2010).

Despite whale sharks spending the majority of their time at the surface (Gunn

et al., 1999; Rowat et al., 2007; Thums et al., 2012), they can also dive deeply. This

opens speculation to the potential for this species to use different depths (where water

is cooler) as surface water temperatures increase. However, oxygen limitation might

pose a problem for extended time spent in deeper waters (Pörtner, 2001; Prince &

Goodyear, 2006). For most of the reasons given above, whale sharks have been

referred to “as potentially the most vulnerable species [to climate change] in the

pelagic ecological group” (Chin et al., 2010; in their assessment of the vulnerability of

sharks and rays in the Great Barrier Reef), and present a challenging case study in the

context of climate change and expected shifts in marine species distribution.

Whale shark biology and ecology

Whale sharks (Rhincodon typus) (Smith, 1828) are the largest of all fish (Figure 2),

with records reporting animals as large as 18.8 and 20 m total length (Borrell et al.,

Page 36: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

38

2011; Chen et al., 1997). However, due to difficulties associated with measuring such

a large animal (Wintner, 2000), and lack of specification of the methods used in some

studies reporting individual lengths, a commonly accepted maximum size is around 12

m (Stevens, 2007). The time it takes for a whale shark to reach such a size is still

unknown. Some studies suggest that growth rates are slow (~20 to 34 cm year-1)

(Wintner, 2000), concluding that the species must be long-lived. In fact, a life span of

60 to 100 years has been suggested for this species (Pauly, 1997), and maturity is

expected at lengths ≥ 9 m (Norman & Stevens, 2007).

A large body size, long life span and probable late maturity are characteristic

traits of a K-selected (‘slow’) life history (Stearns, 1976). Species with such traits are

also expected to produce only a few offspring. Despite the limited data on whale

sharks reproduction, there is evidence this might not be the case for whale sharks.

Most of what is known about whale shark reproduction derives from a single pregnant

Figure 2: Whale shark (Rhincodon typus) (Smith, 1828) observed at Ningaloo Reef, Western Australia. Photography by Wayne Osborne from the Australian Institute of Marine Science.

A NOTE:

This figure/table/image has been removed to comply with copyright regulations. It is included in the print copy of the thesis held by the University of Adelaide Library.

Page 37: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

39

female caught in Taiwan back in 1995 (Joung et al., 1996). When this specimen was

examined, more than 300 embryos were found (1:1 female:male ratio), possibly all

fertilised by the same male (Schmidt et al., 2010). They were at different stages of

development with some in egg cases with yolk sac (~ 40-50 cm), and others ready to

be born, free from egg cases and yolk sacs (~ 60 cm). This discovery led to the

conclusion that the species is ovoviviparous, and possibly monoandric and with

capacity to store sperm (Schmidt et al., 2010). Monoandry is usually observed when

mating is only occasional, which might be the reason behind our still limited knowledge

on mating, reproduction, and early life stages. Observation of neonates (≤ 1 m) is also

rare, with only a few records (<20; Rowat & Brooks, 2012) mostly caught by fishers in

the open ocean (e.g., Wolfson, 1983). Despite the rarity of such occurrences, reports

of neonates span the three major oceans (Rowat et al., 2008; Wolfson, 1983).

Whale sharks are found in the Atlantic, Indian and Pacific Oceans. Being

poikilothermes, their body temperature depends on the external water temperature,

and their range of occurrence is considered to be largely restricted to tropical and

temperate waters, between latitudes of 30° north and south (Rowat & Brooks, 2012).

Despite their widespread occurrence, they are mostly observed in specific near-shore

locations where they form seasonal aggregations for reasons that remain unclear, e.g.,

in Ningaloo (Taylor, 1996), Gulf of Mexico (Hoffmayer et al., 2007) or Gulf of California

(Nelson & Eckert, 2007). Several studies have focused on monitoring movement

patterns starting at each of these locations (Eckert & Stewart, 2001; Hueter et al.,

2008; Wilson et al., 2006). However, only a few have captured movement beyond the

aggregation areas (Eckert & Stewart, 2001; Hueter et al., 2008; Rowat & Gore, 2007),

and even fewer have re-sighted the same shark in subsequent seasons (Graham et

Page 38: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

40

al., 2006; Rowat et al., 2011). The few tracks showing movement beyond aggregations

have revealed that whale sharks are possibly able to travel between ocean basins

(Eckert & Stewart, 2001), but due to the paucity of tracking results obtained to date,

data are still insufficient to confirm whether there is a general, recurrent movement of

this species between major ocean masses. Although whale sharks spend most of their

time at the surface (Gunn et al., 1999), they can frequently ‘dive’ deeply (> 1000 m)

(Brunnschweiler & Sims, 2012; Rowat & Gore, 2007). The reasons behind their

movement behaviour are yet to be understood, but some have suggested they might

be associated with foraging and feeding (Brunnschweiler et al., 2009; Sleeman et al.,

2010a).

Aggregations of whale sharks usually coincide with higher concentrations of

plankton in surrounding waters (Heyman et al., 2001; Sleeman et al., 2007). Being

filter feeders, they generally only consume small secondary grazers (e.g., krill or crab

larvae; small fish) (Meekan et al., 2009; Wilson & Newbound, 2001), and therefore

other species are commonly associated with them, including commercially valuable

anchovy (Duffy, 2002) and herring (Wilson, 2002). This characteristic has been

particularly helpful in oceanic tuna fisheries where whale sharks are used to spot tuna

(Matsunaga et al., 2003) commonly aggregating under the shark (possibly foraging on

similar prey) – in other words, whale sharks are used as fish aggregating devices.

Despite the evidence for their ability to embark on large migrations, some

whale sharks are also commonly resighted within the same near-shore locations in

different years (Norman, 2007), such as in Ningaloo (Holmberg et al., 2008; Taylor,

1996), the Seychelles (Rowat et al., 2009b), and Belize (Graham & Roberts, 2007).

However, most of the sharks seen at these aggregations are immature males.

Females are observed mostly in the Gulf of California (Eckert & Stewart, 2001;

Page 39: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

41

Ramírez-Macías et al., 2007), and the Galapagos (Green, 2011), but they possibly

also occur in South East Asia (Taiwan; Joung et al., 1996), and where a 46-cm

newborn whale shark was recently seen (Philippines; Aca & Schmidt, 2011). Besides

this possible sex segregation, separation by age or size has also been postulated

(Ramírez-Macías et al., 2007; Rowat et al., 2011).

Due to their predictable aggregation behaviour, together with their harmless

characteristics, whale sharks have become of economically important in several

coastal locations around the world. Since 1993, when they first became the attraction

of tourism in Ningaloo, Western Australia (Davis et al., 1997), this source of income,

ranging from US$5 million (over a 14-week season in the Seychelles; Rowat &

Engelhardt, 2007) to US$50 million (per year in Ningaloo; ABC News, 2005), has

spread throughout aggregation areas. Whale shark-dedicated tourism now occurs in

the Philippines (Quiros, 2005), Seychelles and Djibouti (Rowat et al., 2011),

Mozambique (Pierce et al., 2010), Gulf of Mexico and Caribbean Sea (Heyman et al.,

2010; Quiros, 2005), and Gulf of California (Cárdenas-Torres et al., 2007).

In the past, whale sharks have been targeted by fisheries mostly in southeast

Asia where a single whale shark fin was valued at US$316 (White & Cavanagh, 2007).

Due to declining landings (e.g., Chen & Phipps, 2002), commercial fisheries for whale

sharks are now banned, with the last ban occurring in Taiwan as recently as 2008

(Council of Agriculture, 2007) (Articles 44 and 46 of the Fisheries Act,

www.fa.gov.tw/en/LegalsActs/index.aspx). However, some of these fisheries continued

their activity illegally (Riley et al., 2009). For a long time, whale sharks have also been

used to assist the catch of commercially valuable fishes (Iwasaki, 1970). But while

acting as fish aggregating devices, they are also caught (by-catch) in the tuna purse-

seine nets (Matsunaga et al., 2003). Although they are commonly released from these

Page 40: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

42

nets (Chassot et al., 2009), the extent of damage is unknown and survival rates have

never been assessed.

Since 1990, whale sharks have been listed on the IUCN Red List of

Threatened Species (IUCN, 2010), and since 2000 are classified as Vulnerable (i.e., at

risk of extinction in the wild). Almost 200 years after the species was first described,

basic knowledge of these sharks is still lacking, as is our understanding of the

consequences of human activities on their long-term persistence. While impacts from

targeted fisheries soon became evident through declining landings, the effects of

current fishing (by-catch and illegal) are difficult to measure. The expansion of tourism

at both new and previously known aggregations, might also have detrimental effects

(Heyman et al., 2010), especially if management is poor.

Changes observed in whale shark populations from near-shore locations have

been associated with a general decline (Bradshaw et al., 2008). With the broad-range

of occurrence of whale sharks, pressure to develop large-scale conservation

measures has increased (Rowat, 2007), however the currently unknown connectivity

among aggregations (Figure 3) makes the assessment of the regional conservation

status of whale sharks challenging. Recent genetic evidence demonstrated high

connectivity between whale sharks from populations among different oceans (Castro

et al., 2007; Schmidt et al., 2009); hence, one can hypothesise that whale sharks form

a single, global population at least over generational scales. Most whale shark studies

to date have focused on single aggregation locations, and are therefore strongly

biased towards specific size, gender and habitat classes. Consequently, they have

failed to capture global patterns of whale shark population ecology and connectivity.

Page 41: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

43

Figure 3: Whale shark distribution range and sighting reports known as of 2009. Black dots represent locations where whale sharks seasonally aggregate. N: free swimming neonates < 1m (Rowat et al., 2008; Wolfson, 1983); E: large female with 300 embryos (Joung et al., 1996); J: group of juveniles (~ 1 m) swimming with an adult whale shark (Pillai, 1998).

Page 42: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

44

Outline of the thesis

My thesis describes and quantifies the global patterns of whale shark population

distribution and connectivity for the first time, contextualising them within the theme of

global change. Besides presenting a challenging case study in the context of climate

change and likely shifts in marine species distribution, the elusive whale shark also

represents a good model to understand biotic responses to oceanographic patterns;

being a filter feeder, they should track primary production more closely than higher-

order predators (Grémillet et al., 2008).

The main component of my thesis begins with formulating a hypothesis on the

possible global connectivity of whale sharks. This is presented as a review (Chapter II)

where I examined from a global perspective all of the whale shark literature and

distributional data available to date, and then summarised this in a conceptual model

of global whale shark connectivity. In this review I pointed out the major gaps in

knowledge, and suggested some ways to fill these gaps. Later, I validated or explored

these hypothetical links in subsequent chapters (Chapters III, IV, V and VI).

With most publications and data available for the Indian Ocean, I began here

for my global analysis. In my first data chapter (Chapter III), I predicted the spatial and

seasonal patterns in whale shark distribution within the Indian Ocean based on

fisheries-collected observational data. Following this spatial analysis, and the current

concerns about declining abundance (Bradshaw, 2007; Bradshaw et al., 2008), I

assessed temporal trends in the Indian Ocean whale shark population, and compared

the variation in whale shark sightings with the variation in climatic signals (Chapter IV).

To test if my hypotheses regarding global connectivity are biologically realistic on a

temporal scale, I analysed both the spatial and temporal patterns of whale shark

occurrence also in the Atlantic and Pacific Oceans, again based on fisheries-collected

Page 43: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

45

observational data (Chapter V). My goal here was to look for asynchronies in the

probability of whale shark occurrence in the different oceans. This is to be expected if

whale sharks are travelling between oceans. In the last data chapter (Chapter VI), I

built a global distribution model to assess current global patterns of whale shark

habitat suitability. Here I also projected whale shark distribution under a climate

change scenario to provide the first quantitative prediction of how increasing sea

surface temperatures might affect the distribution of this species. Finally, I revisit the

general context in which this thesis is inserted to reveal the novelties of my approach. I

point out how to overcome problems associated with incomplete, and often biased,

data on highly migratory marine species, to generate appropriate statistical

approaches which allow assessment of the possible climate change impacts on these

species.

All my data chapters were written as ‘stand-alone’ manuscripts, and therefore

some of the material presented in the Introduction and Methods sections are repeated

in each chapter. These manuscripts are now published (Chapter III), in press (Chapter

II and IV), submitted (Chapter VI) or in preparation for submission (Chapter V).

The main objectives of my Ph.D. were to:

1. provide a modelling framework to allow assessment of possible

climate change impacts on wide-ranging elusive marine species

(using the whale shark as a case study);

2. analyse the current knowledge about whale shark at a global scale

and show how they might be connected within an hypothesised global

meta-population;

Page 44: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

46

3. assess the influence of environmental conditions on whale shark spatial

distribution at a regional and global scale;

4. determine temporal trends in whale shark occurrence in the three major

oceans and assess asynchrony in occurrences (as expected for a global

migrating population) or evidence of declining numbers; and

5. derive realistic predictions of this species current geographical range and

provide a modelling tool for exploring whale shark distribution under current

and future climate conditions.

Page 45: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

47

CHAPTER II. Inferred global connectivity of whale shark populations

A. Sequeiraa, C. Mellina,b, M. Meekanc, D. Simsd, C. Bradshawa,e

aThe Environment Institute and School of Earth and Environmental Sciences, The University of Adelaide, South Australia 5005, Australia bAustralian Institute of Marine Science, PMB No. 3, Townsville MC, Townsville, Queensland 4810, Australia cAustralian Institute of Marine Science, UWA Oceans Institute (MO96), 35 Stirling Hwy, Crawley, Western Australia 6009, Australia. dMarine Biological Association of the United Kingdom, The Laboratory, Citadel Hill, Plymouth PL1 2PB, UK dSouth Australian Research and Development Institute P.O. Box 120, Henley Beach,

South Australia 5022, Australia

Journal of Fish Biology 2013; 82:2, 595 – 617

Page 46: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

48

Page 47: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

49

Statement of authorship

INFERRED GLOBAL CONNECTIVITY OF WHALE SHARK POPULATIONS

Journal of Fish Biology 2013; 82:2, 595 – 617

SEQUEIRA, A.M.M. (Candidate) Planned the article, performed the analysis, interpreted data, wrote the manuscript and acted as corresponding author. I hereby certify that the statement of contribution is accurate.

Signed ………………………………………………………………….Date…………………

MELLIN, C. Helped with data interpretation and provided critical evaluation. I hereby certify that the statement of contribution is accurate and I give permission for the inclusion of the manuscript in the thesis.

Signed ………………………………………………………………….Date…………………

MEEKAN, M. G. Supervised development of work and provided evaluation. I hereby certify that the statement of contribution is accurate and I give permission for the inclusion of the manuscript in the thesis.

Signed …………… ................................Date…………………

Page 48: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

50

SIMS, D.W. Provided evaluation of the manuscript. I hereby certify that the statement of contribution is accurate and I give permission for the inclusion of the manuscript in the thesis.

Signed ………………………………………………………………….Date…………………

BRADSHAW, C.J.A. Supervised development of work, helped in data interpretation and with manuscript evaluation and writing. I hereby certify that the statement of contribution is accurate and I give permission for the inclusion of the manuscript in the thesis.

Signed ………………………………………………………………….Date…………………

Page 49: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

51

Inferred global connectivity of whale shark populations

Abstract

Ten years have passed since the last synopsis of whale shark (Rhincodon typus, Smith

1828) biogeography. While a recent review of the species’ biology and ecology

summarized the vast data collected since then, it is clear that information on population

geographic connectivity, migration and demography of whale sharks is still limited and

scattered. Understanding whale shark migratory behaviour is central to its conservation

management considering the genetic evidence suggesting local aggregations are

connected at the generational scale over entire ocean basins. By collating available data

on sightings, tracked movements and distribution information, our review provides

evidence for the hypothesis of broad-scale connectivity among populations, and generates

a model describing how the world’s whale sharks might be part of a single, global meta-

population. Whale shark occurrence timings and distribution patterns make possible a

connection between several aggregation sites in the Indian Ocean. Our conceptual model

and validating data lend support to the hypothesis that whale sharks are able to move

among the three largest ocean basins with a minimum total travelling time of around two to

four years. Our model provides a worldwide perspective of possible whale shark migration

routes, and suggests a modified focus for additional research to test its predictions. Our

framework can be used to trim the hypotheses for whale shark movements and

aggregation timings, thereby isolating possible mating and breeding areas that are

currently unknown. This will assist endeavours to predict the longer-term response of the

species to ocean warming and changing patterns of human-induced mortality.

Page 50: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

52

KEYWORDS: meta-population; migration; movement; sea surface temperature;

Rhincodon typus

Introduction

The large and docile whale shark (Rhincodon typus, Smith 1828) is a filter-feeding

chondrichthyan that can reach over 18 m in total length (Borrell et al., 2011; Chen et al.,

1997; Compagno, 2001), making it the largest extant fish species. Its geographic range is

more less known across all tropical and warm temperate waters (Compagno, 2001; Last &

Stevens, 2009), especially around sites where individuals aggregate seasonally. Ironically,

little is known about how these aggregations are connected via migration, apart from

preliminary population genetic evidence suggesting connectivity at the generational scale

(Castro et al., 2007; Schmidt et al., 2009). Further, their pelagic distribution is poorly

described, as is the true extent of their range. Rising concerns about how warming seas

will affect whale shark distributions are potentially being manifested already; for example,

occasionally individuals are sighted at latitudes higher than their nominal 30 º – in the Bay

of Fundy, Canada 44 º N (Turnbull & Randell, 2006), in the southern Azores islands,

Portugal 41 º N (Sa, 2008) or in the north-east of New Zealand > 35 º S (Duffy, 2002). This

lack of understanding of their distribution, connectivity and migration pathways therefore

severely limits our capacity to broker international conservation plans for the species.

Whale shark sightings are most often reported near shore because they often

aggregate in specific locations along the coast within warm/temperate waters at

approximately the same time every year (e.g., Rowat, 2007). The incentive for aggregating

is not yet fully understood, but most studies suggest that such occurrences are related to

Page 51: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

53

food blooms, currents or temperature variation (Cárdenas-Palomo et al., 2010; Colman,

1997; Kumari & Raman, 2010; Meekan et al., 2009; Sleeman et al., 2010b; Wilson, 2002).

In contrast, whale sharks are rarely recorded on the high seas, largely because of the

challenges of sampling the ocean realm (Richardson & Poloczanska, 2008). Due to the

drive to chase ever-more-elusive commercial fish species, the fishing industry seems to be

the only sector capable of recording marine species regularly in the open ocean (Sequeira

et al., 2012). Whale shark-targeted fisheries ran mainly in South-East Asia and India prior

to the turn of the last century (Joung et al., 1996; Pravin, 2000; Stevens, 2007; White &

Cavanagh, 2007). As a consequence of declining landings, the species was classified as

Vulnerable in 2000 in the IUCN Red List (www.iucnredlist.org) and the fisheries were

eventually banned – the last bans occurred in India in 2001 (as per the Indian Wildlife

Protection Act, Schedule I, amended in 2001) and Taiwan in 2007 (Council of Agriculture,

2007). Even though most of the bans happened more than a decade ago, a recent review

of this classification resulted in no change of status (IUCN, 2010). Together with the

Taiwanese commercial catch lasting until 2007, the persistent high threat risk of the

migratory whale shark (UNCLOS, 1982) might be due to a low scrutiny of the ban’s

implementation (Stewart & Wilson, 2005), and continued illegal fishing in areas such as

eastern Indonesia (White & Cavanagh, 2007) and the Maldives (Riley et al., 2009). The

unintentional catch of whale sharks is also occurring in other fisheries of large ocean

coverage (e.g., purse seiners; Romanov, 2002).

Whale shark-based tourism has proven the worth of live sharks (e.g., Anderson &

Waheed, 2001), with related incomes replacing those once provided by targeted fisheries

(Quiros, 2007). Established tourism operations continue to expand throughout the world at

Page 52: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

54

known whale shark aggregation sites, such as in the Seychelles (Rowat & Engelhardt,

2007), Australia (Catlin & Jones, 2009) and Mexico (de la Parra Venegas et al., 2011), and

they play an important conservation role because of the positive example of economic

benefits from renewable tourism. Tourism also assists whale shark research because

photographs taken during dive operations are used to identify returning individuals (Speed

et al., 2007), thus providing data essential for estimating abundance, vital rates and

persistence probability (Bradshaw, 2007; Meekan et al., 2006; Rowat et al., 2009a). By

collating tourism-derived photographic data from different locations, research is currently

pursuing evidence of inter-aggregation migration hypothesised to occur given the finding

of ocean-scale genetic connectivity (Castro et al., 2007; Schmidt et al., 2009); however,

none has yet been found (Brooks et al., 2010).

Examining whale sharks at the scale of single aggregations cannot adequately

describe the species’ life history because it encapsulates only a small proportion of the life

stages. Collecting data outside aggregation areas is therefore essential. Continual

developments in tagging technology to improve estimates of home range size, movement

patterns and habitat use (Hammerschlag et al., 2011) have been partially successful in

this regard, although despite > 3000 whale shark tagged to date (e.g.,

www.whaleshark.org), trajectories have not revealed reliable evidence for inter-

aggregation connection. This is not entirely surprising given the low probability of

resighting migratory individuals in widely spaced aggregations. The lack of bio-logging

efficiency is mostly due to premature detachment and limited spatial coverage of acquired

data (e.g., Brunnschweiler et al., 2009), tag removal and/or damage (Fitzpatrick et al.,

2006; Hays et al., 2007) which might result in part from attacks from other sharks or killer

Page 53: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

55

whales (Fitzpatrick et al., 2006; Speed et al., 2008), and the accumulation of bio-fouling

organisms causing tag malfunction (Hays et al., 2007). Moreover, even though tracking

methods can provide some insight into the movement behaviour of whale sharks, they are

still unlikely to encompass the full range of the population’s distribution.

To investigate putative links among aggregations, mitochondrial and microsatellite

DNA studies have revealed that whale sharks from the three major ocean basins have low

genetic differentiation (Schmidt et al., 2009), albeit sharks in the Atlantic Ocean had

moderately different mitochondrial haplotype frequencies from those in the Indian and

Pacific Oceans (Castro et al., 2007). These results partially validate the prediction that

whale sharks have at least generational-scale migrations that connect populations among

the world’s oceans. According to Hartl and Clark (1989), population differentiation can be

prevented with only a few breeding migrants per generation, and panmixia can occur even

with only about four breeding migrants per generation (Hartl & Clark, 1989; Morjan &

Rieseberg, 2004). Given the species’ long generation time (ranging from 15 to 37 years;

Bradshaw et al., 2007), only rare dispersal would be required to demonstrate equivalent

gene flow. This suggests that current bottom-up approaches based on collecting difficult-

to-obtain tagging data to estimate vital rates and life history traits have a low probability of

characterizing broad-scale migratory patterns. Since Smith first described the species in

1828, we still know next to nothing about their physiology (e.g., growth rates) or

reproduction (e.g., mating areas, variation in pup production, breeding frequency). The

only clue that whale sharks are ovoviviparous was provided by a single pregnant female

specimen carrying 300 embryos in different stages of development (Joung et al., 1996).

Page 54: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

56

In this overview and analysis, we propose instead a top-down approach to infer

whale shark occurrence probability and global patterns of movement. We suggest that the

primary focus of research should be shifted to broader-scale approaches that generate

hypotheses based on conceptual models of migration. We argue that testing these

hypotheses indirectly through modelling approaches, or directly via improved technologies

that provide higher-quality and longer-term migration data, will ultimately provide a more

realistic picture of the species’ distribution throughout its life stages. This will assist in

making predictions on population viability and redistribution resulting from warming oceans

and changing patterns of human-induced mortality.

Knowledge base

Our understanding of whale shark ecology and biology has accelerated from an average

of < 3 papers year-1 between 1992 and 2005 to a mean of 15.7 year-1 over the following 6

years (source: Thomson-Reuters ISI Web of Knowledge consulted 23 March 2011 using

“whale shark” and “Rhincodon typus” as key words). A peak in the number of papers

occurred in 2007 when a special issue on the subject was published (Fisheries Research,

2007) following the first International Whale Shark Conference in Perth, Western Australia

(Irvine & Keesing, 2007).

Since whale sharks were first identified by Andrew Smith in Table Bay, South

Africa (Smith, 1828), E. W. Gudger has still published the highest number of peer-

reviewed papers on the species, mainly describing occurrences in Florida, the Gulf of

California (Sea of Cortés), Seychelles, Galapagos, Hawaii, The Bahamas and the

Caribbean Sea (as reviewed in Gudger, 1934), and was the first to attempt to portray

Page 55: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

57

whale shark “migratory behaviour” partially as a function of ocean current patterns (e.g.,

Gudger, 1932, 1934). Iwasaki (1970) attempted to surmise the distribution of whale sharks

based on data collected by Japanese tuna fishing vessels, from which he concluded that

their occurrence around Japan was seasonal and depended mainly on the behaviour of

the Kuroshio Current, temperature and prevailing winds. Uchida et al. (1984) suggested

that whale sharks are annually regular in their appearance on the east coast of Taiwan,

but that seasonality in presence on the west coast depended on temperature. Later, Taylor

(1996) published the first paper describing the regular autumn aggregation of whale

sharks at Ningaloo Reef, Australia, and associated it with possible zooplankton blooms

following the coral spawn. Since then, Ningaloo Reef has been the site of the most

targeted studies (Figure 4), even if much research has been done elsewhere relating

mainly to occurrence, movement, tourism, mortality and feeding behaviour (Figure 4 and

Table A1 in Appendix A). Few data are available on whale shark life-history traits such as

reproduction (Chang et al., 1997; Joung et al., 1996; Schmidt et al., 2010) and growth

(Pauly, 1997; Wintner, 2000), or internal biology (Dove et al., 2010; Wilson & Martin, 2001;

Yopak & Frank, 2009), which reflects the difficulty of obtaining specimens for detailed

examination and measurement. From the summary of recent studies per site (Figure 4;

Table A1), the following trends emerge: (1) knowledge of whale shark ecology is most

advanced in the Indian Ocean, and (2) similar work has been done in the different

locations worldwide, but (3) mostly at the scale of individual aggregations.

Recent genetic studies (Castro et al., 2007; Schmidt et al., 2009) provide the only

clue so far that the world’s whale shark populations are connected via dispersal occurring

during some stage of their life. However, there has been no effort so far to synthesise all

Page 56: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

58

the disparate information available into a conceptual model of dispersal and population

connectivity, even after the conservation of whale sharks was highlighted as requiring both

“regional cooperation and conservation initiatives” (Rowat, 2007).

Figure 4: Graphical illustration of the number of representative whale shark studies (since 1997). Grouped by (1) occurrence, population and correlates for occurrence, (2) movement, (3) tourism and conservation, (4) mortality and (5) foraging, in different locations worldwide. References provided in Table A1 in Appendix A. .

Page 57: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

59

Global occurrences

The first 76 recorded whale shark occurrences worldwide were compiled in the early

1930s by Gudger (e.g., Gudger, 1932). He suggested the Sulu Sea (between southern

Philippines and northern Borneo) as the single location from which all whale sharks

originate, dispersing from there depending on environmental conditions (Gudger, 1934).

After Gudger, Wolfson and Notarbartolo-di-Sciara (1981) counted 345 records worldwide

(including repeat sightings) until 1980. Currently, whale sharks have been recorded in the

three major oceans spanning the Equator and have been seen, even if only occasionally,

near the shore of over 100 countries in five continents (Figure 5) (Compagno, 2001;

Martin, 2007; Rowat, 2007; Stacey et al., 2008).

Although existing photo-identification libraries have yet to report a match from

sharks seen at distant locations, there is some evidence for short-distance movements.

This is the case for resights of the same shark in Gladden Spit (Belize), Isla Contoy

(Mexico) and Utila (Honduras) (Graham & Roberts, 2007), and a resighting in

Mozambique of a shark tagged in the Seychelles 11 months earlier (Rowat & Gore, 2007).

Peaks in whale shark occurrence appear to happen synchronously in different

locations around the world (Figure 6). For example, in January they occur in KwaZulu-

Natal (South Africa), Djibouti and Christmas Island (Australia); in March to May, they occur

at Gladden Spit (Belize), Gujarat (India), Ningaloo (Australia) and around the Philippines;

in August to October, whale sharks have been reported in Portugal (around the Azores

islands), Mozambique, Seychelles and Bahia de Los Angeles (Gulf of California, Mexico).

Page 58: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

60

Figu

re 5:

Com

pilat

ion

of w

orld

wide

wha

le sh

ark o

ccur

renc

es.

From

Com

pagn

o (2

001)

, Mar

tin (2

007)

, St

acey

(200

8) an

d Ro

wat (

2007

) ove

rlaid

wi

th th

e IUC

N (2

001)

di

strib

utio

n (h

atch

ed

area

). Sq

uare

size

re

pres

ents

size

cla

sse,

and

colo

urs

repr

esen

t gen

der

(whe

re av

ailab

le):

blue

= m

ale, r

ed =

fe

male

and

blac

k =

gend

er n

ot sp

ecifi

ed.

Squa

re d

ots r

epre

sent

ju

veni

le wh

ale sh

ark

aggr

egat

ions

and

the

�� re

pres

ents

the

sight

ing

of p

regn

ant

fem

ale w

ith 30

0 em

bryo

s (Jo

ung

et al

., 19

96).

Small

blac

k do

ts re

pres

ent l

ong-

term

(198

0 to

2010

) wh

ale sh

ark s

ight

ings

fro

m tu

na fi

sher

ies.

Page 59: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

61

However, when examining the timings over several years, a pattern of sequential monthly

peaks at neighbouring locations emerges (Figure 6).

A highly skewed sex ratio is common in certain aggregations (e.g., Ningaloo,

Seychelles, Mozambique, Belize) where mostly immature males are found (Graham &

Roberts, 2007; Meekan et al., 2006; Rowat et al., 2011). Rowat et al. (2001) provided

some evidence for size segregation, stating that individuals observed in the Djibouti

aggregation (mainly immature males) are smaller than those from the Seychelles (and

other locations in the Indian Ocean). They suggested that the Djibouti population might be

a ‘staging’ group for other regional aggregations (i.e., segregation might therefore be

another feature contributing to migration among sites). In the Gulf of California (Mexico),

whale sharks segregate both by sex (Eckert & Stewart, 2001) and size, with larger adult

females mainly seen in the southern part of the Sea of Cortés (Ramírez-Macías et al.,

2007).

Smaller whale sharks (< 2.5 to 3 m total length) have been observed only rarely

(Colman, 1997) and almost all neonates reported so far have been caught as by-catch in

purse-seine fisheries (Wolfson, 1983). The smallest neonate ever recorded (46 cm total

length) was recently found swimming freely in Sorsogon, Philippines (Aca & Schmidt,

2011), and the only fertility data available for the species is derived from a single, 10.6 m-

long female caught off Taiwan containing 300 ‘ready-to-be-born’ embryos measuring 58 to

64 cm (Joung et al., 1996).

Page 60: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

62

Tracking studies

Gunn et al. (1999) and Eckert and Stewart (2001) published the first whale shark tracking

results derived from satellite tags deployed at Ningaloo Reef (Western Australia) and the

Sea of Cortés (Gulf of California), respectively. Despite the > 3000 data-logging and

satellite tags deployed to date (www.whaleshark.org, consulted 5 August 2011), we found

only 109 whale shark tracks publicly available (Table 1, Table 2 and Table 3), with only 75

published in peer-reviewed papers.

Most of the tracking results available are from the Indian Ocean (47 tracks), with

Ningaloo being the main tagging location. Next was the Pacific Ocean from where 30

tracks have been published (16 starting at the Sea of Cortés and 14 from the Galapagos

Islands), followed by 22 in the Atlantic Ocean deployed mostly in Belize, and finally 10 in

the Indo-Pacific area (from the Philippines, Malaysia and Taiwan). The sex ratio of these

tracked individuals is 33 females:31 males:45 unknown (Tables 1-3). Average track

duration is around 90 days, although they range from only a few hours (Gunn et al., 1999)

to more than 3 years (Eckert & Stewart, 2001), with about 30 % of the tracks lasting < 1

month and only 2 lasting > 9 months. Distances travelled varied between 4 and 12,620 km

and averaged about 1,545 km, with only 10 sharks travelling > 2,000 km.

Page 61: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

63

Fi

gure

6: G

loba

l seq

uenc

e of w

hale

shar

k mon

thly

appe

aran

ce ti

min

gs (3

year

s pre

sent

ed) i

n di

ffere

nt lo

catio

ns w

orld

wide

. Lo

catio

ns ar

e ord

ered

geo

grap

hica

lly fr

om th

e nor

th A

tlant

ic Oc

ean,

thro

ugh

the I

ndian

Oce

an to

the e

ast a

nd w

est P

acifi

c Oce

an. D

ark r

ed re

pres

ent t

he se

ason

al pe

aks,

dark

pin

k re

pres

ent t

he fu

ll sea

son,

and

light

pin

k rep

rese

nt o

ccas

iona

l sig

htin

gs. G

rey

area

s rep

rese

nt la

rge d

istan

ces b

etwe

en th

e loc

atio

ns ab

ove a

nd b

elow

the g

rey c

ells.

1 Mac

hete

and

Afon

so, p

ers.

com

m.;

2 Mot

ta et

al., 2

010;

3 Hey

man

et al

., 200

1 and

Gra

ham

& R

ober

ts, 2

007;

4 Har

born

e et a

l., 20

01; 5 H

azin

et al

., 200

8; 6 B

eckle

y et a

l., 19

97, C

liff e

t al.,

2007

and

Rowa

t, 20

07; 7 P

ierce

et al

., 201

0 and

Giff

ord

et al

., 200

7; 8 J

onah

son

& Ha

rdin

g, 20

07; 9 R

owat

& E

ngelh

ard,

2007

, Row

at et

al., 2

011 a

nd R

owat

, 200

7; 10

, 11, 12

Row

at, 2

007;

13 R

owat

et al

., 201

1 an

d Ro

wat e

t al.,

2007

(ane

doct

al re

ports

); 14

Pra

vin, 2

000;

15 R

owat

, 200

7; 16

And

erso

n &

Ahm

ed, 1

993;

17 A

nder

son

& Ah

med

, 199

3; 18

Pra

vin, 2

000;

19, 2

0 Row

at, 2

007;

21 H

obbs

et al

., 20

09; 22

Tay

lor,

1996

; 23 Q

uiro

s, 20

07; 24

, 25 C

hang

et al

., 199

7; 26

, 27, 28

Iwas

aki, 1

970;

29 G

unn

et al

., 199

2 in

Colm

an, 1

997;

30 D

uffy

, 200

2; 31

Nels

on &

Eck

ert,

2007

and

Card

enas

-Tor

res e

t al.

, 200

7; 32

Gre

en, 2

011.

Page 62: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

64

Table 1: List of published whale shark tracks from the Indian Ocean. * indicates tracks recreated and displayed in Figure 7; numbers in superscript indicate the reference for each track where there are multiple references for the same tagging location.

# Length (m)

Sex Tagging Date

Pop-up Date

Duration (days)

Distance (km)

Mean speed (km/day)

South Africa - Cape Vidal (Gifford et al., 2007a) 1* 7 F 08.03.1998 25.03.1998 17 199 - 2* 7 M 25.10.1998 01.11.1998 7 90 - 3* 8 M 23.01.1999 25.01.1999 < 3 65 - Mozambique - Tofo (Brunnschweiller et al., 2009) 4* 6-7 F 18.02.2006 16.5.2006 87 1200 31.2 Madagascar – Nosy Be (Graham et al., 2008) 5 - - 5-12.12.2005 15.04.2006 - - - 6 - - 5-12.12.2005 07.02.2006 - - - 7 4 M 5-12.12.2005 24.11.2005 - - - 8 8 F 5-12.12.2005 19.12.2005 - - - 9 - - 5-12.12.2005 15.03.2006 - - - Djibouti - Arta Beach (Rowat et al., 2007) 10 3 M 22.01.2006 31.01.2006 9 - 10.1

Arabian Sea (MML, 2010) 11* 4.2 F 18.03.2010 20.04.2010 33 - - Seychelles - Isle of Mahe (Rowat & Gore, 2007) 12* 5 - 02.09.2001 17.10.2001 46 1422.74 33.1 (max) 13* 7 - 25.10.2001 12.11.2001 19 502.1 23.8 (max)

14* 6.5 M 25.10.2001 23.12.2001 60 3382.62 56.9 (max) Western Australia - Ningaloo (2CSIRO, 2005; 1Gunn et al., 1999; 3Wilson et al., 2006; 4Wilson et al., 2007) 151 - - 28.03.1994 - <1 - 60.5 161 - - 29.03.1994 - <2 - 51.3 171 - - 06.04.1997 - <2 - 64.8 181 - - 10.04.1997 - <1 - 73.5 19*2 7 F 22.04.2002 17.07.2002 85 3130 36 20*2 4.2 M 05.05.2005 29.09.2005 147 - - 21*2 4 F 06.05.2005 19.09.2005 136 - - 22*2 7.5 - 05.05.2005 30.08.2005 118 - - 23*2 4.2 M 07.05.2005 10.06.2005 34 - - 24*2 - - 02.05.2005 10.06.2005 39 - - 25*2 4.2 - 06.05.2005 04.06.2005 29 - - 26*2 7.3 M 28.06.2005 31.07.2005 33 1740 52.8 273 8-10 F 02.05.2003 10.05.2003 8 85 -

Page 63: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

65

283 6 F 03.05.2003 04.05.2003 1 16 - 293 4.5 F 05.05.2003 - - - - 30*3 6-8 - 05.05.2003 11.11.2003 191 942 - 313 4.5 M 03.05.2004 03.05.2004 0 4 - 323 8.5 F 03.05.2004 03.05.2004 0 7 - 333 7.5 F 03.05.2004 07.05.2004 4 93 - 343 4.7 F 04.05.2004 - - - - 353 4.5 F 04.05.2004 - - - - 36*3 11 F 04.05.2004 31.08.2004 119 1302 - 373 7.3 F 04.05.2004 - - - - 38*3 4.7 - 05.05.2004 24.11.2004 203 673 - 393 7.6 F 05.05.2004 - - - - 403 5.6 F 07.05.2004 - - - - 413 7.6 F 07.05.2004 10.05.2004 3 16 - 42*3 6.2 F 08.05.2004 06.09.2004 119 738 - 43*3 7.2 M 08.05.2004 04.07.2004 57 1501 - 443 7 F 08.05.2004 - - - - 45*3 5.3 F 09.05.2004 11.12.2004 216 443 - 464 4-5 M 05.05.2005 29.09.2005 147 - - Christmas Island (Australian Institute of Marine Sciences ^) 47* 5 M 17.01.2008 14.04.2008 88 - - ^ whale shark tagged by M.M., unpublished.

Page 64: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

66

Table 2: List of published whale shark tracks from the Pacific Ocean and Indo-Pacific area. * indicates tracks displayed in Figure 7.

# Length (m)

Sex Tagging Date

Pop-up Date

Duration (days)

Distance (km)

Mean speed (km/day)

Hainan – China (MCSS, 2009) 1* - M 28.06.2009 31.07.2009 33 - - Taiwan (Hsu et al., 2007) 2* 4.5 M 08.04.2002 01.11.2002 208 5896 28.3 3 4.2 F 17.04.2002 - - - - 4* 4.2 M 13.11.2002 28.02.2003 108 3732 34.6 5* 4 M 09.04.2004 01.08.2004 114 3121 28.41 Philippines - Salay (Eckert et al., 2002) 6 7 - 05.02.1997 - 10 900 18.9 7* 3 - 18.02.1997 02.05.1997 73 1567 31.9 Philippines – Donsol (Eckert et al., 2002) 8* 5 - 23.02.1997 03.03.1997 8 54 13.7 Malaysia - Usukan Island (Eckert et al., 2002) 9* 7 - 03.02.1997 9.02.1997 6 220 6.9 10* 7 - 05.02.1997 12.06.1997 127 8025 23.3

11* 7 - 05.02.1997 15.02.1997 10 900 18.9 Gulf of California – Sea of Cortez (Eckert & Stewart, 2001) 12* 3.7 F 06.10.1994 07.10.1994 1 57.5 - 13* 4.6 - 22.9.1994 14.10.1994 22 - - 14* 3.0 F 19.9.1994 06.10.1994 17 320.3 21.7 15 4.6 - 20.9.1994 - 17 31.7 3.2 16* 4.3 - 20.9.1994 29.10.1994 39 818.3 23.5 17 4.0 - 23.9.1994 - 5-7 23.3 8.4 18 4.0 - 07.10.1994 - 1 11.2 - 19* 6.1 F 10.9.1995 22.09.1995 12 404.6 23.6 20* 3.7 - 10.9.1995 13.09.1995 3 8.5 11.1 21* 7.1 - 12.9.1995 26.11.1998 1144 12620 17.1 22* 3.7 F 10.9.1995 28.09.1995 18 46.3 2 23* - - 09.9.1995 07.10.1995 28 199.6 - 24 15.0 F 20.6.1996 - 111 2863.6 28.8 25* 18.0 F 19.6.1996 20.07.1996 30 206.8 18.2 26* - F 19.6.1996 15.04.1998 665 7762 23.3 Galapagos – aTotal of 14 tracks (Green, 2011) 27*a - - 07.2011 - - - -

Page 65: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

67

Table 3: List of published whale shark tracks from the Atlantic Ocean. * indicates tracks displayed in Figure 7; numbers in superscript indicate the reference for each track where there are multiple references for the same tagging location.

# Length (m)

Sex Tagging Date

Pop-up Date

Duration (days)

Distance (km)

Mean speed (km/day)

Gulf of Mexico - Gladden Spit (1Graham et al., 2006); Utila, Honduras (2Gifford et al., 2007b); Florida (3MML, 2010); Quitana Roo (4Hueter et al., 2008) 11 5.5 M 23.04.2000 07.05.2000 14 - - 21 3.6 M 23.04.2000 02.06.2000 40 - - 31 9.7 M 11.04.2001 15.10.2001 188 - - 41 6.7 M 11.04.2001 removed 248 - - 51 6.7 M 21.04.2000 06.05.2000 14 - - 61 5.5 - 25.04.2000 08.08.2000 105 - - 71 5.5 - 25.04.2000 15.06.2000 40 - - 81 5.2 M 16.03.2001 31.07.2001 127 - - 91 6.7 M 11.04.2001 30.06.2001 187 - - 101 5.2 M 11.04.2001 - 163 - - 111 5.5 - 10.05.2001 03.07.2001 249 - - 12*2 8 M 18.02.1999 29.06.1999 132 - - 13*2 8 M 31.12.1999 31.01.2000 31 - - 14*3 7.5 F 28.05.2010 20.08.2010 84 - - 15*3 7.5 M 18.06.2010 04.10.2010 108 - - 16*4 6 M 31.08.2005 01.10.2005 31 889 28.7 17*4 5.5 M 02.09.2005 01.10.2005 29 260 9.0 18*4 7 M 23.07.2006 27.12.2006 157 970 6.2 19*4 8 M 13.09.2006 22.10.2006 39 712 18.3 20*4 7 F 13.09.2006 25.11.2006 73 1027 14.1 21*4 8 F 14.09.2006 10.01.2007 118 371 3.14 22*4 7.5 F 31.08.2007 24.01.2008 150 7213 48.1

Compiling all published tracks at a global scale (Figure 7) shows that the current

information on whale shark movement range and habitat use is still highly scattered. In the

Atlantic ocean for example, there are tracks in the area of the Gulf of Mexico (Gifford et al.,

2007b) that confirm the link between Honduras, Belize and Mexico suggested by Graham

and Roberts (2007). There is also a > 7000 km track from a tag deployed at Quintana Roo

(Mexico) finishing in the central Atlantic Ocean south of the Saint Peter and Saint Paul

Archipelago (Brazil) about 5 months later (Hueter et al., 2008). However, there are no

Page 66: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

68

tracks available from locations in the eastern Atlantic where whale sharks have been

reported (e.g., Portugal or the west coast of Africa – Figure 5). In the Indian Ocean,

existing tracks link different aggregations. This is the case for South Africa and

Mozambique (Gifford et al., 2007b), Mozambique and Madagascar (Brunnschweiler et al.,

2009), United Arab Emirates and Qatar (MML, 2010) and Ningaloo with Timor, Indonesia

and Christmas Island (CSIRO, 2005; Wilson et al., 2006; Wilson et al., 2007). There is

also a 60-day track suggesting a connection between Seychelles and Sri Lanka (Rowat &

Gore, 2007), although it was inferred based on only one transmission obtained after an

inactive period of 1.5 months (represented as a straight line in Figure 7). Likewise, the

tags deployed in western Philippines (Eckert et al., 2002) show a possible connection with

Malaysia and south Vietnam, while sharks tagged in Taiwan (Hsu et al., 2007) reached

waters off east Philippines.

Several tags deployed in the Gulf of California (Eckert & Stewart, 2001) showed

that most sharks remain in the Sea of Cortés; however, four left the Gulf and one was

reported to have travelled almost 13,000 km from the east to the west Pacific Ocean in

1144 days (< 37 months). Eckert and Stewart (2001) asserted that due to data-storage

limitations, sharks with longer tracks (and not surfacing regularly enough to transmit stored

data), were located less frequently. However, whale sharks spend most of their time at the

surface (Sleeman et al., 2010a; Sleeman et al., 2010b; Wilson et al., 2006), so such long

periods without surfacing for satellite position acquisition are unlikely. For this reason, the

37-month track with an atypically straight trajectory, lack of surface intervals, and

unprecedented rate and magnitude of travel, should be deemed biologically unrealistic

Page 67: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

69

Figu

re 7:

Glo

bal o

verv

iew o

f pub

lishe

d wh

ale sh

ark t

rack

s. (C

olou

r cod

e ind

icate

s are

a of t

ag d

eplo

ymen

t. Tr

acks

adap

ted

from

orig

inal

sour

ces a

s det

ailed

in T

ables

1-3.)

Page 68: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

70

(e.g., it is more likely that the satellite tag had detached and was floating passively with the

currents). Recently, some tags deployed near the Galapagos (Table 3 and Figure 7) have

revealed that most sharks head west after congregating at this location, although the

information available is still incomplete (www.galapagoswhaleshark.com).

The big picture: formulating global hypotheses

After collating all available whale shark occurrences and tracking data, the question of

whether or not whale sharks are migrating across or within oceans remains unanswered.

One must conclude therefore that such conventional methods to study whale shark

movement are still of insufficient temporal and spatial coverage to provide useful

conclusions at the ocean-basin scale. We thus propose a different approach by shifting

research from a local to a global perspective.

The Indian Ocean is the region from which the most whale shark data have been

collected; therefore, it is the best region in which to begin formulating hypotheses

regarding whale shark movement patterns. We know that some locations are linked via

direct observation such as Mozambique and Seychelles (Rowat & Gore, 2007) or Ningaloo

and Christmas Island (Figure 7). It is clear that appearance timings occur sequentially

(Figure 6). Therefore, we hypothesise that at least some sharks move in a general

clockwise pattern from the southwest (KwaZulu-Natal) in January, to the southeast Indian

Ocean (Ningaloo) in March over approximately two years. According to Eckert and Stewart

(2001), the mean pace at which whale sharks travel varies from about 2 to 30 km day-1,

with more horizontal distance covered when closer to the surface. Average speeds of

around 30 km day-1 have also been reported by others (Brunnschweiler et al., 2009; Eckert

Page 69: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

71

et al., 2002; Hsu et al., 2007; Hueter et al., 2008; Rowat & Gore, 2007), while faster

speeds have been reported for short-term movements (Gunn et al., 1999). Assuming

constant travel speeds, a dominance of surface swimming and otherwise ideal conditions,

a shark moving at about 30 km day-1 would be able to travel > 10,000 km year-1 –

approximately the distance between South Africa and southern Western Australia. Using

this logic, two years to complete the longer clockwise migration from South Africa to

Ningaloo is biologically plausible.

Once in Ningaloo, even though the strength of the southward-flowing Leeuwin

Current along the western coast of Australia might influence the degree to which they

penetrate southward, previous tagging results demonstrate that the predominant

movements are northward towards Indonesia, and then either westward or eastward

(CSIRO, 2005) (Figure 7). Tagging data show that whale sharks take about five months to

return to the central Indian Ocean (passing near Christmas Island much earlier than the

seasonal peak in abundance) and about four months to travel east as far as East Timor.

Even though peaks of occurrence are not published for the southern areas of the Malay

Archipelago/Timor, whale sharks are found occasionally in the Coral Sea, about four to

five months after the Ningaloo peak (Figure 6). From the Coral Sea, environmental

conditions are likely influential on the patterns of subsequent movement. Either they (1)

move farther south towards the north coast of New Zealand (surpassing 35 ° S) where

their occurrence peaks in February (Duffy, 2002), (2) return to the Indian Ocean or (3)

migrate eastward (whale sharks are occasionally seen around the islands of the south

west and central Pacific Ocean (Compagno, 2001; and data collected by purse seine

fisheries) (Figure 5).

Page 70: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

72

In the central Indo-Pacific, whale sharks tracked from Taiwan travelled close to

the Philippines even outside the peak abundance season (March and June). Eckert et al.

(2002) tracked whale sharks from the Philippines in February, from where they travelled

west passing Malaysia and south Vietnam toward the Gulf of Thailand in about 4.5

months. There are many routes that potentially link the Pacific and Indian Oceans in the

Malay Archipelago (e.g., Andaman and Java Seas), such that whale sharks can cross to

the Indian Ocean towards Christmas Island or the north of Australia.

From a temporal perspective (Figure 6), there are no obvious paths from eastern

locations in the Indian Ocean to the west through to Bangladesh. That lead us to the

hypothesis that whale sharks could swim straight towards the Maldives (as was the case

for a shark tagged off Ningaloo moving far into the central Indian Ocean). This would also

be consistent with the lower numbers of whale sharks sighted on the eastern coast of India

compared to the western coast (Pravin, 2000). Additionally, whale sharks could also move

from the Maldives through to the Seychelles and then to South Africa based on observed

temporal patterns of occurrence (Figure 6).

For there to be population connectivity between the Indian/Pacific and Atlantic

Oceans, whale sharks would need to travel through the Cape of Good Hope in South

Africa. Beckley et al. (1997) suggested that whale shark strandings on the east coast of

South Africa might be associated with changes in temperature associated with Agulhas

Current from the Indian Ocean mixing with the cold water upwelled by the Benguela

Current in the southeast Atlantic. Beckley et al. (1997) also stated that stranded animals

were generally small, which suggests that only larger animals would survive the crossing

from the Indian to the Atlantic Ocean. The greater thermal inertia of larger sharks might

Page 71: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

73

allow them to transit in the southern Atlantic Ocean until they reach the warmer waters of,

for example, Gabon (peak of occurrences registered by purse seine fisheries in the

Atlantic - Figure 5), or the region around the Brazilian Saint Peter and Saint Paul

Archipelago where sharks are observed year-round, but mainly during the first semester

with a peak in June (Hazin et al., 2008) – six months after the peak in South Africa.

Within the entire Atlantic Ocean, there are tracks available mostly for the area of

the Gulf of Mexico/Caribbean Sea that show some linkages between the aggregations of

Central America (Figure 7). This suggests that at least here there is a single meta-

population. The odd long track starting from Mexico in August and finishing south of the

Saint Peter and Saint Paul Archipelago in January suggests that whale sharks from this

Central American population can travel great distances within the Atlantic Ocean. Year-

round sightings of whale sharks in the Archipelago outside of peak months (Hazin et al.,

2008) hint at the possibility of a trans-Atlantic thoroughfare.

Because the occurrence of whale sharks in the Azores islands in the eastern

Atlantic is seasonal (mostly at the end of August and September) and does not occur

every year (Machete & Afonso, pers. comm.), it is possible that whale sharks only travel

there when environmental conditions (especially water temperature) are most suitable (Sa,

2008). The Azores current, flowing south of the Azores Archipelago (where most whale

sharks are spotted by tuna fisheries), originates from the Gulf Stream (Klein & Siedler,

1989). When the Gulf Stream is strong, the warmer-than-usual waters near the Azores

might encourage more northerly forays (Sa, 2008). We therefore hypothesise that

individuals travelling to the Azores might originate from the Gulf of Mexico. Whale sharks

are also occasionally seen in the Madeira Islands (Portugal) to the south-east of the

Page 72: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

74

Azores (Wirtz et al., 2008). Because water temperatures in these parts of the Atlantic are

generally cooler, movements toward the warmer southerly waters closer to the Equator

near the Saint Peter and Saint Paul Archipelago and West Africa would be more typical –

a prediction supported by observation from fishing fleets in those areas.

Breeding within the Atlantic is supported by the discovery of a single neonate (58

cm total length) specimen inside the stomach of a captured blue shark (Kukuyev, 1996;

Martin, 2007). Other neonates have been caught near the Equator in this ocean (Wolfson,

1983) and an egg case was found in the Gulf of Mexico in the early 1950s (Baughman,

1955). Neonates have also been found in the eastern Pacific Ocean (Wolfson, 1983),

where large adult females are often observed (Eckert & Stewart, 2001). Although more is

known about whale sharks inhabiting the Indian Ocean, only 2 neonates (< 1 m) have ever

been reported there (Rowat et al. 2008). Despite young juveniles (~ 1 m total length)

occasionally observed swimming with larger individuals (e.g., Pillai, 1998), most of the

Indian Ocean aggregations are comprised mainly of immature males, which makes it

unlikely that the region includes a permanent breeding area. If there is a nursery in the

Indo-Pacific area (e.g., around The Philippines), and if whale sharks are able to cross to

the Indian Ocean, small juveniles could potentially occur in the latter even in the absence

of a nursery.

Notwithstanding these hypothesised linkages, we know that whale sharks often

return to the same aggregation at least semi-annually (e.g., Maldives, Ningaloo; Meekan

et al., 2006) based on photographic matching (Speed et al., 2007). Such repeat sightings

do not support the notion that all individuals regularly travel around ocean basins over two-

to four-year cycles. We hypothesise instead that even though whale sharks can travel over

Page 73: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

75

entire or even between ocean basins, many (perhaps most) remain near single

aggregation sites for several months or years. For example in the Maldives, whale sharks

might remain in the general area, travelling west (December to April) to east (May to

November) of the islands over one or more years (Figure 6), or extend the travelling

farther from Gujarat (northwest side of India) to west of the Maldives in the first half of the

year, and then east of the Maldives and Tamil-Nadu (east side of India) in the second.

From here they would have the option to move east (e.g., to Bangladesh in December) or

back to the west of the Maldives, south India and Gujarat (Figure 6). This back-and-forth

movement around India and the Maldives accords with the higher re-sighting rate in the

Maldives relative to other nearby locations (Riley et al., 2010). Other examples of these

hypothesised shorter migration routes and possible populations are depicted in Figure 8.

Page 74: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

76

Fi

gure

8: W

hale

shar

k mig

ratio

n pa

ttern

s. De

pict

ing

likely

sub-

popu

latio

ns (r

ed ci

rcles

). Bl

ue ar

rows

indi

cate

pos

sible

links

bet

ween

sub-

popu

latio

ns (a

ggre

gatio

ns).

Blac

k arro

ws d

enot

e mov

emen

ts re

veale

d fro

m ta

ggin

g da

ta, a

nd d

ashe

d lin

es in

dica

te n

o cu

rrent

evid

ence

for a

mig

rato

ry p

athw

ay (e

xcep

t the

que

stio

nabl

e 13,0

00-k

m tr

ack a

cros

s the

Pac

ific f

rom

Eck

ert a

nd S

tewa

rt (2

001)

show

n as

a ba

ck d

otte

d lin

e)

Page 75: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

77

Foreseeing change

The global distribution of whale sharks lies predominantly within tropical and warm

temperate waters between ~ 30 º north and south (Compagno, 2001; Last & Stevens,

2009). Since the late 1980s, however, individual sharks have been observed well pole-

ward of these latitudes. Indeed, Wolfson (1986) recorded whale shark at latitudes of 41 º N

and 36.5 º S, and more recently, Duffy (2002) reported them in New Zealand south of 35 °

S. One whale shark has even been sighted as far north as 44 º in the Bay of Fundy,

Canada (Turnbull & Randell, 2006). Whale sharks are also occasionally sighted south of

the Azores (~ 41 ° N), mainly by tuna fishermen (Machete & Afonso, pers. comm.) and

recently, many sightings (> 400) have been made mainly around the island of Santa Maria

from June to September (fisheries data), likely in response to an unusual pulse of warm

water (Sa, 2008).

Whale sharks mainly stay within a narrow range of sea surface temperatures. In

the Sea of Cortés, most tracked individuals were in waters between 28 and 32 º C (Eckert

& Stewart, 2001); in the Seychelles, most were between 25 and 35 ºC (Rowat & Gore,

2007); in the north-western Pacific, most were between 23 and 32 º C (Hsu et al., 2007);

for pelagic sightings derived from fisheries records, > 90 % of 1185 records were in

surface waters between 26.5 and 30 o C (Sequeira et al., 2012). If water temperature plays

an important role in modifying whale shark distribution, the rapid warming of the world’s

oceans arising from anthropogenic climate change (Hegerl & Bindoff, 2005) will likely

affect their future distribution. Because whale sharks seem to avoid higher temperatures

(e.g., around the Equator; Sequeira et al., 2012), poleward shifts are more probable than

an overall expansion of their current distribution. In addition to these expected

Page 76: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

78

distributional changes, shifts in both abundance and distribution are expected from

previous (Bradshaw et al., 2007; Romanov, 2002) and ongoing (Riley et al., 2009; White &

Cavanagh, 2007) commercial fishing, potentially from excessive disturbance arising from

ecotourism at aggregation sites (Cárdenas-Torres et al., 2007; Pierce et al., 2010; Quiros,

2007; Rowat & Engelhardt, 2007) and other human-related disturbances (Haetrakul et al.,

2007; World Widlife Foundation, 2012).

Conclusions

Despite ample evidence that whale sharks are capable of long-distance travel, confirmed

by genetic evidence that the world’s populations are connected at least at the generational

scale, we know that many individuals spend long periods within the immediate surrounds

of particular locations. We have thus provided a conceptual model based on known

movement patterns, aggregation timings and oceanic structure demonstrating how within-

and inter-ocean movements might occur, even if they do so rarely and without non-genetic

evidence yet to test our hypotheses. This lack of evidence however does not counter our

model. It mostly derives from the thousands (i.e., not millions or tens of millions) of photos

used in photo identification-based studies. Photo-matching is only semi-automated (e.g.,

I3S; Van Tienhoven et al., 2007), thus limiting the probability of photographing ‘migrating’

individuals, and making the probability of resighting these individuals in other aggregations

even smaller. Moreover, flank mismatch (right or left side only) can also limit the number of

individuals that can be compared.

Viewing whale shark ecology from our global perspective therefore engenders

many related hypotheses that can be tested as technology improves and genetic evidence

Page 77: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

79

is refined, and it isolates the most important questions to determine the long-term

conservation status of this species. A natural progression of our global movement model is

the construction of a broad-scale population model that connects sub-populations based

on inferred demographic and movement rate data. Application and continual refinement of

biologically plausible rates will further hone the global model and its predictions, thereby

unfolding key aspects of the species cryptic life history so that research can target the

most relevant aspects. It is clear that more data are required from locations close to key

aggregation sites where hypothesised connections can be validated. Tagging and

photographic-identification data within these regions should be prioritized based on their

putative connections predicted under the global model. For example, areas such as

Madagascar, the Seychelles and Tanzania, or the Philippines, Malaysia and Gulf of

Thailand require much broader research coverage to validate existing hypotheses of

connectivity.

Tagging studies should also target many sharks within the same aggregation to

chronicle the highest number of potential movement patterns (cf. Eckert & Stewart, 2001;

Wilson et al., 2006), and tags should be deployed as close to the end of the peak

aggregation season as possible to maximize the potential for measuring long-distance

trajectories, especially given the short monitoring periods characteristic of such studies

(Hammerschlag et al., 2011). Other improvements to maximize monitoring time include the

application of anti-fouling paint to tracking equipment (Hammerschlag et al. 2011), using

fast-loc GPS technology (Hays et al. 2007) and switching from steel to copper salt-water

switches used to improve the efficiency of duty cycling (Hays et al. 2007). Improvements in

Page 78: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

80

tag attachment methods are also needed to prevent premature detachment (e.g.,

Brunnschweiler et al., 2009).

Another important consideration when describing migratory patterns is that longer

migrations might reflect sex- and age-specific behaviour. Indeed, 3 of the 4 whale shark

tracks exceeding 3000 km were large females (~ 7 m) (Tables 1 - 3). As suggested

previously, female natal philopatry might result in long migrations only for the purposes of

occasional breeding (Ramírez-Macías et al., 2007). This sort of sex-specific behaviour

would also explain the high resighting rates in some aggregations dominated by immature

males. While a similar number of males and females tracks have been published to date,

different size classes (≤ 6 m or > 6 m) have different average track lengths (200-300 and

1200 km, respectively). Of course, there are exceptions to this trend; some juveniles

tagged in Taiwan had tracks > 3000 km even though they remained resident (Hsu et al.,

2007). More tracking studies of longer duration will assist in describing these stage-

specific movement trends and capacities.

Given the large distances between some of the known aggregations, another way

to refine our knowledge of this species is to profit from the opportunities provided by

certain pelagic fisheries. Expansive commercial fisheries (e.g., tuna purse seiners) do

release whale sharks that are accidentally captured or encircled by their nets. Tagging

endeavours associated with these releases could potentially provide the data necessary to

estimate vital rates and a better description of pelagic movement patterns. Further,

focussing tag deployments and tissue collections near potential thoroughfares, such as

those hypothesised under our model like the Saint Peter and Saint Paul Archipelago and

Page 79: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

81

Christmas Island, would likely increase the probability of capturing pan- or trans-oceanic

movements.

Page 80: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

82

CHAPTER III. Ocean-scale prediction of whale shark distribution

A. Sequeiraa, C. Mellina,b, D. Rowatc, M. Meekand, C. Bradshawa,e

aThe Environment Institute and School of Earth and Environmental Sciences, The University of Adelaide, South Australia 5005, Australia bAustralian Institute of Marine Science, PMB No. 3, Townsville MC, Townsville, Queensland 4810, Australia cMarine Conservation Society, Seychelles, P.O. Box 1299, Victoria, Mahe, Seychelles

dAustralian Institute of Marine Science, UWA Oceans Institute (MO96), 35 Stirling Hwy, Crawley, Western Australia 6009, Australia. eSouth Australian Research and Development Institute P.O. Box 120, Henley Beach, South Australia 5022, Australia

Diversity and Distributions 2012; 18 : 504 – 519

Page 81: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

83

Statement of authorship

OCEAN-SCALE PREDICTION OF WHALE SHARK DISTRIBUTION

Diversity and Distributions 2012; 18 : 504 – 519

SEQUEIRA, A.M.M. (Candidate) Planned the article, performed the analysis, interpreted data, developed the models, wrote manuscript and acted as corresponding author. I hereby certify that the statement of contribution is accurate.

Signed…………………………………………………………………….Date………………

MELLIN, C. Helped planning the article, helped with data interpretation, analysis and development of the models, and assisted the writing. I hereby certify that the statement of contribution is accurate and I give permission for the inclusion of the manuscript in the thesis.

Signed…………………………………………………………………….Date………………

ROWAT, D. Provided data and assisted interpretation. I hereby certify that the statement of contribution is accurate and I give permission for the inclusion of the manuscript in the thesis.

Signed…………………………………………………………………….Date………………

Page 82: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

84

MEEKAN, M. G. Provided critical evaluation and interpretation. I hereby certify that the statement of contribution is accurate and I give permission for the inclusion of the manuscript in the thesis.

Signed……………………………………………………………………….Date……………

BRADSHAW, C.J.A. Supervised development of work, helped in data interpretation, models development and manuscript evaluation and writing. I hereby certify that the statement of contribution is accurate and I give permission for the inclusion of the manuscript in the thesis.

Signed……………………………………………………………………….Date……………

Page 83: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

85

Ocean-scale prediction of whale shark distribution

Abstract

Aim Predicting distribution patterns of whale sharks (Rhincodon typus, Smith 1828) in

the open ocean remains elusive due to few pelagic records. We developed

multivariate distribution models of seasonally variant whale shark distributions derived

from tuna purse-seine fishery data. We tested the hypotheses that whale sharks use a

narrow temperature range, are more abundant in productive waters and select sites

closer to continents than the open ocean.

Location Indian Ocean

Methods We compared a 17-year time series of observations of whale sharks

associated with tuna purse-seine sets with chlorophyll-a concentration and sea surface

temperature data extracted from satellite images. Different sets of pseudo-absences

based on random distributions, distance to shark locations and tuna catch were

generated to account for spatio-temporal variation in sampling effort and probability of

detection. We applied generalised linear, spatial mixed-effect and Maximum Entropy

models to predict seasonal variation in habitat suitability and produced maps of

distribution.

Results The saturated generalised linear models including bathymetric slope, depth,

distance to shore, the quadratic of mean sea surface temperature, sea surface

temperature variance and chlorophyll-a had the highest relative statistical support, with

the highest percent deviance explained when using random pseudo-absences with

fixed effect-only models and the tuna pseudo-absences with mixed-effects models

(e.g., 58% and 26% in autumn, respectively). Maximum Entropy results suggested that

Page 84: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

86

whale sharks responded mainly to variation in depth, chlorophyll-a and temperature in

all seasons. Bathymetric slope had only a minor influence on presence.

Main conclusions Whale shark habitat suitability in the Indian Ocean is mainly

correlated with spatial variation in sea surface temperature. The relative influence of

this predictor provides a basis for predicting habitat suitability in the open ocean,

possibly giving insights into the migratory behaviour of the world’s largest fish. Our

results also provide a baseline for temperature-dependent predictions of distributional

changes in the future.

KEYWORDS colour imagery; Indian Ocean; Rhincodon typus; satellite data; sea

surface temperature; species distribution models; tuna purse-seine fisheries

Introduction

Species distribution modelling has been applied widely in conservation ecology

(Guisan & Thuiller, 2005; Guisan & Zimmermann, 2000; Hirzel et al., 2002; Phillips et

al., 2009) to understand the role of environmental conditions driving distribution and

abundance patterns. Predictions arising from this approach are essential to determine

the likely effects of habitat change on persistence probability or community structure

(Araújo & Williams, 2000; Beaumont et al., 2007). Such models incorporate

information on environmental conditions and combine these with the known

distribution of a species or population to define its ecological niche (Hutchison, 1957;

Kearney & Porter, 2009) and predict its probability of occurrence in a location where

no biological information is currently available (Robertson et al., 2003).

Page 85: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

87

The use of these models in terrestrial systems is commonplace (Lehmann et

al., 2002); however, their application in non-terrestrial systems is rare with only a few

marine examples such as seagrass beds (Kelly et al., 2001), bryophytes (Sérgio et al.,

2007), coral communities (Beger & Possingham, 2008; Garza-Pérez et al., 2004;

Mellin et al., 2010b; Tittensor et al., 2009), shellfish in coastal ecosystems (Sequeira et

al., 2008) and dolphins / whales in the Mediterranean Sea (Praca & Gannier, 2007)

and in the eastern tropical Pacific Ocean (Oviedo & Solís, 2008). To our knowledge,

they have not been applied to the distribution of widely dispersed pelagic fishes in

open oceans. Data collection in marine environments is often “very difficult, resource-

intensive and expensive” (Richardson & Poloczanska, 2008) and the logistics of this

task is much greater in the pelagic realm where work is typically based aboard large

vessels. There is limited opportunity for synoptic sampling because few research

programs can afford multiple ships operating simultaneously, and datasets of

distribution of fishes typically contain only reports of presence; therefore, confirming

true absences can be difficult (Zaniewski et al., 2002).

Although some statistical techniques can cope with presence-only data – e.g.,

ENFA (Hirzel et al., 2002), BIOCLIM (Thuiller et al., 2009), MaxEnt (Phillips & Dudík,

2008) – and produce acceptable results (Elith et al., 2006; Zaniewski et al., 2002),

regression methods such as boosted regression trees (De’ath, 2007; Elith et al.,

2008), multivariate adaptive regression splines (Leathwick et al., 2005), or generalised

linear or additive models (Hastie & Tibshirani, 1986; Nelder & Wedderburn, 1972) tend

to provide more realistic projections when using reliable and accurate absence data

(Brotons et al., 2004; Zaniewski et al., 2002). Where such data are unavailable to use

in regression models, an alternative approach is to generate pseudo-absences that

Page 86: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

88

should, ideally, also account for any spatial bias in the sampling effort (Phillips et al.,

2009).

The whale shark (Rhincodon typus, Smith 1828) is the largest of all fish and

can reach more than 12 m in total length (Stevens, 2007). Although little is known

about its habitat selection or migration patterns, the whale shark appears to be a

highly mobile species. It is distributed in oceanic and coastal tropical waters

(Compagno, 2001; Last & Stevens, 1994) from 30 ºN to 35 ºS, spending most of its

time (often > 80%) at the surface (waters < 20 m deep), with regular excursions to

depths of 300-500 m during the day when transiting the open ocean (Brunnschweiler

et al., 2009; Wilson et al., 2006). Whale sharks are known to aggregate nearshore in a

number of coastal locations (e.g. see Rowat, 2007). This behaviour makes them the

subject of highly lucrative ecotourism industries around the world (Martin, 2007; Rowat

& Engelhardt, 2007), although in the past such aggregations have also been targeted

by commercial fishing (Joung et al., 1996; Pravin, 2000; Stevens, 2007; White &

Cavanagh, 2007). While large-scale fisheries have closed, artisanal and small-scale

fisheries for the species still exist in many parts of the tropics (Riley et al., 2009;

Stewart & Wilson, 2005; White & Cavanagh, 2007). As the whale shark is considered

a Vulnerable species (IUCN, 2010), Bradshaw et al. (2008) stressed the importance of

understanding their migratory behaviour because they can travel from regions where

they are protected to those where they are still under threat.

At present there is little information on the habitat requirements and pelagic

distribution of any filter-feeding shark (Sims et al., 2003; Southall et al., 2006),

although movements of whale sharks have been associated with climate and surface

water conditions (Cárdenas-Palomo et al., 2010; Sleeman et al., 2010a), plankton

blooms (Colman, 1997; Kumari & Raman, 2010) and other potential food sources

Page 87: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

89

(Meekan et al., 2009; Wilson, 2002). These fine-scale studies (100s m – 100s km) and

other hypotheses have never been formally tested at the ocean-basin scale (1000s

km) (Cárdenas-Torres et al., 2007; Rowat, 2007; Southall et al., 2006).

Access to a 17-year dataset of whale shark sightings generated by the tuna

fishing industry in the Indian Ocean offered the opportunity to examine distribution

patterns of this species in the open ocean. By combining these data with

environmental time series measured via remote sensing, we can potentially infer

seasonal patterns in habitat suitability for whale sharks at spatial scales likely to

encompass the range of entire populations. Our main objectives were to: (i) determine

the spatio-temporal patterns in habitat suitability for whale sharks over a broad (ocean

basin) spatial scale and (ii) identify the most important determinants of whale shark

distribution; we tested the hypotheses that specific temperature ranges, productivity

values and distance from shore functions drive seasonal spatial distribution of whale

sharks.

Methods

Dataset collected by tuna fisheries (whale shark presences)

Whale sharks, mainly associated with net-sets for tuna catch (hereafter referred to as

whale shark sightings), were recorded by purse-seine fishers registered at the Indian

Ocean Tuna Commission (IOTC; Pianet et al., 2009) and these data were made

available by vessel owners through the Institut de Recherche pour le Développement

(IRD; France). The dataset consisted of a 17-year (1991-2007) time series of (i) whale

shark sightings, date and location (longitude and latitude at 0.01° precision) for a total

Page 88: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

90

of 1185 whale sharks; and (ii) tuna catch (tonnes) and fishing effort (days) pooled for

each 5° grid cell, per month and per year (6618 records total).

The dataset covered 172,800 km2 of the Indian Ocean between 30 °N to 30

°S and 35 to 100 °E (Figure 9a). When all years were combined, a total of 811 whale

sharks were recorded in autumn (Apr- Jun; 68 % of records), 68 in winter (Jul- Sep; 6

%), 191 in spring (Oct-Dec; 16 %) and 115 in summer (Jan-Mar; 10 %) – austral

seasons. We divided the number of sightings by the fishing effort of associated

catches to calculate a sightings-per-unit-effort index (SPUE - with unit effort being

fishing days), for each month across the study area (Figure 9a). We applied a Kruskal-

Wallis analysis to the whale shark locations per season to test whether the

distributions of whale shark SPUE varied among seasons.

We also obtained a second dataset of tuna catches in the Indian Ocean derived

from the same raw data as the whale shark-associated sets, covering the same spatial

extent but aggregated at a finer scale (1° resolution) (Pianet et al., 2009). This dataset

consisted of: total tuna catch (tonnes) and fishing effort (days) pooled for each 1° grid

cell per month (Figure 9b). As for the 5° resolution dataset, we estimated the

normalized catch-per-unit-effort (CPUE) by dividing the total tuna catch by the fishing

effort in each grid cell for each month across the study area.

Page 89: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

91

To reduce bias associated with temporal autocorrelation and test the hypothesis

that specific environmental parameters (e.g., SST, or Chl a – used here as a proxy for

food availability) affect whale shark distributions in a specific season, we split the data

into four groups each of them corresponding to quarters of the year and defined here

a) b)

Figure 9: Indian Ocean Tuna Commission (IOTC) datasets collected from 1991 to 2007 on tuna purse-seine fisheries in the Indian Ocean. (a) Tuna-associated fisheries per 5-° grid cell including whale shark (Rhincodon typus, Smith 1828) associated net-sets (hereafter referred to as sightings): top panel – area covered by the fisheries (grey) and 1185 whale shark sightings (blue dots); middle panel – associated effort in days spent fishing within the area sampled; bottom panel – whale shark sightings per unit effort (SPUE) of associated fisheries (in sightings per fishing day). Grey areas represent sampled area where no whale sharks were sighted. (b) Tuna total fisheries per 1-° grid cell: top panel – total tuna catch; middle panel – total fishing effort in days spent; bottom panel – total tuna catch per unit effort (tons per fishing day).

20˚

-20˚

Page 90: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

92

as summer (January to March), autumn (April to June), winter (July to September) and

spring (October to December).

Pseudo-absence dataset generation

Whale shark dataset were presence-only, as true absences were unknown. We

therefore generated pseudo-absences for use in generalised linear models (GLM) and

spatial generalised linear mixed-effect models (GLMM). The pseudo-absence

locations were also included as background data in additional MaxEnt runs, to get

more comparable results between the two modelling procedures.

To assess the influence of the method used to generate pseudo-absences on

model outputs, we generated the same number of pseudo-absences as presence

records for each season following three different methods: (i) random selection of non-

presence grid cells within the 9-km grid over the entire area accounting for uneven

sampling bias by weighting pseudo-absences by the fishing effort in the same area

(random) , (ii) selection of non-presence grid cells with a probability that was weighted

by the inverse distance to the whale shark presences (IDW), assuming that detection

probability was higher near recorded presences and to account for the same spatial

bias, and (iii) selection weighted by the total tuna catch from the 1° dataset (tuna),

assuming that whale sharks and tuna would have similar distribution patterns due to

their association – it is common practice among tuna fishers to use whale sharks as

indicator species when targeting tuna (Matsunaga et al., 2003), making whale shark

absences more likely to occur in areas of high tuna catch (and high fishing effort), but

where no whale shark sightings were recorded.

We generated each set of pseudo-absences 100 times using the srswor

function (simple random sampling without replacement) from the Sampling package in

R (R Development Core Team, 2010) prior to their use in GLM and GLMM.

Page 91: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

93

Environmental variables

We collated a environmental dataset at 9-km resolution over the study area that

included bathymetry (including mean depth in m, and slope in º), distance to shore

(km), distance to shelf (km), seasonal mean and standard deviation of chlorophyll-a

concentration (Chl a in mg m-3) and sea surface temperature (SST in °C). We initially

collated bathymetry data (see Figure B1 in Appendix B) across the study area at

approximately 1.7-km resolution using the one-minute grid for the General Bathymetry

Chart of the Oceans (GEBCO, 2003). Mean depth and slope were then calculated for

each 9-km grid cell. For each whale shark sighting, we calculated the shortest

distances to the coast and to the continental shelf in ArcGIS 9.2 with the Near tool

using a World Equidistant Cylindrical coordinate system.

We obtained Chl a and SST data at a 9-km resolution from relayed image

composites from the Sea-viewing Wide Field-of-view Sensor (SeaWiFS) and Moderate

Resolution Imaging Spectroradiometer (MODIS-Aqua), respectively (see Figure B2

and Figure B3). We averaged weekly daytime measures available since September

1997 for SeaWiFS and July 2002 for MODIS-Aqua for each month. For each season,

we calculated the resulting Chl a and SST mean and standard deviation in R (R

Development Core Team, 2010). Due to non-Gaussian data, we investigated

monotonic relationships among environmental predictors using Spearman’s rank

correlation coefficient (��.

Models

Generalised linear and mixed effects models

We set GLM with a binomial error distribution and a logit link function to compare, for

each season, the predictive ability of all possible combinations of environmental

Page 92: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

94

predictors. We included a quadratic term for SST to account for the possibility of a

selected temperature range given that they are poikilotherms and external

temperatures likely affect their metabolic processes (Bullock, 1955).We compared

models based on two bias-corrected indices of parsimony (Burnham & Anderson,

2004): the information-theoretic Akaike’s information criterion corrected for small

sample sizes (AICc) and the Bayesian information criterion (BIC). AICc favours more

complex models (with higher predictive capacity) when tapering effects exist and

sample sizes are large, whereas BIC tends to identify the main drivers of complex

relationships (Link & Barker, 2006). We assessed each model’s strength of evidence

relative to the entire model set by calculating relative model weights (wAICc and

wBIC). We used the percentage of deviance explained (De) to quantify each model’s

explanatory power.

We also applied a 10-fold cross-validation using 1000 iterations to assess the

mean prediction error of the model that maximized De for 10 % of the observations

that were randomly selected and left out of the training dataset. We assessed the

predictive power of the models according to Cohen’s Kappa statistics κ that measures

agreement/accuracy (Cohen, 1960) and varies from ≤ 0 for no agreement between

observed data and hypothetical probability, to 1 when perfect agreement occurs.

Following Woodby et al. (2009), we considered that a model had a poor performance

when κ < 0.4, good when 0.4 < κ < 0.75, and excellent when κ > 0.75; these ranges

were based on the κ agreement scale proposed by Landis & Koch (1977). For each

index of model performance, we calculated the median across the results obtained for

each of the 100 replicates of each pseudo-absence dataset.

We assessed potential spatial autocorrelation in both observations and GLM

residuals as a function of distance between sites based on Moran’s I (Diggle & Ribeiro,

Page 93: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

95

2007) after a Bonferroni correction (Legendre & Legendre, 1998). We used the spatial

correlation structure that gave the best fit to the null model to define the error

covariance matrix in a spatial GLMM, coding the 1° grid cell as a random effect within

each GLM fitted previously (e.g. Mellin et al., 2010b). We fit GLMM using the

penalized quasi-likelihood (Venables & Ripley, 2002a) and derived predictions for the

entire area sampled in each season. We then built prediction maps in ArcGIS 9.2.

Following Araújo & New (2006), we used an ensemble approach to combine the

full range of results obtained by the different techniques used to account for pseudo-

absences. We weighted the contribution of each model according to its percentage of

deviance explained to build an ensemble prediction for seasonal distribution of whale

sharks in the Indian Ocean.

MaxEnt (Maximum Entropy)

We compared GLM results to predictions obtained using a presence-only modelling

technique, Maximum Entropy(Phillips et al., 2006; Phillips & Dudík, 2008; Phillips et

al., 2009; Phillips et al., 2004), using the software provided by the authors (MaxEnt

version 3.3.3e November 2010, AT&T Labs Research). MaxEnt is a tool for generating

species distribution models from presences-only data. This modelling tool uses

covariate data from species presence locations and background sampling to estimate

habitat suitability for the species occurrence (for a detailed statistical explanation of

MaxEnt, see Elith et al., 2011). To make MaxEnt models more comparable to the

GLM/ GLMM, we used the same datasets for presences and pseudo-absences to

generate similar models per season with restricted settings (i.e., we fitted only linear

and quadratic features) in MaxEnt (for details on MaxEnt features options see Elith et

al., 2011). Due to the functionality of MaxEnt to use presences-only data, we ran

additional models making use of the entire background where environmental data

Page 94: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

96

were available and allowing the model to use all the features available in the console:

auto-features (linear, quadratic, product, threshold and hinge).

We projected results to the entire area sampled within each season, using the

area under the curve (AUC) to measure model performance. For models using the

same presence and pseudo-absences datasets, we calculated the Kappa statistic to

compare results directly to the GLM performance. We selected the jack-knife test

option in all model sets to infer the relative importance of each variable.

Results

Due to the uneven fishing effort in each 5° cell within the area covered by the tuna

purse-seine fisheries (e.g., ≤100 days in the eastern side and ≥ 11,000 days in the

western side of the Indian Ocean; Figure 9a), the aggregated normalized whale shark

data was used instead of the whale shark sightings alone. SPUE varied from 2.10 ×10-

3 to a maximum of 444 × 10-3 in each grid cell during the season when most whale

sharks were sighted (autumn), followed by SPUE between 2.45 × 10-3 and 200 × 10-3

in summer, 0.86 × 10-3 to 160 × 10-3 in winter and 0. 46 × 10-3 to 37.9 × 10-3 in spring.

Sighting (SPUE) patterns of whale sharks shifted between seasons (Figure 10);

there was a non-random (Kruskal-Wallis H3 = 33.702 ; P < 0.001) clockwise shift in

relative occurrence from the Mozambique Channel in autumn, through the areas

around the Equator in the western Indian Ocean in winter, spreading east in spring

and returning to the Mozambique Channel in summer. We found a strong positive

correlation (Spearman’s rho = 0.89) between whale shark SPUE and tuna CPUE

recorded by the same tuna-purse fleets. We used the tuna catch in the 1° resolution

Page 95: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

97

dataset to weight the probability of selecting a pseudo-absence in the following

analyses, as these data were available at a much finer spatial resolution.

Seasonal standard deviation of Chl a and its mean value were highly correlated

in all seasons (-0.92 < Spearman’s ρ < 0.87; P < 0.001), as well as the distance to the

continental shelf and to the shore (ρ = 0.9115; P < 0.001), and depth and distance to

the shelf (ρ = 0.6146; P < 0.001). Even though distance to shelf is potentially more

informative regarding whale shark distribution than distance to shore, we omitted the

standard deviation of Chl a and distance to the shelf (instead of standard deviation of

Chl a and both distance to shore and depth) from the list of candidate predictors to

construct the model set with the lowest number of uncorrelated variables.

Figure 10: Seasonal variation in aggregated whale shark (Rhincodon typus, Smith 1828) sightings per unit effort (SPUE, where effort corresponds to the number of fishing days) in the area of the Indian Ocean sampled by the Indian Ocean Tuna Commission (IOTC) from 1991 to 2007. (a) autumn (Apr-Jun) : 811 whale sharks recorded (68% of sightings); (b) winter (Jul-Sep): 68 whale sharks sighted (6% of sightings); (c) spring (Oct-Dec): 191 whale sharks sighted (16 % of sightings); (d) summer (Jan-Mar): 115 whale sharks sighted (10% of total sightings).

Page 96: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

98

Generalised linear and mixed-effects models

The percentage of deviance explained was highest for the generalised linear model

including all environmental variables in all seasons, irrespective of the technique used

for generating pseudo-absences (Table 4). The highest values were obtained with

random pseudo-absence (57% in autumn and 20% in summer) and lowest with IDW

pseudo-absences (< 15% in all seasons).

Statistical support (wAICc) was greatest for the model including all

environmental variables as well, but only when using random and the tuna pseudo-

absences in all seasons, except in spring when removing Chl a increased support

using the tuna dataset (Table 4). The top-ranked model according to wBIC only

matched the one that also maximized wAICc in autumn for the three pseudo-absence

datasets (Figure B4:), and in winter and summer when using random pseudo-

absences (Table 4). Both observations and GLM residuals were spatially

autocorrelated (P < 0.001). The spatial correlation structure that gave the best fit to the

null model varied between the exponential and spherical, with a shape parameter of 5

or 10, depending on the method used for generating pseudo-absences (an example of

the resulting Moran’s I plots is shown in Figure B5).

The pseudo-absence dataset that provided the best results also differed

between modelling techniques; for example in autumn, the highest deviance was

explained with random when using GLM, and with tuna when using GLMM. Percent

deviance explained was generally higher for all GLM in all seasons. After accounting

for the spatial autocorrelation (Table 5), GLMM using tuna pseudo-absences explained

the highest deviance (25.5% in autumn, 23.7% in summer, 11.1% in spring and 5.3%

in winter).

Page 97: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

99

Tabl

e 4: S

umm

ary o

f gen

erali

sed

linea

r mod

els re

latin

g pr

obab

ility o

f wha

le sh

ark (

Rhin

codo

n ty

pus,

Smith

1828

) occ

urre

nce t

o oc

ean

prop

ertie

s. -

slope

, dep

th, d

istan

ce to

shor

e (sh

ore)

, mea

n se

a sur

face

tem

pera

ture

(SST

mea

n) an

d its

qua

drat

ic te

rm (S

ST m

ean2

), se

a sur

face

tem

pera

ture

stan

dard

dev

iatio

n (S

ST S

D)

and

chlo

roph

yll a

(Chl

a). S

hown

for e

ach

mod

el ar

e the

bias

ed-c

orre

cted

mod

el ev

iden

ce b

ased

on

weig

hts o

f Aka

ike’s

info

rmat

ion

crite

rion

corre

cted

for s

mall

sam

ple s

izes

(wAI

Cc),

weig

hts o

f Bay

esian

info

rmat

ion

crite

rion

(wBI

C) an

d th

e per

cent

age o

f dev

iance

expl

ained

(%DE

). Th

ree d

iffer

ent m

etho

ds w

ere u

sed

for g

ener

atin

g ps

eudo

-ab

senc

es: r

ando

m, in

vers

ely d

istan

t to

whale

shar

k sig

htin

g lo

catio

ns (I

DW) a

nd b

ased

on

tota

l tun

a cat

ch (t

una)

. Res

ults

onl

y for

case

s whe

re w

AICc

> 0.

001,

and

valu

es ≤

0.1

are s

hown

in lig

ht g

rey.

Note

: slo

pe, d

epth

, dist

ance

to sh

ore r

efer

red

to to

geth

er as

phy

sical

varia

bles

(phy

s); m

ean

sea s

urfa

ce te

mpe

ratu

re (S

ST m

ean)

, its q

uadr

atic

term

(SST

mea

n2) a

nd se

a sur

face

tem

pera

ture

stan

dard

dev

iatio

n (S

ST S

D) re

ferre

d to

toge

ther

as S

ST va

riabl

es (S

ST va

r).

Se

ason

au

tum

n

wi

nter

sp

ring

sum

mer

Mod

el

wAIC

c wB

IC

%DE

wA

ICc

wBIC

%

DE

wAIC

c wB

IC

%DE

wA

ICc

wBIC

%

DE

rand

om

phys

+ S

ST va

r -

- -

0.25

0.65

34.04

0.0

1 0.0

4 18

.00

- -

- sa

turate

d 1.0

0 1.0

0 57

.05

0.99

0.99

34.97

0.7

5 0.3

5 34

.72

0.99

0.96

19.88

ID

W

slo

pe

- -

- 0.0

5 0.0

7 0.1

5 -

- -

0.01

0.04

0.20

de

pth

- -

- 0.0

8 0.0

9 0.5

5 -

- -

0.04

0.11

1.08

sh

ore

- -

- 0.0

5 0.0

6 0.1

1 -

- -

0.03

0.07

0.70

SS

T va

r -

- -

0.21

0.02

4.51

0.20

0.89

11.80

0.2

5 0.0

3 3.9

8

Chl a

-

- -

0.07

0.07

0.32

- -

- 0.0

4 0.0

9 0.9

7

phys

+ S

ST va

riable

s -

- -

0.07

0.00

7.47

0.51

0.01

13.93

0.1

4 0.0

7 6.3

4

satur

ated

0.99

0.99

11.88

0.0

9

8.91

0.24

0.00

14.16

0.1

4 0.0

0 7.1

4

null

- -

- 0.1

3 0.6

0 0.0

0 -

- -

0.02

0.34

0.00

tuna

SST

varia

bles

- -

- 0.0

6 0.9

0 17

.92

0.00

0.37

22.64

-

- -

ph

ys +

SST

varia

bles

- -

- 0.2

6 0.0

0 20

.60

0.66

0.55

27.37

0.4

5 0.3

2 18

.19

sa

turate

d 1.0

0 1.0

0 42

.61

0.86

0.08

26.72

0.3

2 0.0

5 27

.77

0.54

0.10

19.10

Page 98: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

100

For each season predictive maps derived from different pseudo-absence datasets

resulted in similar patterns (Figure 11), while greater differences occurred among

seasons (Figure 12). During autumn, highly suitable areas were concentrated near the

Mozambique Channel and close to shore in the south-eastern side of the African

continent. A shift in habitat suitability occurred in winter towards the north and central

western Indian Ocean, spreading towards the east in a ‘C’ shape (surrounding the

Equator) in spring and stretching from east to west south of the Equator in summer.

GLM / GLMM MaxEnt

Figure 11: Habitat suitability of whale sharks (Rhincodon typus, Smith 1828) in the Indian Ocean during autumn. The prediction maps in the left panel are derived from generalised linear mixed-effects models, and those in the right panel are derived from MaxEnt when using three different techniques for generating pseudo-absences: i) randomly (random), ii) based on probability weighted by the inverse distance to whale shark presence locations (IDW to shark) and by iii) a probability directly weighted by total tuna catch (Tuna catch), respectively per row.

Page 99: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

101

Maximum Entropy

The pseudo-absences datasets derived from the three different techniques resulted in

similar maps in MaxEnt in each season (Figure 11). MaxEnt prediction maps were

generally consistent with that predicted by the GLM and GLMM (Figure 11), except for

spring - where MaxEnt presented the central equatorial area of the Indian Ocean as

the most suitable habitat. Within the total area sampled, the region likely to be more

suitable for whale shark occurrence in autumn was the Mozambique Channel, followed

by the western equatorial Indian Ocean in winter, the central area of the Indian Ocean

in spring, and more dispersed, but already including the Mozambique Channel again,

in summer. The MaxEnt variable importance ranking differed both among seasons and

the pseudo-absences dataset used in the model, while the percent of variable

contribution varied mainly within seasons (Table 6 and Supplementary information and

Figure B6 in Appendix B). In autumn the more important variable was the quadratic

term of sea surface temperature for all techniques, while the variable with the highest

percentage of contribution for the model results was distance to shore. For both winter

and spring, the highest ranked variables varied mainly between Chl a (mean), SST

(mean), SST (SD) and the quadratic SST term, while for summer depth was an

important variable with the highest percentage contribution.

The jack-knife test identified Chl a (in winter and spring) and physical

variables (in both autumn and summer) as those with important individual effects in all

model sets (Table 6). The most important single variables were mainly the producing

the poorest model results when excluded from the set of predictors.

AUC obtained with the MaxEnt models was generally low, varying from 0.574

to 0.721 (Table 6). In all seasons, the highest AUC values were obtained for models

with the random, followed by tuna pseudo-absences datasets (around 0.7), while the

Page 100: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

102

lower scores were obtained when the IDW dataset was used (below 0.63). k obtained

for MaxEnt were nearly always less than those obtained from GLM, and generally

higher than those from the GLMM accounting for spatial autocorrelation (Table 5).

MaxEnt models using the full background data available but with restricted

settings, i.e., keeping the features restricted to linear and quadratic, produced similar

results to the models using random and tuna pseudo-absences in terms of variables

percentage of contribution and permutation importance (Table 6 / Table B1). When the

same model was allowed to use auto-features, i.e., all the feature types available,

results were somewhat different and mostly only Chl a (mean) came out as an

important predictor (Table 6). Despite having the highest AUC (> 0.92), the predicted

suitable area seemed to be more restricted with this model set.

Page 101: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

103

Tabl

e 5: C

ompa

rison

of r

esul

ts fo

r the

Max

Ent a

nd g

ener

alise

d lin

ear (

GLM)

and

mixe

d ef

fect

s (GL

MM) m

odels

. W

here

spat

ial au

toco

rrelat

ion

was i

nclu

ded

for t

he th

ree p

seud

o-ab

senc

e tec

hniq

ues u

sed

(rand

om, ID

W an

d tu

na).

Show

n fo

r eac

h m

odel

is th

e Coh

en’s

Kapp

a sta

tistic

(κ) a

nd it

s st

anda

rd d

eviat

ion

(SD)

. The

per

cent

age o

f dev

iance

expl

ained

(%DE

) for

each

tech

niqu

e (sh

own

for t

he av

erag

e mod

el pr

edict

ions

- we

ight

ed b

y wAI

Cc) i

s also

show

n fo

r the

GL

M an

d GL

MM.

Se

ason

au

tum

n

wi

nter

sp

ring

sum

mer

Mod

el

κ SD

%

DE

κ SD

%

DE

κ SD

%

DE

κ SD

%

DE

rand

om

Ma

xEnt

0.54

0.02

- 0.5

7 0.0

8 -

0.50

0.05

- 0.4

9 0.0

6 -

GL

M 0.7

9 0.0

9 58

.0 0.5

1 0.0

9 29

.5 0.5

9 0.0

5 32

.8 0.5

1 0.0

6 17

.7

GL

MM

0.36

0.02

23.9

0.04

0.07

1.3

0.00

0.01

0.0

0.22

0.07

6.3

IDW

MaxE

nt 0.3

4 0.0

3 -

0.10

0.09

- 0.2

5 0.0

5 -

0.31

0.06

-

GLM

0.35

0.02

10.9

0.00

0.00

0.0

0.33

0.06

13.3

0.00

0.00

0.0

GLMM

0.2

1 0.0

2 4.9

0.1

7 0.0

9 10

.5 0.2

6 0.0

6 2.6

0.0

7 0.0

6 8.5

tuna

MaxE

nt 0.6

4 0.0

2 -

0.37

0.08

- 0.4

8 0.0

5 -

0.43

0.06

-

GLM

0.68

0.02

42.3

0.47

0.08

30.9

0.51

0.05

24.4

0.56

0.56

22.5

GLMM

0.5

9 0.0

2 25

.5 0.5

6 0.0

7 5.3

0.1

7 0.0

3 11

.1 0.4

3 0.0

6 23

.7

Page 102: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

104

Tabl

e 6: S

umm

ary o

f the

Max

Ent m

odels

relat

ing

prob

abilit

y of w

hale

shar

k (Rh

inco

don

typu

s, Sm

ith 18

28) o

ccur

renc

e to

indi

vidua

l oce

an p

rope

rties

. -

slope

, dep

th, d

istan

ce to

shor

e (sh

ore)

, mea

n se

a sur

face

tem

pera

ture

(SST

mea

n) an

d its

qua

drat

ic te

rm (S

ST m

ean2

), se

a sur

face

tem

pera

ture

stan

dard

dev

iatio

n (S

ST S

D)

and

chlo

roph

yll a

(Chl

a). T

hree

diff

eren

t met

hods

wer

e use

d fo

r gen

erat

ing

pseu

do-a

bsen

ces:

rand

om, in

vers

ely d

istan

t to

whale

shar

k sig

htin

g lo

catio

ns (I

DW) a

nd b

ased

on

tota

l tun

a cat

ch (t

una)

, and

resu

lts ar

e sho

wn in

rows

1 to

3 wh

en o

nly l

inea

r and

qua

drat

ic fe

atur

es w

ere u

sed.

The

two

last r

ows s

how

resu

lts w

hen

MaxE

nt m

odel

was

give

n th

e ful

l bac

kgro

und

(bac

kgro

und)

with

cova

riate

dat

a ava

ilabl

e and

vary

ing

the f

eatu

re ty

pe u

sed

in th

e mod

el. T

he J

ack-

knife

test

resu

lts sh

owin

g wh

ich en

viron

men

tal

varia

ble h

ad th

e hig

hest

gain

whe

n us

ed in

isol

atio

n (b

est a

lone

) and

whi

ch en

viron

men

tal v

ariab

les d

ecre

ased

the g

ain th

e mos

t whe

n om

itted

(wor

st w

ithou

t) ar

e also

sh

own

toge

ther

with

the v

alue o

btain

ed fo

r the

area

und

er th

e cur

ve (A

UC) t

est.

Only

top

rank

ed va

riabl

es ac

cord

ing

to p

erm

utat

ion

impo

rtanc

e are

show

n he

re -

the

com

plet

e tab

le wi

th sc

ores

per

varia

ble c

an b

e see

n in

the s

uppl

emen

tary

mat

erial

(Tab

le B1

).

Se

ason

au

tum

n wi

nter

sp

ring

sum

mer

ra

ndom

– lin

ear a

nd q

uadr

atic

feat

ures

To

p ran

ked

SST

mean

2 Ch

l a

SST

mean

de

pth

Be

st /W

orst

sh

ore

Chl a

Ch

l a / S

ST S

D de

pth

AU

C 0.7

21

0.701

0.6

91

0.668

ID

W –

linea

r and

qua

drat

ic fe

atur

es

Top r

anke

d SS

T me

an2

depth

Ch

l a

depth

Best

/ Wor

st

depth

Ch

l a

Chl a

de

pth

AU

C 0.6

25

0.574

0.6

27

0.607

tu

na –

linea

r and

qua

drati

c fea

ture

s

Top r

anke

d SS

T me

an2

SST

mean

2 SS

T me

an

depth

Best

/ Wor

st

shor

e Ch

l a

Chl a

/ SST

SD

depth

AUC

0.710

0.6

73

0.672

0.6

76

Page 103: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

105

back

grou

nd –

linea

r and

quad

ratic

featu

res

Top r

anke

d SS

T me

an2

SST

mean

2 SS

T me

an

SST

mean

Best

/ Wor

st

depth

Ch

l a

SST

mean

2 / S

hore

de

pth

AU

C 0.9

28

0.837

0.8

58

0.843

ba

ckgr

ound

– au

to fe

atur

es

To

p ran

ked

Chl a

Ch

l a

SST

mean

2 Ch

l a

Be

st / W

orst

Ch

l a

Chl a

SS

T me

an / C

hl a

Chl a

AUC

0.961

0.9

20

0.956

0.9

28

Page 104: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

106

Discussion

Niche-based models provide an alternative means for generating information about

species distributions when conventional sampling methods are expensive, logistically

difficult and produce unreliable results (e.g. when sampling for rare species - Edwards

Jr. et al., 2005; Guisan et al., 2006). In the past, occurrence data for whale sharks

have been collected at spatial scales that constitute only a small part of the animal’s

range, usually within coastal waters where nearshore aggregations form (Beckley et

al., 1997; Graham & Roberts, 2007; Jonahson & Harding, 2007; Meekan et al., 2006;

Rowat et al., 2009a). This analysis of sightings collected by fisheries in the open

ocean provides the first opportunity to predict the pelagic distribution of this wide-

ranging species, even if collected at a coarse spatial resolution (see Barbosa et al.,

2010 on downscaled projections derived from low resolution data). The predictive

maps produced by our models revealed a seasonal shift in whale shark habitat

suitability following a clockwise direction from the south-west Indian Ocean in autumn,

to the central (north and south of the equator) Indian Ocean in winter and spring, and

then back to the southern Indian Ocean in summer (Figure 12). Given that our analysis

accounted as much as possible for seasonal differences in sampling effort, this

clockwise shift likely results from seasonal changes in environmental conditions, such

as variation in temperature, that seem to be driving whale shark distribution patterns

within the Indian Ocean. Indeed, surface water properties were used before to explain

variation in the temporal distribution patterns of whale sharks (Cárdenas-Palomo et al.,

2010; Sleeman et al., 2007; Wilson, 2002; Wilson et al., 2001), although our study is

the first to test these hypotheses spatially and by season at the scale of almost an

entire ocean basin.

Page 105: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

107

Figure 12: Seasonal ensemble result for whale sharks (Rhincodon typus, Smith 1828) habitat suitability. Overlayed by sea surface temperature isolines of 26 and 29 °C. Standard deviation of the estimated probability of habitat suitability is shown with fading colours as per figure legend.

Sea surface temperature was the main variable affecting the relative

occurrence of whale sharks, with the resulting predictive maps reflecting the ’C’ shape

of the sea surface temperature patterns (Figure B3). Despite average temperatures

ranging between 23 and 34 oC (Figure B3), around 65 % of the whale shark sightings

occurred between just 27.5 and 29 oC, and 90 % occurred between 26.5 and 30 oC

(PathFinder AVHRR - temperatures recorded during the same weeks whale sharks

were spotted). It seems therefore that whale sharks use only a narrow sea surface

temperature range, which is in accordance with our hypothesis that a restricted

temperature regime exists for this species thus justifying the inclusion of the quadratic

term in our models. Whale sharks appear to avoid high temperatures which might

Page 106: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

108

elevate metabolic rates and food requirements, and excessively low temperatures that

limit metabolic function. However, even though whale sharks used only a small band

of averaged temperatures, these are not exclusive (e.g. Turnbull & Randell, 2006), and

they might move outside this envelope for other reasons such as foraging. Other

species have previously been described as predominantly occurring in a small range

of temperatures, e.g., leatherback turtle (15-33˚C; McMahon & Hays, 2006), salmon

sharks (2-8˚C in winter; Weng et al., 2005), white sharks (10-14˚C; Boustany et al.,

2002).Our results are consistent with much of what is known about the occurrence of

these sharks from finer-scale (km – 100s km) studies of nearshore aggregations in the

Indian Ocean (Figure B7). According to Rowat (2007), whale sharks aggregate off

South Africa, Mozambique and Kenya mainly in summer where our models predicted a

relatively high probability of occurrence. South African and Mozambique aggregations

persist into autumn, which is consistent with our predictions. A higher density of whale

sharks between January and May was also detected by Beckley et al. (1997) and Cliff

et al. (2007) in the far north of the South African coast. During autumn there are peaks

in whale shark abundance off Gujarat (India) and Thailand, another observation that is

in accordance with our model predictions (Theberge & Dearden, 2006 recorded an

increase in whale shark sightings starting in October and peaking in May). Peak

aggregations off Tanzania, Kenya and Seychelles occur in winter, when our 0.4

probability isoline covers these areas. In the Seychelles, whale sharks peak in

abundance during the spring and winter (Rowat & Engelhardt, 2007), while in the

Maldives, abundances peak in spring, as our models predict (Figure B7 - the 0.4

probability isoline covers the referred locations in the correspondent seasons). These

aggregations are both consistent with our model outcomes. However, peaks in

abundance of sharks observed off the coasts of Bangladesh and in the Mozambique

Page 107: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

109

Channel in spring could not be compared to our model outputs because the available

data did not extend to these regions at these times.

Our prediction of the occurrence of whale sharks along the Madagascar

coastline throughout the remainder of the year is corroborated by opportunistic

observations (Jonahson & Harding, 2007). The Mozambique Channel also has

suitable conditions for whale sharks almost year-round (Pierce et al., 2010). The

presence of sharks in the Channel might be influenced by series of rotating gyres that

spin off sequentially southwards down into the Channel (DiMarco et al., 2002); these

are thought to entrain tuna and might do the same for whale sharks. In our models, the

high likelihood of whale shark occurrence in this area was driven mainly by the

suitable sea surface temperature and/or productivity ranges observed there (we did

not have access to geostrophic current data). However, we found no strong evidence

to explain why the Channel is an important whale shark habitat, except that Chl a

never dropped below 0.1 mg m-3 there (Figure B2).

Generally, we found that sea surface temperature was a better predictor of

whale shark distributions than chlorophyll a. The latter was used as a proxy for food

availability (zooplankton) for whale sharks; although, trophic links between

phytoplankton and zooplankton are not necessarily direct, strong or immediate. There

are likely at least to be time (and therefore, spatial) lags between peaks of chlorophyll

a and zooplankton (Runge, 1998; Sleeman et al., 2010b).The filtering effects of

zooplankton on algae (sensu Runge, 1998) suggest that direct measurements of

zooplankton abundance would provide a much better predictor of whale shark

distribution than chlorophyll a per se, which appears to be the case for filter-feeding

basking sharks (Sims et al., 2005). However, such data are only available over

relatively restricted spatial scales; remote sensing provides the only means by which

Page 108: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

110

estimates of food availability can be obtained at the spatial scales relevant to oceanic

patterns.

Both genetic (Castro et al., 2007; Schmidt et al., 2009) and satellite tracking

data (Sleeman et al., 2010b; Wilson et al., 2007) indicate a capacity for long-distance

dispersal in whale sharks. Despite no observations of individually identified (Speed et

al., 2007) whale sharks traversing the Indian Ocean basin, the basin-scale shift in

distributions predicted by our models suggest at least some migratory behaviour at the

individual level, implying that broad-scale movements are possible such that

individuals could visit several known aggregation sites as they follow suitable

environmental conditions among seasons. We identified sea surface temperature as a

key determinant of whale shark distribution, thus forewarning that current aggregation

locations might shift with a changing climate.

Our models of seasonal distribution were able to predict habitat suitability for

whale sharks over a more extensive area than that covered by the sightings/tagging

data alone. They can be used to assess inter-annual variability in sightings at an

ocean scale. Fluctuating conditions measured at inter-annual scales, using remote

sensing can be used to infer inter-annual differences in the probability of occurrence

over time and space outside of our study area. Such results would assist in predicting

how seasonal aggregations might shift over space and time.

All three pseudo-absence techniques (random, IDW and tuna) resulted in

similar predictive maps for each season. However, we found major differences among

seasons within each technique, mainly in terms of the regression models’ deviance

explained. It should be noted here that because pseudo-absence locations vary within

each model, the resultant deviance explained and AUC are not directly comparable.

Nevertheless, they are useful to determine how absence locations influence the

Page 109: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

111

models results and we have used them to weight the predicted probability of

occurrence derived from each model when generating the final ensemble prediction.

GLM explained the highest deviance for almost all models, but this approach

does not account for spatial autocorrelation that can lead to an inflated explanatory

power in models of species abundance (Lichstein et al., 2002; Wintle & Bardos, 2006).

The inclusion of a spatial correlation structure is usually necessary because it likely

arises from the ecological processes that drive population dynamics (Mellin et al.,

2010). Our GLMM approach resulted in lower kappa statistics and deviance explained

by the models. This change in results is not easy to interpret (Dormann et al., 2007),

however we contend that the GLMM approach that accounted for some of the potential

spatial bias provides predictions of higher confidence, as GLM residuals are highly

spatially autocorrelated, confirming that these models are biased.

Our results show that the MaxEnt model can produce similar prediction maps

to those generated by GLM based on the same input datasets. Being a much easier

tool to employ, MaxEnt is useful to develop species distribution models quickly that

give results analogous to more robust regression models. It is noteworthy, however,

that assessment results (made here by means of Kappa statistics) were slightly lower

with MaxEnt when compared to GLM. When spatial autocorrelation is not accounted

for (either with MaxEnt or GLMs), the random method for pseudo-absences selection

generally resulted in better performances. In this context, MaxEnt models can be used

even more efficiently by using the full set of available covariate data and letting the

model randomly select the points used as ‘pseudo-absences’ (i.e., background).

Results obtained by the runs with the full dataset and the linear and quadratic features

gave similar maps and better AUC results. When the MaxEnt model used all features,

the same sort of pattern in habitat suitability was apparent; however, it produced much

Page 110: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

112

steeper spatial gradients across the region (Figure B7). This resulted in a reduced

capacity to predict some known whale shark areas of occurrence relative to GLM and

GLMM predictions (Figure B7).Additionally, MaxEnt identified depth, chlorophyll a and

sea surface temperature as the main predictors, which contrasts somewhat with the

support for more complex models using GLM and GLMM.

The coarse resolution of the input data, the necessity of relying mainly on

surface data from remote sensing (as opposed to data integrated over all depths

exploited by this species) and the lack of true absences mean that predictions should

be taken only as an index of relative probability of occurrence. The tuna purse seine

fisheries covered a large area (172,800 km2), but not all of the Indian Ocean. For this

reason there is less uncertainty in predictions for the western than the eastern part of

the ocean basin. Additionally, about 70 % of the sightings occurred during autumn,

making predictions for other seasons relatively less robust.

There is a growing demand by managers for ecologists to supply more

accurate results on the area of occurrence and distribution of ecological niches of

species. Such data are fundamental in generating appropriate protection rules for

management strategies (Beger & Possingham, 2008; Lehmann et al., 2002; Urbina-

Cardona & Flores-Villela, 2009). Combining data collected by the Regional Fisheries

Management Organisations with our modelling approach, the timing of whale shark

appearances at specific sites (e.g., at sites where they are still currently fished) can

also be predicted and subsequently used to examine the drivers of observed

population trends (Bradshaw, 2007; Bradshaw et al., 2008). Further, whale sharks are

frequently seen with large wounds or scars clearly derived from collision with boats or

ship propellers (Speed et al., 2008). Predicted areas for whale shark occurrence could

thus be used as input information for management of shipping routes. In general,

Page 111: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

113

understanding the distribution and migration patterns of whale sharks is an essential

precursor to identify possible mating and breeding areas, and to understand the

potential effects of fisheries and eco-tourism on the probability of the long-term

persistence of the species.

Page 112: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

114

CHAPTER IV. Spatial and temporal predictions of inter-decadal

trends in Indian Ocean whale sharks

A. Sequeiraa, C. Mellina,b, S. Deleana, M. Meekanc, C. Bradshawa,d

aThe Environment Institute and School of Earth and Environmental Sciences, The University of Adelaide, South Australia 5005, Australia bAustralian Institute of Marine Science, PMB No. 3, Townsville MC, Townsville, Queensland 4810, Australia cAustralian Institute of Marine Science, UWA Oceans Institute (MO96), 35 Stirling Hwy, Crawley, Western Australia 6009, Australia. dSouth Australian Research and Development Institute P.O. Box 120, Henley Beach, South Australia 5022, Australia

Marine Ecology Progress Series (in press)

Page 113: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

115

Statement of authorship

SPATIAL AND TEMPORAL PREDICTIONS OF INTER-DECADAL TRENDS IN INDIAN OCEAN WHALE SHARKS

Text in manuscript SEQUEIRA, A.M.M. (Candidate) Planned the article, performed the analysis, interpreted data, developed the models, wrote manuscript and acted as corresponding author. I hereby certify that the statement of contribution is accurate.

Signed……………………………………………………………………….Date……………

MELLIN, C. Helped generating the hypothesis, developing the models and assisted writing. I hereby certify that the statement of contribution is accurate and I give permission for the inclusion of the manuscript in the thesis.

Signed……………………………………………………………………….Date……………

DELEAN, S. Assisted data interpretation and helped with the development of the models. I hereby certify that the statement of contribution is accurate and I give permission for the inclusion of the manuscript in the thesis.

Signed……………………………………………………………………….Date……………

Page 114: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

116

MEEKAN, M. G. Supervised development of work and provided evaluation. I hereby certify that the statement of contribution is accurate and I give permission for the inclusion of the manuscript in the thesis.

Signed……………………………………………………………………….Date……………

BRADSHAW, C.J.A. Supervised development of work, helped with data and results interpretation, and assisted writing. I hereby certify that the statement of contribution is accurate and I give permission for the inclusion of the manuscript in the thesis. Signed…………………………………………………………………….Date………………

Page 115: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

117

Spatial and temporal predictions of inter-decadal trends in Indian Ocean whale sharks

Abstract

The processes driving temporal distribution and abundance patterns of whale sharks

remain largely unexplained. We present a temporal trend analysis of whale shark

occurrence at an ocean-basin scale, incorporating both spatial and temporal elements.

We test the hypothesis that the average sighting probability of sharks in the western

Indian Ocean has not changed over nearly two decades, and whether variance in this

parameter can be partially explained by climate signals. We used a 17-year dataset

(1991-2007, autumn only) of whale shark observations recorded in logbooks of tuna

purse-seiners covering most of the western Indian Ocean. We randomly generated

pseudo-absences both in space and time, and applied sequential generalised linear

mixed-effects models within a multi-model information-theoretic framework, accounting

for sampling effort and random annual variation, to evaluate relative importance of

temporal and climatic predictors on sighting probability. After accounting for seasonal

patterns in distribution, we found evidence that whale shark sighting probability varied

with fishing effort, and increased slightly in the first half of the sampling interval (1991

– 2000) to decrease thereafter (2000 – 2007). The highest-ranked model (including a

spatial predictor of occurrence, fishing effort, a quadratic term for time and a random

spatial effect) explained ~ 60 % of the deviance in sighting probability. The inclusion of

El Niño variation in the central Pacific resulted in the highest model rank, with a weak

positive effect on the sighting probability in later years. Changes in seasonal

distribution explained 27 % of the deviance in annual occurrence. We found that

sighting probability increases slightly with rising sea surface temperature in the central

Pacific Ocean and reduced temperatures in the Indian Ocean. The declining phase of

Page 116: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

118

the peak, concurrent with recent accounts of declines in population size at near-shore

aggregations and with the most pronounced global warming, deserves continued

investigation. Teasing apart the legacy effects of past exploitation and those arising

from on-going climate changes will be a major challenge for the successful long-term

management of the species.

KEYWORDS temporal trends; Rhincodon typus; tuna purse-seine fisheries;

generalised linear mixed-effects models; spatial distribution; satellite data

Introduction

Understanding the mechanisms driving observed patterns in species occurrence in

space and time is a key and challenging objective in ecology (Gotelli et al., 2010;

Pimm et al., 1995). While global declines in exploited marine fish species are well-

documented (Jackson et al., 2001; Pauly et al., 2005; Roberts, 2002), the evidence for

large shifts in distribution and abundance of much of the world’s biota arising from a

warming climate is also mounting (e.g., Parmesan, 2006; Traill et al., 2010; Walther et

al., 2002). There is, however, considerably less evidence for climate change-induced

shifts in the marine environment (but see Hoegh-Guldberg & Bruno, 2010; Last et al.,

2011; Sumaila et al., 2011; Wernberg et al., 2011), principally due to the physical and

economic constraints of collecting long-term datasets in the marine realm (Richardson

& Poloczanska, 2008). These problems are exacerbated for elusive migratory marine

species because the low probability of detection can become an important issue in

quantifying trends (Gotelli et al., 2010) and in disentangling spatial and temporal

patterns.

Page 117: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

119

The whale shark (Rhincodon typus Smith, 1828) is a highly migratory species

(Sequeira et al., 2013) found in warm and temperate waters around the globe (Last &

Stevens, 1994). Aggregating seasonally near the shore at specific coastal locations

(e.g., Rowat, 2007), it became an important species both for fishing (Pravin, 2000) and

ecotourism industries (Rowat & Engelhardt, 2007). Due to the species’ poorly

quantified population size, demography and behaviour, as well as evidence for

regional declines (Bradshaw et al., 2008; Bradshaw et al., 2007), the legal targeted

commercial fisheries have been banned (Bradshaw et al., 2008; Theberge & Dearden,

2006), and whale sharks are now classified as Vulnerable in the IUCN Red List

(www.iucnredlist.org). In contrast, whale shark ecotourism is growing worldwide (e.g.

Hsu et al., 2012; Pierce et al., 2010; Rowat & Engelhardt, 2007), but is highly

dependent on the expectation that animals will return to the same locations every year.

Reported whale shark sighting rates within these locations are highly variable,

even within the same months (e.g., de la Parra Venegas et al., 2011). Such variation

in local occurrence has been associated with fluctuation in climatic signals such as El

Niño events and the Southern Oscillation index (Sleeman et al., 2010a; Wilson et al.,

2001). There have also been several attempts to quantify trends in whale shark

population size and abundance (Meekan et al., 2006; Rowat et al., 2009a), although

these studies have been based mostly on data from near-shore aggregations

composed largely of juvenile males (Meekan et al., 2006). Due to the transitory nature

of whale shark occurrence, some have suggested that regional approaches should

instead be used to quantify broader-scale patterns spatial patterns and temporal

trends (Rowat et al., 2009b). However, expanding a study site from one aggregation to

a region (which might include multiple aggregation sites) inevitably results in adding

spatial complexity to the process.

Page 118: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

120

Partitioning variance across spatial and temporal gradients to detect patterns

in species occurrence is not usually straightforward, but can be addressed statistically

through the use of random-effects models (see Ogle, 2009; Qian et al., 2010). These

multilevel, mixed-effects or hierarchical models have been used extensively to

understand temporal trends in species assemblages (Gotelli et al., 2010), community

structure and patterns in the marine environment (e.g., MacNeil et al., 2009; Mellin et

al., 2010a), abundance and biomass of species (Ruiz & Laplanche, 2010) and

biological responses to different environmental conditions (Bedoya et al., 2011).

Using a long-term (1991 – 2007) and wide-extent dataset of whale shark

sightings in the Indian Ocean collected by the tuna purse-seine industry, we present

an analysis incorporating both spatial and temporal elements to examine temporal

trends of this species at the ocean-basin scale. Following previous work quantifying

the habitat suitability and seasonal variation in whale shark relative abundance in the

Indian Ocean (Sequeira et al., 2012), our latest approach now partitions the complex

spatial and temporal variation in whale shark occurrence patterns. Specifically, we test

the hypothesis that sighting probability remains constant over time, and quantify the

influence of global climatic signals on temporal patterns of occurrence at a broad

spatial scale.

Methods

We developed generalised linear mixed-effects models (GLMM) sequentially, with the

first step assessing the evidence for a temporal trend in whale shark occurrence, and

a second testing the hypothesis that sighting probability is correlated with variation in

Page 119: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

121

climatic indices. Below we detail the datasets used (presence/absence data and

sampling effort) and the modelling steps (predictors and model development).

Whale shark dataset

We used data recorded in logbooks from purse-seine fishers registered under the

Indian Ocean Tuna Commission (Pianet et al., 2009). These logbooks contain long-

term (1991 - 2007) data on whale shark occurrences (hereafter referred to as

‘sightings’) derived from associated net-sets for tuna catch using whale sharks as fish

aggregation devices. A total of 1185 sightings were recorded during the sampling

period, including date and location at a 0.01-° resolution (i.e., latitude and longitude

data were collected using the GCS WGS84 system and made available in units of

decimal degrees to a precision of 1/100th of a degree). The dataset provided no

indication of gender for the sighted sharks. Due to substantial fluctuation over time and

higher numbers of sightings occurring mostly in autumn (Figure 13 – top), we used

data from this season to examine the inter-annual trends in autumn whale shark

occurrence (Figure C1 in Appendix C shows spatial variation in occurrences).

Seasonal patterns of whale shark occurrence in the same dataset were previously

described by Sequeira et al. (2012).

Page 120: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

122

Figure 13: Temporal variation in whale sharks sighted by the tuna purse-seine fisheries and comparison with climatic indices. Top chart: Total number of whale sharks sighted by tuna purse-seine fisheries per month and per year. Grey bars represent autumn (Apr-Jun from 1991 to 2007). Bottom charts: Total number of whale sharks sighted in autumn per year, overlaid with the values for the Indian Ocean Dipole (IOD; left) and sea surface temperature variation in Region 4 of the central Pacific due to El Niño/La Niña events (NINO4; right).

Pseudo-absence generation

Whale shark sightings were presence-only, so we generated pseudo-absences to

produce the denominator of the logit function that allows for the binomial estimation in

our GLMM detailed below. For each presence recorded, we randomly generated 100

pseudo-absences (1:100 ratio) both (i) in space by randomly choosing non-presence

grid cells over the western Indian Ocean (function srswor – simple random sampling

without replacement – from the {Sampling} package in R), and (ii) in time by randomly

assigning to the selected point an autumn date within the 17-year interval (function

Page 121: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

123

srswr – simple random sampling with replacement – from the {Sampling} package in R)

(R Development Core Team, 2011). The high presence/pseudo-absence ratio (1:100),

which inevitably results in low prevalence (0.01), allows a better representation of the

‘background’ available, which consisted of both spatial (each grid cell) and temporal (a

specific date within the time period covered in the dataset) components.

Fisheries effort data

Effort data (number of fishing days per month) were recorded with a resolution of 5-°

within the area covered by the fisheries (30 ° N – 30 ° S and 35 °– 100 ° E; grey area

in Figure 14), giving a total of 1638 records (autumn only) and 13674 fishing days with

associated net-sets. The variables effort and number of sightings per year are

illustrated in Figure 14. Because the eastern part of the Indian Ocean (east of the

Maldives) was sampled only in one year during autumn (1998 – Figure C1), we used

only the western area of the Indian Ocean (west of the Maldives as depicted in Figure

15) in our temporal analysis.

Figure 14: Example of variation in fishing effort (days – top row) and number of whale sharks sighted (bottom row).

Page 122: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

124

Only two years of autumn (April-June from 1991 to 2007) data when effort was highest (1997 - left) and lowest (2005 - right) shown. Grey area depicts the total area sampled by the tune purse-seiners.

Figure 15: Results of the fishing effort analysis across the western Indian Ocean where the purse-seine fisheries operated in more than one year. Green line = limit of the western Indian Ocean area considered. Squares represent the 5-˚ area for which fishing effort data were available, and colours represent the effort analysis results based on the GLM evidence ratios: effort ~ Time against the null model. Values inside each square are the estimates obtained for the coefficient of the Time variable (when the null models were not ranked higher) after Holm correction.

Spatio-temporal variation in sampling effort can affect the ability to detect

temporal trends in occurrence (Phillips et al., 2009), so we developed a series of

generalised linear models (GLM) with a Poisson error distribution using time in years

(Time) as a predictor for effort (in each of the 5-° grid squares sampled more than

ER < 1 “T” positive; ER >> 150 “T” negative; ER >> 150 1 < ER < ~10

Evidence ratio (ER)

Page 123: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

125

once during autumn), to test if the spatial patterns in sampling effort were evenly

distributed throughout the 17-year period. We compared the results of these models

with a null (intercept-only) model by calculating the evidence ratio – a bias-corrected

index of the likelihood of one model over another (wAICc (GLM Time) / wAICc (GLM null)) – for

each model. To control for inflation of type I errors due to multiple testing across grid

squares, we used the Holm correction through the Bioconductor {multtest} package

(Pollard et al., 2005) in R (R Development Core Team, 2011).

Model predictors

This section describes the predictors used in each model step, detailing how we first

accounted for temporal variation both in effort and sightings (Step 1), and then tested

for correlations between observed trends and changes in climatic predictors (Step 2).

Step 1

Because temporal and spatial variation are seldom dissociated, we needed to include

a term covering the variation in spatial probability of whale shark occurrence to test the

hypothesis that the average probability of sightings remained constant over time within

the large area under study. For this we used the results derived from Sequeira et al.

(2012), where we fitted models of seasonal spatial distribution of whale sharks in the

Indian Ocean. Here we re-fitted the whale shark distribution model for autumn

(Sequeira et al., 2012), including some modifications to the likelihood estimation,

pseudo-absence ratio, and covariate treatments (see Supplementary Information in

Appendix C for a detailed description of these changes). We then used logit-scale

predictions of the likelihood of whale shark occurrence in autumn (spatial probability -

Page 124: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

126

SpatialP; Figure C2) within each 9-km grid cell (resolution used in Sequeira et al.,

2012) as an explanatory variable of whale shark occurrence in the temporal models

developed herein. The inclusion of SpatialP in the GLMM (see below) accounts for

possible spatial autocorrelation.

Temporal changes in effort can explain some of the inter-annual variation

detected in sightings; therefore, we also added effort as a predictor in our temporal

models to account for its potential effect on the temporal patterns observed. Because

the mean effort for autumn was already accounted for within the spatial predictor, we

only included the temporal variation around the mean (i.e., the zero-centred inter-

annual deviations around mean effort). We included both a fixed and a random effect

for time (Time and year, respectively) to account for inter-annual variability in

presences, allowing the random structure to contain only information that could not

have been modelled with fixed effects (following Zuur et al., 2009).

Step 2

To test the hypothesis that inter-annual variation in whale shark sightings was

correlated with variation in indices of sea surface temperature and air pressure as

reported previously in a near-shore aggregation (Sleeman et al., 2010a; Wilson et al.,

2001), we considered variation in climatic indices in both the Indian and the Pacific

Oceans. We tested 4 indices: (i) the Indian Ocean Dipole (IOD – Saji et al., 1999), (ii)

El Niño variation in the central Pacific Region 4, 160 °E – 150 °W / 5 °S – 5 °N

(NINO4, Burgers & Stephenson, 1999), (iii) the Oceanic Niño Index, ONI – three-

month moving averages of sea surface temperature in the Niño 3.4 Region -170 – 120

°W / 5 °S – 5 °N and the (iv) Southern Oscillation Index (SOI – Walker, 1925). We

Page 125: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

127

collected the online climatic indices for the total period covered by the purse-seine

fisheries from the Earth System Research Laboratory (U. S. Department of

Commerce, National Oceanic and Atmospheric Administration, Boulder, Colorado,

www.esrl.noaa.gov/index.html), the Japanese Agency for Marine-Earth Science and

Technology (www.jamstec.go.jp/frsgc/research/d1/iod) and the Australian Government

Bureau of Meteorology (www.bom.gov.au). To understand how the different indices

are associated with each other and to assist interpretation of the model results, we

investigated the correlation between the climatic predictors using the pairs.panels

function in the package {psych} in R (R Development Core Team, 2011) and their

monotonic relationships using Spearman’s rank correlation (ρ). IOD is weakly collinear

with both NINO4 and ONI (Pearson coefficients = 0.06 and -0.06, respectively) while

the two latter predictors are highly correlated (Pearson coefficient = 0.80) (Figure C3).

Spearman coefficients also showed collinearity only between NINO4 and ONI (ρ =

0.8).

Models

We applied generalised linear mixed-effects models (GLMM) with a binomial error

distribution and a logit link function to compare the predictive ability of different

combinations of the predictors. The mixed-effects models we developed in each step

included all possible combinations of the fixed and random effects (Step 1), and each

of the four individual climatic predictors (Step 2). By including climatic variables one at

a time, we concurrently tested whether replacing Time by any of the climatic variables

could explain away any trends observed in Step 1.

In each step, we compared models based on the Akaike’s information criterion

corrected for small sample sizes (AICc) (Burnham & Anderson, 2004), which favours

Page 126: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

128

models with higher predictive capacity when sample sizes are large and tapering

effects exist – as expected to occur in our spatio-temporal models. We assessed each

model’s strength of evidence relative to the entire model set by calculating AICc model

weights (wAICc) and used the percentage of deviance explained (%De) to quantify

each model’s goodness-of-fit. We retained the AICc top-ranked model from the first

step and used it in the second step as a control model for the more complex

combinations with climatic predictors. We developed all models in R version 2.11.1(R

Development Core Team, 2011).

Results

Fisheries effort analysis

The spatial analysis of effort (GLM models with Poisson distribution) per grid cell (5-°

resolution) demonstrated no evidence of temporal trend in effort in approximately 60 %

of the grid cells (Figure 15). In the remaining cells, the evidence ratios of the models

including Time were high (>> 150; Figure 15), and effort generally increased over time

(estimated coefficients ranging from 0.03 ± 0.007 to 0.23 ± 0.046; Figure 15) with the

exception of only four cells where effort decreased with time (coefficients ranging from

-0.04 ± 0.005 to -0.12 ± 0.02; Figure 15). Two of these four declining-effort cells

corresponded to the area where most of the whale shark sightings occurred (compare

Figure 15 and Figure C1).

Page 127: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

129

Models

Step 1

The GLMM with the highest information-theoretic support (wAICc = 0.45; Step 1)

included the spatial predictor (i.e., spatial predictions of the probability of whale shark

occurrence in autumn on a logit scale – SpatialP; Figure C2), the zero-centred effort

(inter-annual variation around the autumn mean, i.e., the variation not accounted for by

the spatial predictor), and a quadratic temporal trend with a random intercept term

accounting for the among-year variance (Table 7). This model explained almost 60 %

of the deviance in whale shark sighting probability. We found evidence for a slight

increase in the probability of whale shark sightings for the first half of the sampling

period (1991-2000), and for a declining trend over the last half (2001-2007) (Figure

16). However, total wAICc was shared approximately evenly among the two top-ranked

models (Table 7; only one of them including the Time predictor), indicating that the

temporal trend was only weak. Results for models not including the spatial term

(SpatialP) generally performed poorly relative to the other models in the set, with this

predictor

Page 128: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

130

Tabl

e 7: S

umm

ary o

f the

gen

erali

sed

linea

r mixe

d-ef

fect

s mod

els re

latin

g pr

obab

ility o

f wha

le sh

ark o

ccur

renc

e to

spat

ial an

d te

mpo

ral p

redi

ctor

s. “S

patia

lP” –

a sp

atial

pre

dict

or d

erive

d fro

m p

revio

us sp

atial

dist

ribut

ion

mod

els (S

eque

ira et

al., 2

012)

, effo

rt, T

ime f

ixed

effe

ct p

redi

ctor

for t

ime (

in ye

ars)

, and

a ra

ndom

effe

ct

for y

ear (

Step

1) an

d gl

obal

clim

atic

pred

ictor

s: In

dian

Oce

an D

ipol

e (IO

D), S

outh

ern

Oscil

latio

n In

dex (

SOI),

El N

iño

in th

e cen

tral P

acifi

c - R

egio

n 4 (

NINO

4) an

d Oc

eani

c Niñ

o In

dex (

ONI)

(Ste

p 2)

. Sho

wn fo

r eac

h m

odel

are t

he n

umbe

r of p

aram

eter

s (k)

, log-

likeli

hood

(LL)

, bias

ed-c

orre

cted

mod

el ev

iden

ce b

ased

on

weig

hts o

f Aka

ike’s

info

rmat

ion

crite

rion

corre

cted

for s

mall

sam

ple s

izes (

wAIC

c) an

d th

e per

cent

age o

f dev

iance

expl

ained

(%De

). To

p-ra

nkin

g/be

st-p

erfo

rmin

g m

odels

in ea

ch st

ep ar

e in

bold

and

mod

els ar

e or

dere

d by

dec

reas

ing

wAIC

c. Va

lues

< 0.

001 n

ot sh

own.

M

odel

k

LL

wAIC

c %

De

Step

1

S

patia

lP +

effo

rt +

Tim

e + T

ime2 +

(1| y

ear)

7 -1

956.0

8 0.4

54

57.0

Spat

ialP

+ ef

fort

+ (1

| yea

r) 5

-195

8.29

0.37

0 56

.9

Sp

atial

P +

effo

rt +

Tim

e +

(1| y

ear)

6 -1

958.0

3 0.

176

56.9

Spat

ialP

+ ef

fort

5

-253

1.80

<0

.001

44

.3

Spat

ialP

+ ef

fort

+ Ti

me

6 -2

531.

72

<0.0

01

44.3

Spat

ialP

+ Ti

me

+ Ti

me2 +

(1| y

ear)

6 -2

707.

72

<0.0

01

40.4

Spat

ialP

+ (1

| yea

r) 4

-271

0.93

<0.0

01

40.3

Sp

atial

P +

Tim

e +

(1| y

ear)

5 -2

710.8

7 <0

.001

40

.4

ef

fort

+ Ti

me

+ Ti

me2 +

(1| y

ear)

6 -3

201.4

1 <0

.001

29

.6

ef

fort

+ (1

| yea

r) 4

-320

3.59

<0.0

01

29.5

effo

rt +

Tim

e +

(1| y

ear)

5 -3

202.9

5 <0

.001

29

.5

Spat

ialP

+ Ti

me

5 -3

313.7

7 <0

.001

27

.1

Page 129: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

131

Sp

atial

P 4

-331

7.50

<0

.001

27

.0

Ti

me

+ Ti

me2 +

(1| y

ear)

5 -3

995.6

9 <0

.001

12

.1

1 +

(1| y

ear)

3 -3

998.6

7 <0

.001

12

.0

Tim

e +

(1| y

ear)

4 -3

998.2

3 <0

.001

12

.0

effo

rt

4 -4

025.

96

<0.0

01

11.4

1 3

-454

4.22

<0

.001

0

Step

2

Sp

atial

P +

effo

rt +

Tim

e + T

ime2 +

NIN

O4 +

(1| y

ear)

8 -1

861.7

5 0.8

1 59

.0

Spat

ialP

+ eff

ort +

NIN

O4 +

(1| y

ear)

6 -1

865.1

8 0.

19

59.0

Spat

ialP

+ ef

fort

+ Ti

me

+ Tim

e2 + O

NI +

(1| y

ear)

8 -1

926.1

5 <0

.01

57.6

Spat

ialP

+ ef

fort

+ ON

I + (1

| yea

r) 6

-192

9.97

<0

.01

57.5

Spat

ialP

+ ef

fort

+ Ti

me

+ Tim

e2 + IO

D +

(1| y

ear)

8 -1

954.4

4 <0

.01

57.0

Spat

ialP

+ ef

fort

+ IO

D +

(1| y

ear)

6 -1

956.

63

<0.0

1 56

.9

Sp

atial

P +

effo

rt +

Tim

e +

Time2 +

(1| y

ear)

7 -1

956.

08

<0.0

1 57

.0

Sp

atial

P +

effo

rt +

(1| y

ear)

5 -1

958.2

9 <0

.01

56.9

Spat

ialP

+ ef

fort

+ Ti

me

+ Tim

e2 + S

OI +

(1| y

ear)

8 -1

956.

00

<0.0

1 57

.0

Sp

atial

P +

effo

rt +

SOI +

(1| y

ear)

6 -1

958.

16

<0.0

1 56

.9

1

3 -4

544.

22

<0.0

1 0

Page 130: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

132

alone explaining 27 % of the deviance (Table 7). A positive relationship with the zero-

centred inter-annual deviations from mean fishing effort (coefficient estimate = 1.63 ±

0.06) indicates that the number of sightings is higher when effort is higher than

average and vice versa.

a) b)

Figure 16: Partial effect of time on whale shark presence. (a) on the log-odds scale showing the rate of change and (b) on the probability scale, showing the effect of time on the probability of whale shark presence. Dotted lines represent 95 % confidence interval.

Step 2

Including the index of sea surface temperature in the central Pacific (NINO4) in the

highest-ranked model from Step 1 above resulted in the highest statistical support

(wAICc = 0.81). However, the percentage of deviance explained (59 %) was only

slightly higher than the model excluding the climatic predictor (57 %). All models that

included a climatic predictor reflecting variation in sea surface temperature (NINO4,

IOD, ONI) had higher support than those without climatic variables (from Step 1), while

the model including SOI (relative to air pressure variation) performed poorly even

when compared to models excluding climate signals (Table 7). The partial effects of

Page 131: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

133

the climatic variables showed that an increase in NINO4 (reflecting higher sea surface

temperature in the central Pacific Ocean) had a positive effect on whale shark

probability of occurrence in the western Indian Ocean (Figure 17), and an increase in

IOD (reflecting higher sea surface temperature in the western part of the Indian

Ocean) had a negative effect (Figure C4).

Figure 17: Partial effects of NINO4 and Time (after accounting for climatic contributions) in the probability of whale shark presence. Top row: Partial effect of NINO4 in the probability of whale shark presence. NINO4 is an index representing sea surface temperature variation in the central Pacific Region 4 (160 °E – 150 °W / 5 °S – 5 °N; Burgers and Stephenson, 1999). Bottom row: The partial effect of Time after accounting for climatic contributions to whale shark probability of presence is shown. Dotted lines represent 95 % confidence interval.

Log-odds Probability

NIN

O4

Tim

e (w

ith N

INO

4)

Page 132: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

134

Discussion

Access to a long-term dataset of whale shark sightings covering the western sector of

an entire ocean basin provided a unique opportunity to analyse temporal trends and

variation in whale shark occurrence at a scale more likely to encompass the range of

these highly migratory animals, and for a greater proportion of the population than in

previous studies. Overall, our results highlighted a high inter-annual consistency in

whale shark distribution patterns, with the spatial predictor (i.e., mean seasonal

distribution) alone accounting for 27 % of the deviance. We also found evidence for a

modest peak in whale shark occurrences in the middle of the time series (~ 2000),

which prevailed after accounting for changes in global climatic indices. To date,

analogous temporal analysis of whale shark occurrence have only been done at the

scale of single aggregations (~ 10s of km; Sleeman et al., 2010a) covering only a

small part of this species’ range and only a small proportion of the population (mostly

male juveniles, Meekan et al., 2006). Thus, our results are the first to estimate ocean-

scale trends in occurrence that include several known aggregation sites for the

species, and it is the first analysis to account simultaneously for spatial and temporal

components, including effort, that can mask underlying trends in the probability of

occurrence.

The high percentage of deviance explained by the spatial predictor (27 %)

implies that distribution patterns are annually consistent (autumn only), with a higher

probability of occurrence around the Mozambique Channel. This consistency might be

due to permanent characteristics of the area, such as physical features enhancing

upwelling, local productivity and tolerable sea surface temperatures (Sequeira et al.,

2012).

Page 133: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

135

Modelling must ensure that variation in sampling effort does not confound

results (Dennis et al., 1999). In our dataset, effort generally increased through time,

although there was a decline in the Mozambique Channel (Figure 15) where most

sightings occurred during the 17 years (compare with Figure C1). By including the

zero-centred inter-annual deviations from mean fishing effort in our models (which

represent temporal fluctuations from the mean within each grid cell) (Sequeira et al.,

2012), we accounted for the contribution of effort in the temporal probability of

occurrence. Combining the spatial predictor just with effort, our models explained ~ 44

% of the deviance (Table 7), demonstrating the importance of incorporating both

spatial and sampling effort components in temporal models.

We hypothesise that the non-linear temporal trend we observed (Figure 16),

even though weak, reflects variation in whale shark abundance in the western area of

the Indian Ocean. Such a trend, where the probability of occurrence first increases and

then decreases, could represent an inter-decadal cycle (i.e., cycle over 15 years),

similar to what has been observed for strandings of other large marine species in the

Southern Ocean (Evans et al., 2005) and fish catches in the same region (Jury et al.,

2010). In the latter study, such cycles were only just detected even with annual catch

records spanning 40 years. Our dataset, covering < 20 years, revealed evidence

(although weak) for a peak in whale shark occurrence that is in accordance with the

longer cycles found for fish catches (Jury et al., 2010).

Focusing only on the declining segment of the identified peak, which might be

a reflection of a long-term decline in relative abundance, we suggest three competing

hypotheses to explain it: (i) a downward component of possible inter-decadal

abundance (as presented above), (ii) a real decline in abundance, or (iii) a

distributional shift due to changing habitat characteristics (e.g., via climate change).

Page 134: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

136

Considering the decline observed as reflecting a real decline in abundance, our results

support previous data (Bradshaw et al., 2008) and predictions (Bradshaw et al., 2007)

from Ningaloo in Western Australia suggesting a decline in population size. In fact, the

South-East Asian and Indian whale shark-targeted fisheries (Hsu et al., 2012; Pravin,

2000) were banned due to the low numbers captured even with increasing effort – a

signal that can indicate population decline. Despite the bans, targeted fisheries still

present a risk to whale sharks due to lack of enforcement (Stewart & Wilson, 2005)

and there is evidence that illegal fishing is still occurring (Riley et al., 2009; White &

Cavanagh, 2007). Such illegal fisheries could still be responsible for on-going declines,

although it is currently impossible to determine to what extent this might be still

occurring. The unintentional bycatch of whale sharks in other fisheries (e.g. purse

seiners; Romanov, 2002) also remains unquantified.

The decline segment is also concurrent with the most pronounced warming

observed over the last decade, such that a related hypothesis where latitudinal shifts

in distribution (e.g., to areas not considered in this study) resulting from rising water

temperatures (as suggested in Sequeira et al., 2012) driven by global climate change

could be influential. Our models’ estimates suggest that sighting probability increases

with rising sea surface temperature in the central Pacific Ocean (Figure 16) and

reduced temperatures in the Indian Ocean (Figure C4). Despite IOD being a climatic

index directly measured in the Indian Ocean, our models gave higher support for

NINO4 as a predictor. However, there were some occasions where a few whale

sharks were sighted despite high NINO4 values (Figure 13). This is particularly

interesting for 1997 when not only the NINO4 index was high, effort was also maximal

(Figure 14). Our data covered mainly the western Indian Ocean, but there was a peak

in sightings at Ningaloo in 1997 (Wilson et al., 2001). The same occurred in 1992

Page 135: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

137

(Wilson et al., 2001) and 2002 (see Sleeman et al., 2010a), when NINO4 was high and

few sharks were observed in the western Indian Ocean. In both 1997 and 2002, IOD

was positive (0.04 to 0.88) and coincided with a decrease in probability of occurrence

in the western Indian Ocean (Figure C4). When IOD is positive, the thermocline

deepens in the eastern Indian Ocean, resulting in more intense upwelling and greater

surface productivity (Behera & Yamagata, 2003). Saji et al. (1999) first described IOD

as independent of ENSO (El Nino / Southern Oscillation), although others suggest

possible associations between the two (reviewed by Maity et al., 2007). Understanding

how these coupled sea surface-atmosphere phenomena are related is not our aim, but

the (weak) support for Pacific and Indian Ocean indices suggest some possible effects

of broad-scale oceanic climate events influencing whale shark habitat use.

Despite cycles in fish abundance in the western Indian Ocean being directly

related to environmental fluctuation, most of the variance in fish catch can be

explained with estimates of local productivity (Jury et al., 2010). Changes in climate

are already affecting the base of the principal food web – phytoplankton (e.g. Edwards

& Richardson, 2004). This evidence, coupled with our demonstration of possible

climate-influenced distribution of the world’s largest fish, demonstrates a need to

predict how species will respond to future climate scenarios. As such, long-term

management of whale sharks requires both a better understanding of basic ecology

and demography, and how these will be altered as the oceans warm.

Page 136: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

138

CHAPTER V. Inter-ocean asynchrony in whale shark occurrence

A. Sequeiraa, C. Mellina,b, M. Meekanc, C. Bradshawa,d

aThe Environment Institute and School of Earth and Environmental Sciences, The

University of Adelaide, South Australia 5005, Australia

bAustralian Institute of Marine Science, PMB No. 3, Townsville MC, Townsville,

Queensland 4810, Australia

cAustralian Institute of Marine Science, UWA Oceans Institute (MO96), 35 Stirling

Hwy, Crawley, Western Australia 6009, Australia.

dSouth Australian Research and Development Institute P.O. Box 120, Henley Beach,

South Australia 5022, Australia

(Manuscript in preparation)

Page 137: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

139

Statement of authorship

INTER-OCEAN ASYNCHRONY IN WHALE SHARK OCCURRENCES Text in manuscript SEQUEIRA, A.M.M. (Candidate) Planned the article, performed the analysis, interpreted data, developed the models, wrote manuscript and acted as corresponding author. I hereby certify that the statement of contribution is accurate.

Signed……………………………………………………………………….Date……………

MELLIN, C. Helped planning the article, and assisted data interpretation and models development. I hereby certify that the statement of contribution is accurate and I give permission for the inclusion of the manuscript in the thesis

Signed.…………………………………………………………………….Date………………

MEEKAN, M. G. Provided evaluation. I hereby certify that the statement of contribution is accurate and I give permission for the inclusion of the manuscript in the thesis

Signed……………………………………………………………………….Date……………

Page 138: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

140

BRADSHAW, C.J.A. Supervised development of work, helped in model results interpretation and manuscript evaluation. I hereby certify that the statement of contribution is accurate and I give permission for the inclusion of the manuscript in the thesis.

Signed……………………………………………………………………….Date……………

Page 139: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

141

Inter-ocean asynchrony in whale shark occurrence

Abstract

The whale shark is a highly migratory species, with genetic evidence for inter-ocean

connectivity. Given this migratory behaviour, population trends can only be understood

by examining patterns of occurrence synchrony among aggregation locations. We

present a two-step modelling approach of whale shark spatial and temporal

occurrence in the Atlantic and Pacific Oceans by applying generalised linear mixed-

effects models. To test the hypothesis that the probability of whale shark occurrence is

asynchronous across oceans, as expected if inter-ocean migration is occurring, we

used long-term datasets of whale shark sightings (April to June) derived from tuna

purse-seine logbooks covering most of the central-east Atlantic (1980-2010) and

western Pacific (2000-2010). We first predicted seasonal habitat suitability and

produced maps of shark distribution in each area, and then evaluated the relative

effect of time (year) on sighting probability. Additionally, we applied fast Fourier

transforms to determine if any periodicity was apparent in whale shark occurrences in

each ocean. After partialling out the effects of seasonal patterns in spatial distribution

and sampling effort, we found no evidence for a temporal trend in the Atlantic.

Conversely, we detected an increase in the whale shark probability of occurrence in

the Pacific. The highest-ranked model for the latter included a spatial predictor of

whale shark occurrence along with fishing effort, a linear term for time, and a random

temporal effect (year) (15 % deviance explained). Fast Fourier transforms revealed a

prominent 15-year cycle for shark occurrences in the Atlantic. We further reveal that

the increase in probability of shark occurrence in the Pacific is concurrent with a

decrease in the Indian Ocean. We conclude that cyclic patterns driven by migratory

behaviour would better explain temporal trends in whale shark occurrence at the

Page 140: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

142

oceanic scale. Despite cycles partially explaining observations of fewer sharks in some

years, overall reported sighting rate has been decreasing. More data are needed to

examine the flow of individuals across oceans to isolate mechanisms driving change.

KEYWORDS temporal trends; Rhincodon typus; tuna purse-seine fisheries;

generalised linear mixed-effects models; spatial distribution; satellite data

Introduction

While species loss in terrestrial systems is strongly associated with habitat

degradation (e.g., Pimm & Raven, 2000), most of humanity’s obvious negative impacts

in marine ecosystems result from direct exploitation (Pauly et al., 1998; Worm et al.,

2006) or related by-catch of marine life (Agardy, 2000; Hall et al., 2000), and now

climate change is also beginning to impose its toll (e.g., Sumaila et al., 2011).

Reported declines in marine species challenge the idea that extinctions in the oceans

are unlikely (Hendriks et al., 2006), and raise awareness that understanding

connectivity is paramount to assist sustainable management (Agardy, 2000).

Whale sharks (Rhincodon typus, Smith 1928) travel large distances (e.g.,

Rowat & Gore, 2007), and its sub-populations are expected to be connected across

the world’s oceans (Castro et al., 2007; Sequeira et al., 2013). Due to the species’

highly migratory behaviour, concerns regarding the adequacy of current management

measures have been raised (Rowat, 2007). These measures mostly concern confined

areas where these sharks are economically important (e.g., for tourism) (Pierce et al.,

2010; Quiros, 2007), and might neglect negative impacts occurring elsewhere

(Bradshaw, 2007). Whale shark-based eco-tourism has been developed based on the

anticipation that individuals return to the same location each year at approximately the

Page 141: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

143

same time (Taylor, 1996); however, evidence for declining relative abundance has

been reported at some of these locations (Bradshaw et al., 2008; Theberge &

Dearden, 2006). There is also some quantitative support for a slight reduction in the

probability of occurrence in the Indian Ocean during the last decade (Sequeira et al., in

press).

Whale sharks are potentially affected by a range of human-driven alterations

to the marine environment, including exploitation through direct commercial fisheries

(last ban occurred in Taiwan only in 2008) (Council of Agriculture, 2007), poaching

(Riley et al., 2009), by-catch (Romanov, 2002), and habitat disturbance (Heyman et

al., 2010). Anthropogenic climate disruption is another potential stressor because

temperature is an important predictor of whale shark distribution (Sequeira et al.,

2012) and relative abundance (Sleeman et al., 2010a). Alternatively or synergistically,

observed short-term declines could also be confounded by inter-decadal cycles in

relative abundance (Sequeira et al., in press) associated with broad-scale migration

patterns. Because this species is highly mobile and populations are connected across

oceans at least at the generational scale (Castro et al., 2007; Schmidt et al., 2009),

temporal trends can only be truly revealed by comparing the synchrony of occurrence

patterns among locations across ocean basins.

Temporal trends in species occurrence are seldom dissociated from spatial

processes. Even though statistical models have been mostly used to assess and

predict the spatial distribution of species (Guisan & Thuiller, 2005; Guisan &

Zimmermann, 2000; Hirzel et al., 2002; Phillips et al., 2009) based on the ecological

niche (Kearney & Porter, 2009), they can also be used to assess temporal trends

(Gotelli et al., 2010). For example, species distribution models have indeed been used

to estimate habitat suitability for highly migratory marine species (Elith & Leathwick,

Page 142: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

144

2009; Oviedo & Solís, 2008; Praca & Gannier, 2007), and also to assess their

temporal trends (Sequeira et al., in press).

Access to the logbooks compiled by tuna purse-seiners from the Atlantic and

Pacific Oceans gave us the opportunity to use similar models to estimate broad-scale

trends in whale shark occurrence to complement the assessment made previously for

the Indian Ocean (Sequeira et al., in press). Here we: (1) predict whale shark habitat

suitability within the areas covered by the tuna fisheries in the Atlantic and Pacific, (2)

test the hypothesis of temporal asynchrony in probability of occurrence, and (3) assess

possible cyclic patterns in occurrence. Our main objective is to provide a temporal

assessment of whale shark occurrence probability encompassing most of the species’

known geographical range by comparing their probabilities of occurrence in different

oceans.

Methods

With the main objective to assess temporal trends in whale shark occurrence in the

Atlantic and Pacific oceans, and compare them with the results obtained previously for

the Indian Ocean (Sequeira et al., in press; Sequeira et al., 2012), the models we

developed herein follow a similar approach. First we developed habitat suitability

models and used the resulting habitat suitability as part of the input data to model

temporal trends in whale shark occurrence. Below we describe the data used

(opportunistic fisheries-derived and environmental datasets), the modelling methods,

including how we accounted for pitfalls in the opportunistic dataset used (e.g.,

presence only-data and sampling bias), and how we used the Fourier Transforms to

look for possible cyclic patterns in the whale shark sightings dataset.

Page 143: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

145

Data

We used whale shark occurrence data from the Atlantic and Pacific Oceans recorded

in the logbooks of tuna purse-seiners. Because tuna aggregate underneath whale

sharks (possibly foraging on similar preys), fishers deploy nets around the sharks to

facilitate harvest of the target fish (Matsunaga et al., 2003). Hereafter, we use the term

‘sighting’ to describe logbook records of these net sets associated with whale sharks.

The dataset made available by the Institut de Recherche pour le Développement

(France) and the Secretariat of the Pacific Community comprises most of the central

area of the Atlantic (21 °N to 15 °S and 34 °W to 14 °E) and central western Pacific

(15 °N to 15 °S and 130 °E to 150 °W) (Figure 18). It includes date (month and year),

longitude and latitude information for whale shark sightings (0.01° precision), and

spatially aggregated information on sampling effort (number of days spent fishing) per

month and per grid cell of 1-° resolution in the Atlantic, and 5-° resolution in the Pacific

(Figure 18). No information on individual vessel or trip units was available. To keep the

highest resolution possible in the input data for our models, effort was included as an

offset - as detailed below. The data spanned 1980 to 2010 in the Atlantic (total of

18277 records), and 2000 to 2010 in the western Pacific (total of 2272 records

provided by only part of the fleets registered with the Secretariat of the Pacific

Community, but representative of the fisheries in the area – P. Williams pers. comm.).

To compare our results for the Atlantic and Pacific with previous results for the Indian

Ocean (Sequeira et al., in press), and due to the generally low number of sightings in

other seasons for each ocean, we used data only for the months of April to June

(austral autumn).

Page 144: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

146

Figure 18: Whale shark (Rhincodon typus, Smith 1828) datasets recorded by tuna purse-seiners including location of net-sets in the (left) Atlantic (1980-2010, data provided by the Institut de Recherche pour le Développement - France) and (right) Pacific (2000-2010, data provided by the Secretariat of the Pacific Community). Top panel: area covered by the fisheries (grey) and location of whale shark sightings (blue dots). A total of 1030 sightings were recorded in the Atlantic, and 167 in the Pacific Ocean (between April and June). Bottom panel: number of fishing days (effort) with a resolution of 1-° for the Atlantic and 5-° for the Pacific Ocean.

We assembled environmental data on daytime sea surface temperature (SST

in °C) and chlorophyll a (Chl a in mg m-3) at a 9-km resolution derived from the

Advanced Very High Resolution Radiometer (AVHRR) PathFinder version 5.0 and

Sea-viewing Wide Field-of-view Sensor (SeaWiFS) satellites, respectively. We used

ArcToolBox functions (ArcGIS 9.3.1TM automated with Python scripts) to calculate

mean and standard deviation of SST and Chl a per grid cell for all weekly composites

between April and June for the time period of each ocean dataset. We also derived

depth (m), slope (º) and distance to shore (km; using the Near tool in ArcGIS 9.3.1TM

on a equidistant cylindrical coordinate system) from the General Bathymetry Chart of

Page 145: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

147

the Oceans (GEBCO, 2003). We then collated the full dataset with a common

resolution of 9-km including six predictors: mean depth, slope, distance to shore, mean

SST, SST standard deviation and mean Chl a.

Models

We developed the modelling approach in two steps to (1) compare the spatial

predictive ability of different combinations of the environmental predictors in modelling

whale shark habitat suitability, and (2) assess evidence for a temporal trend in whale

shark occurrence in each ocean. In both steps, we applied generalised linear mixed-

effects models (GLMM) – with a binomial error distribution and a logit link function – to

our presence-only data by generating pseudo-absences for binomial estimation.

The process of generating pseudo-absences differed in each modelling step.

In the first step, we randomly generated 10 pseudo-absences per presence based on

a spatially random distribution within the area covered by the fisheries (excluding all

the presence cells). In the second step, we generated 100 pseudo-absences per

presence based on both temporally and spatially random distributions, that is,

randomly choosing a date within the temporal coverage of each dataset, and then

randomly assigning to it a location within the area covered by the fisheries (for each

ocean). In both steps, we generated the spatially random distributions with the srswor

function (simple random sampling without replacement) from the {sampling} package

in the R programming language (R Development Core Team, 2012). For the

temporally random pseudo-absence distribution, we randomly selected a date within

the temporal coverage of each dataset (April to June only) by using the srswr function

(simple random sampling with replacement) from the same package in R. To assess

the influence of the date and location associated with the selected pseudo-absences,

Page 146: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

148

we generated each set of pseudo-absences 100 times prior to their use in the spatial

models, and 10 times for the temporal models (due to computing time).

In the first modelling step (i.e., spatial models), we included in our list of

predictors the quadratic terms (nonlinear response) for SST and depth to account for a

possible higher suitability for whale shark occurrence within a range of these variables.

We did this by using the poly function (with degree 2) from the {stats} package in R.

We also included a code for each 1-° grid cell as a spatial random effect to reduce

spatial autocorrelation. To account for the sampling bias associated with effort (more

whale shark sightings expected in areas where fishing effort was higher), and to keep

the highest resolution possible in the input data, we have included effort as an offset in

the models.

In the second modelling step (temporal models), we used as predictors the

outputs from step 1 (i.e., the resulting spatial habitat suitability – Hsuit), a zero-centred

effort term (i.e., inter-annual variation around the mean not accounted for within the

spatial predictor) (Sequeira et al., in press) and time, both as a fixed (time) and

random effect (year) to ensure that the random structure contains only information that

could not have been modelled with fixed effects (following Zuur et al., 2009). To

account for a parabolic-like dependence of occurrences with time, we also added the

quadratic term for time using the poly function (with degree 2).

A generic way to write the GLMM (with a logit link function) is:

where Presence is the expected mean probability of sighting occurrence. and

represent the fixed and random covariates used in the models: the environmental

predictors and the spatial grid in the spatial models, and habitat suitability, zero-

Page 147: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

149

centred effort and time, and year in the temporal models. The index i corresponds to

the number of observations among grid-cells in the spatial models, and years in the

temporal models. and represent the parameters of the fixed and random effects,

respectively, and � is the intercept. The log of fishing effort was included as an offset

only in the spatial models.

We compared the models relative strength of evidence by weighting each

model’s Akaike’s information criterion corrected for small sample sizes (wAICc)

(Burnham & Anderson, 2004), and assessed the goodness-of-fit by calculating the

percentage of deviance explained (%De) for each model. We calculated the 10-fold

cross-validation error for the model with highest wAICc support and assessed the

model’s predictive power with the � statistic (Cohen, 1960). To build the habitat

suitability maps (as result of the first spatial modelling step), we weighted each

model’s predictions according to its weight of evidence (wAICc), and used the result of

the weighted mean prediction for all models as input to the temporal models. We also

calculated the weight of evidence for each predictor used in the temporal models by

summing the wAICc weights over all models in which each predictor appeared. This

allows estimation of the predictor with the highest relative importance to explain the

response variable.

Fourier Transforms

Fourier transforms allow for the decomposition of signals (i.e., time series) into the

sum of sinusoidal curves with different frequencies – a highly useful approach for

ecological time series (Platt & Denman, 1975). To analyse possible cyclic variation in

whale shark occurrences, we applied the fast Fourier transform function (FFT;

Page 148: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

150

following Moler, 2004) to the sightings time series after correcting for effort bias (i.e.,

standardizing sightings per unit effort [SPUE], with an effort unit = 1 fishing day). We

used MatLab version 7.12.0.635 (R2011a) (The Mathworks Inc., Natick, MA), and

interpreted the results of the FFT by plotting the periodogram for the function: power

against the inverse of frequency (time). A cycle is then defined by identifying the

strongest frequency.

To examine cyclic trends apart from the known annual cycles (i.e., number of

sightings is usually higher in particular seasons each year), we started by only using

data for the season in study (Apr-Jun), and defining frequency as year. As an

alternative test to include all sightings data available (i.e., all months), and therefore

increase the length of the vector used in the function, we also ran FFT for the sightings

per monthly effort (i.e., defining frequency as month). In this test, we used a running

average including the six months before and after each month to eliminate the known

annual periodicity. The strength of each resulting cycle (periodicity) is given by the

power of the corresponding frequency in the periodogram.

Results

Spatial patterns in whale shark occurrence

In the Atlantic Ocean, whale shark sightings occurred mostly off Gabon in equatorial

eastern Africa, and at around 10 °N between Senegal and Sierra Leone – these areas

comprise most of the fishing effort (Figure 18). Likewise in the western Pacific, more

sightings were recorded around the area where more days were spent fishing,

although sightings were relatively more spread out through the sampled area (Figure

18).

Page 149: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

151

Our spatial models (step 1) revealed higher habitat suitability mostly

corresponding to the areas where more whale sharks were spotted in the Atlantic, and

to the northeast of Indonesia/ Papua New Guinea in the western Pacific also covering

the area with more sightings (Figure 18 and Figure 19). The model with highest

statistical support was the same for both oceans in step 1 (wAICc = 0.83 and 0.33 in

the Atlantic and Pacific, respectively), and included all predictors except chlorophyll a.

This model explained 50 % of deviance in the Atlantic and only ~ 6 % in the Pacific

(Table 8). During the austral autumn, habitat was more suitable in the Pacific (habitat

suitability > 0.7) than in the Atlantic (~ 0.5) (Figure 19), however, κ was more than

double in the Atlantic (κ ~ 0.7: ‘good’ performance compared to κ ~ 0.3 in the Pacific:

‘poor’ performance) (Table 8). The mean prediction error calculated through the 10-

fold cross validation (CVe) was relatively low in both oceans (≤ 0.1).

Figure 19: Predicted habitat suitability of whale sharks (Rhincodon typus, Smith 1828) in the Atlantic (left) and Pacific (right) Oceans during the months of April to June. Values for habitat suitability vary from 0 (low) to 0.51 and 0.72 (high) in the Atlantic and Pacific, respectively.

Page 150: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

152

Tabl

e 8: S

umm

ary o

f spa

tial g

ener

alise

d lin

ear m

odels

(ste

p 1)

relat

ing

prob

abilit

y of w

hale

shar

k (Rh

inco

don

typu

s, Sm

ith 18

28) o

ccur

renc

e to

ocea

n pr

oper

ties.

Ocea

n pr

oper

ties:

slop

e, de

pth,

dep

th2 a

nd d

istan

ce to

shor

e (sh

ore)

- re

ferre

d to

toge

ther

as p

hysic

al va

riabl

es (p

hysv

ar),

mea

n se

a sur

face

tem

pera

ture

(mSS

T) an

d its

qu

adra

tic te

rm (S

ST2)

, sea

surfa

ce te

mpe

ratu

re st

anda

rd d

eviat

ion

- ref

erre

d to

toge

ther

as S

ST va

riabl

es (S

STva

r), an

d m

ean

chlo

roph

yll a

(mCh

l a) i

n th

e Atla

ntic

and

west

Pa

cific

Ocea

ns d

urin

g Ap

ril to

June

. We s

hown

the b

iased

-cor

rect

ed m

odel

evid

ence

bas

ed o

n we

ight

s of A

kaike

’s in

form

atio

n cr

iterio

n co

rrect

ed fo

r sm

all sa

mpl

e size

s (wA

ICc;

on

ly va

lues

≥ 0.

001 s

hown

), th

e per

cent

age o

f dev

iance

expl

ained

(%De

), an

d th

e Coh

en’s

Kapp

a sta

tistic

s (k;

onl

y valu

es ≥

0.1 s

hown

) for

each

mod

el pe

rform

ing

bette

r tha

n th

e nu

ll mod

el (in

cludi

ng o

nly a

n of

fset

for e

ffort

and

the s

patia

l ran

dom

effe

ct) i

n ea

ch o

cean

.

Atlan

tic O

cean

Pacif

ic Oc

ean

M

odel

wA

ICc

%De

κ

wAIC

c

%De

κ

ph

ysva

r + S

STva

r 0.8

3 50

.4 0.7

± 0.

05

0.33

6.4

0.3 ±

0.08

phys

var +

mSS

T +

SST2

0.17

50

.3 0.6

± 0.

06

0.33

6.1

0.3

± 0.

08

m

SST

+ SS

T2 + m

Chl a

-

- -

0.

28

5.4

0.3 ±

0.08

phys

var +

mCh

l a

- 49

.5 0.2

± 0.

02

0.03

5.4

0.3

± 0.

08

m

Chl a

-

47.0

-

0.01

4.3

0.2

± 0.

08

m

SST

+ SS

T2 -

47.4

0.1 ±

0.01

0.

004

4.3

0.3 ±

0.07

phys

var

- 48

.3 0.4

± 0.

05

0.00

4 4.7

0.3

± 0.

08

SS

T var

-

47.1

0.1 ±

0.01

0.

003

4.4

0.3 ±

0.08

shor

e -

- -

0.

001

3.8

0.2 ±

0.07

depth

-

- -

0.

001

3.6

0.2 ±

0.08

depth

2 -

49.5

-

- -

-

slo

pe

- 47

.3 -

-

- -

Page 151: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

153

Temporal trends in whale shark occurrence

The total number of annual whale sharks sighted in the Atlantic varied from 0 to 137

(33 ± 39; Figure 20a) between April and June. Sightings peaked (~ 120 individuals) in

1995, even though total fishing effort (sum of fishing days spent fishing in all 1-º grid

cells) remained relatively constant between 1985 and 2000 at about 6000 fishing days

(5.8 ± 2.6×103; Figure 20c). During this peak, the number of occurrences was similar

to the one recorded in the early 1980s (137 sightings) when effort was close to 10000

days (Figure 20). Between 2000 and 2010, effort dropped to ~ 2000 days in the

Atlantic. In contrast, fishing effort increased steadily in the Pacific from year 2000

(Figure 20c). The number of sightings in the latter ocean also increased with time,

from around 10 sightings per year in the early 2000s to a maximum of only 35 in 2010

(Figure 20a).

The temporal model (step 2) with highest statistical support differed between

oceans (Table 9). In the Atlantic, this model (wAICc = 0.633) included as predictors the

habitat suitability derived from step 1 (Hsuit), effort, and year (as a random intercept).

The models including the linear and quadratic terms for time (time and time2) resulted

in similar log-likelihood (Table 9), and explained similar percentage of deviance, but

received lower support from wAICc. In the Pacific, the highest-ranked model (wAICc =

0.472) also included the time predictor, although it only explained < 15 % of the

deviance. The partial effect of time on the probability of whale shark occurrence during

the months of April to June in the western Pacific is depicted in Figure 21b, and shows

an increase from 0.003 to 0.012 between 2000 and 2010 (Figure 21b).

Page 152: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

154

Figure 20: Yearly variation in (a) number of whale sharks sighted, (b) relative whale shark occurrences, and (c) days spent fishing. During the months of April to June in the Atlantic (1-º resolution), Indian (5-º resolution) (from Sequeira et al., in press; included here for comparison only), and western Pacific (5-º resolution) Oceans.

Page 153: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

155

Tabl

e 9: S

umm

ary o

f the

tem

pora

l gen

erali

sed

linea

r mixe

d-ef

fect

s mod

els (s

tep

2).

Relat

ing

prob

abilit

y of w

hale

shar

k occ

urre

nce i

n th

e Atla

ntic

and

Pacif

ic Oc

eans

(fro

m A

pril t

o Ju

ne) t

o: a

spat

ial p

redi

ctor

der

ived

from

the s

patia

l dist

ribut

ion

mod

els (H

suit)

, ef

fort

(tem

pora

l var

iatio

n in

fish

ing

effo

rt), t

ime (

year

s), a

nd a

rand

om ef

fect

for y

ear.

Show

n fo

r eac

h m

odel

are t

he n

umbe

r of p

aram

eter

s (k)

, log-

likeli

hood

(LL)

, bias

ed-c

orre

cted

m

odel

evid

ence

bas

ed o

n we

ight

s of A

kaike

’s in

form

atio

n cr

iterio

n co

rrect

ed fo

r sm

all sa

mpl

e size

s (wA

ICc)

and

the p

erce

ntag

e of d

evian

ce ex

plain

ed (%

De).

Top-

rank

ed m

odels

in

each

step

are i

n bo

ld an

d or

dere

d by

%De

. Mod

els w

ith w

AIC c

< 0.

05 n

ot sh

own.

A

tlant

ic Oc

ean

Pa

cific

Ocea

n

Mod

el

k

LL

wAIC

c %

De

LL

wAIC

c %

De

Hsui

t + ef

fort

+ tim

e + ti

me2 +

(1| y

ear)

7

-279

0.90

0.133

51

.70

-753

.55

0.247

14

.99

Hsu

it +

effo

rt +

time +

(1| y

ear)

6

-279

1.34

0.234

51

.69

-753

.91

0.472

14

.95

effo

rt +

time +

time2 +

(1| y

ear)

6

- -

- -7

55.67

0.0

81

14.75

e

ffort

+ tim

e + (1

| yea

r)

5 -

- -

-756

.07

0.148

14

.70

Hsu

it +

effo

rt +

(1| y

ear)

5

-279

1.34

0.633

51

.69

- -

-

Ta

ble 1

0: E

stim

ated

weig

ht o

f evid

ence

for e

ach

tem

pora

l pre

dict

or u

sed

in th

e gen

erali

sed

linea

r mixe

d-ef

fect

s mod

els: H

suit

- spa

tial p

redi

ctor

der

ived

from

the s

patia

l di

strib

utio

n m

odels

, effo

rt - t

empo

ral v

ariat

ion

in fi

shin

g ef

fort

and

a lin

ear a

nd q

uadr

atic

term

for t

ime (

year

s).

Ocea

n we

ight

of e

viden

ce

Hsui

t ef

fort

time

time2

Atlan

tic

1 0.9

99

0.367

0.1

33

Pacif

ic 0.7

61

1 0.9

53

0.328

Page 154: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

156

a) Indian Ocean (Sequeira et al., in press)

b) Pacific Ocean

Figure 21: Partial effect of time on the probability of whale shark presence during the months of April to June. (a) the western Indian Ocean (from Sequeira et al., in press; included here for comparison only); (b) the western Pacific Ocean according to the model with highest information-theoretic support: Presence ~ Hsuit + effort + time + (1| year) (Table 2). Dashed lines indicate 95 % confidence intervals. Note: the increasing trend in the Pacific Ocean starts from 2000, being concurrent with the declining segment previously observed in the Indian Ocean.

The weight of evidence we estimated for each temporal predictor is shown in

Table 10. Habitat suitability (Hsuit) and effort received the highest weight (1 and 0.999,

respectively) in the Atlantic Oceans, while in the Pacific we found the highest weight

for effort (weight of evidence = 1) followed by the linear term for time (time: 0.953).

Cyclic patterns in whale shark occurrence

The fast Fourier transform applied to the time series of sightings per unit effort in the

Atlantic (Figure 22a, left) revealed a prominent peak at 15 years (Figure 22a, centre),

both with yearly and monthly frequencies. This periodicity also received the highest

power when plotting the possible cycles present in the dataset (Figure 22a, right);

however, its relative magnitude was similar to other possible cycles present in the time

series. In the Pacific, even though some cycles were identified (Fig 5b), their power

was consistently low (< 1×10-3).

Page 155: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

157

a)

b)

Fi

gure

22: W

hale

shar

k sig

htin

gs p

er u

nit e

ffort

durin

g th

e mon

ths o

f Apr

il to

June

in th

e Atla

ntic

(top

row;

1981

- 20

10) a

nd P

acifi

c (bo

ttom

row;

2000

- 20

10).

Left

- var

iatio

n in

sigh

tings

per

uni

t effo

rt wi

th ti

me;

cent

re –

stro

nges

t fre

quen

cy o

bser

ved;

and

right

– cy

clic d

escr

iptio

n in

wha

le sh

ark s

ight

ings

per

uni

t effo

rt as

resu

lt of

the

Four

ier tr

ansf

orm

s.

Page 156: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

158

Discussion

The detrimental effects of a changing climate expected for some species (Walther et

al., 2002) has stimulated a recent boom in efforts to examine temporal patterns in

species occurrence (e.g., Gotelli et al., 2010). We have mapped habitat suitability for

whale sharks and quantified temporal variation in their occurrence both in the Atlantic

and Pacific Oceans for the first time. Our modelling results showed no linear (or

quadratic) temporal trends in the Atlantic, although there was some evidence for a

cyclic pattern, and a slight increase (linear trend) in the Pacific since 2000. The latter

result contrasts with the decreasing trend detected in the Indian Ocean during the last

decade (Sequeira et al., in press). Given these opposing results, the evidence for an

inter-decadal occurrence pattern in the Indian Ocean (Sequeira et al., in press), the

low genetic differentiation of whale sharks sampled in different oceans (Castro et al.,

2007; Schmidt et al., 2009), and the implicit notion that whale sharks travel across

oceans (Hueter et al., 2008; Rowat & Gore, 2007), we hypothesise that whale shark

occurrence might be asynchronously cyclical across oceans. Such a pattern would be

consistent with the inter-ocean migratory behaviour we have previously hypothesised

(Sequeira et al., 2013).

Despite genetic clues (Castro et al., 2007; Schmidt et al., 2009) being the only

evidence to date corroborating inter-ocean migration, these results could also have

arisen even if only a small number of individuals were moving between ocean basins

(Hartl & Clark, 1989). However, the number of whale sharks occurring in aggregation

locations within the predictable season is highly variable among years (Rowat et al.,

2009a; Wilson et al., 2001), and this variability holds for the number of whale sharks

sighted at the ocean-basin scale (Figure 20a and b). The opposing trends we detected

between the Indian and the western Pacific Oceans during the last decade might

Page 157: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

159

therefore reflect a migration of sharks between oceans. Similar distributional shifts,

promoting asynchrony in occurrences, have also been suggested for other filter-

feeding sharks (basking sharks Cetorhinus maximus) between the west coast of

Republic of Ireland and the Norwegian Sea (Sims & Reid, 2002), and were potentially

associated with spatial changes in foraging conditions. If whale sharks are migrating

between oceans (Sequeira et al., 2013), a cyclic pattern such as the one observed in

the Atlantic Ocean would explain both the interannual variation in whale sharks

numbers at the ocean-basin scale (Figure 20a and b) and the asynchrony in their

occurrences in the Indian and Pacific Oceans. Moreover, such large-group and multi-

year migrations would still be consistent with the low differentiation found in genetic

studies (Castro et al., 2007; Schmidt et al., 2009). These whale shark migrations could

be associated with changes in environmental conditions (Sleeman et al., 2010b;

Wilson et al., 2001), or with sex/age or reproduction-related behaviour (Ramírez-

Macías et al., 2007; Sequeira et al., 2013). Reproduction-associated multi-year cycles

have been observed in anadromous fish (e.g., sockeye salmon Oncorhynchus nerka)

(Dingle, 1996). Multi-year cycles have also been reported for fish catches in the

western Indian Ocean (Jury et al., 2010), and for other marine megafauna, such as the

decadal cycle in cetacean strandings in southeast Australia (Evans et al., 2005).

We found a stronger cycle with a rhythmicity of 15 years in the Atlantic Ocean

using the fast Fourier transform, which corresponds to half of the full dataset’s

temporal span, and is therefore close to the highest frequency that can be detected by

this method (Nyquist limit) (Moler, 2004). Analogous studies reporting decadal cycles

in marine species have used much longer datasets (~ 40 years; Evans et al., 2005),

and so obtaining longer time series for whale sharks in the Atlantic will be required to

reveal whether this cycle persists. The western Pacific dataset was only a third the

Page 158: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

160

span of the Atlantic dataset, and so we did not expect to reveal meaningful cycles.

Nonetheless, we detected possible cycles for the Pacific at 5.5 and 2.2 years (Figure

22), but with only weak differences in power among them (< 10-3). Our hypothesis of

cyclic occurrence among ocean basins could mean that the light increase in relative

abundance observed in the Pacific Ocean over the last 10 years might correspond to

only a segment of an inter-decadal cycle (> 15 years), but this would be undetectable

with such short dataset.

The results from the spatial models (step 1) agreed well with the data in the

Atlantic (κ ~ 0.7) where higher-resolution data (1-° grid cell) were available. In this

ocean, we predicted higher suitable habitat mostly close to shore around Gabon,

Congo and Equatorial Guinea, and between Cotê d’Ivoire and Mauritania. Due to the

paucity of studies on whale sharks in the Atlantic (excluding the Gulf of Mexico and

Caribbean Sea which were not covered in our dataset), we found no independent data

to validate these results. Higher suitability was also predicted further from shore at

around the Equator and between 15–20 °W (Figure 19, left panel). Whale sharks

occur close to this area in the Saint Peter Saint Paul archipelago, peaking in

occurrence at the end of June (Hazin et al., 2008).

In the western Pacific, the resulting habitat maps show mostly suitable habitat

within the area covered by the fisheries, but despite the higher suitability (up to 0.72),

model accuracy was poor (κ ~ 0.3). Possible reasons for the poor performance include

poorer resolution of the data or unmeasured environmental covariates that could be

better predictors, such as current or wind conditions (Wilson et al., 2001). During the

season we considered (April to June), the number of sightings in the western Pacific

(167) was about 6 times lower than in the Atlantic (1030; Figure 18), which affects the

performance of the western Pacific model. Because the data available for this ocean

Page 159: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

161

represent only partial coverage, access to the full dataset (giving more sightings in this

season), would likely improve model performance by refining the distinction between

presence and absence locations. Nevertheless, the available data are representative

of the total fisheries in the area (P. Williams pers. comm.), and so application of the

same models to other seasons when more sightings are recorded might improve

results. However to compare synchronous seasonal occurrences in the three major

oceans, data covering the months of April to June were required. Due to the lack of

independent studies reporting whale shark occurrence off Indonesia / Papua New

Guinea, we could not externally validate this model’s predictions.

We found a lack of support for the linear or quadratic time predictor in the

Atlantic Ocean from April to June (Table 9) where habitat suitability (Hsuit) and effort

received the highest weight of evidence (Table 10). At least for the last decade in the

western Pacific, there was evidence for a linear, positive trend in the probability of

whale shark occurrence (Table 9; Figure 21). In the latter ocean, even though habitat

suitability was also ranked as an important variable (weight of evidence = 0.761), we

found the highest weight of evidence for the predictors effort and time (Table 10).

Despite the modest trend we found (from 0.003 to 0.012), the difference between

maximum and minimum values (~0.009) is similar to the prevalence assumed in our

temporal models (0.01, corresponding to 100 absences to each presence). Although

the Pacific positive trend is concurrent with the increase observed in fishing days, we

accounted for both spatial and temporal effort bias. Because the habitat suitability

outputs from the spatial model were inputs for the temporal model, the increase in the

whale shark probability of occurrence in the Pacific should be interpreted with caution.

However, if this increase is real, the trend is opposite to that identified in the Indian

Ocean over the same interval (Sequeira et al., in press) (Figure 21), and could indicate

Page 160: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

162

a quasi-decadal shift in sharks between oceans as referred above. A slight lag in

occurrences between the Indian and Atlantic oceans is also evident in some years.

Even if whale shark occurrence is cyclic, which could explain some of the

declines observed in some years / locations, the total number of sightings reported in

the fisheries datasets (ocean-scale) is also lower in the last decade (< 50 per year in

all oceans; Figure 20a). In the 1990s, there were a total of about 500 sightings in the

Atlantic Ocean, and in the early 2000s, only around 150 (both in the Atlantic and west

Pacific) during the study season. We observed a similar pattern in the Indian Ocean

(Sequeira et al., in press), with 600 sightings reported in the 1990s and ~ 200 in

following years (Figure 20a). The reduced sightings in the last decade agree with the

recent accounts of declines from near-shore aggregations (Bradshaw et al., 2008).

The datasets we used were opportunistically collected and therefore lead to

inherent complications. Differential sampling effort is a major issue for which we

accounted at least partially in both modelling steps to reduce potential bias in spatial

and temporal effort. However, an underlying assumption of our models is that failure to

report a whale shark presence (whether due to a failure in detecting a shark or in

reporting a detection) is evenly distributed across the sampling area and study period.

Access to higher-resolution data would likely improve the results (exemplified by the

Atlantic Ocean models), and can be achieved through establishing collaboration

between researchers and fisheries-management organisations. Commercial-in-

confidence restrictions on access to fisheries data (especially sensitive data such as

fish catch) delay scientific research from providing more accurate models on fisheries

impacts. Such collaborations would also allow researchers to give greater insight on

how better datasets could be collated for specific objectives using the resources

currently in place. However, while data collected by fisheries might not be as precise

Page 161: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

163

as those derived from scientifically designed surveys, they are still essential sources of

information given the logistic challenges of surveying widely distributed species over

their entire range. Fisheries data therefore present unique opportunities to advance

our understanding of such species.

Because the datasets we used contained only recorded presences, the

generation of pseudo-absences was necessary to allow for the binomial estimation in

the models. Differences associated with the method to generate pseudo-absences

was assessed in a previous study using similar datasets (Sequeira et al., 2012) where

we found that model performance was not reduced by using random pseudo-absences

selection, and was therefore the method used in this analysis. This accords with the

results from another recent publication specifically addressing pseudo-absence

selection issues (Barbet-Massin et al., 2012). The number of pseudo-absences

selected is another issue to consider when generated. Barbet-Massin et al. (2012)

reported that model accuracy increases until the presence to pseudo-absence ratio

(i.e., prevalence) equals 0.1, and remains constant for lower ratios. In the models we

developed, prevalence was 0.1 in the spatial assessment, and 0.01 in the temporal

assessment (to allow enough points for the spatially explicit temporal analysis);

therefore in both cases, we do not expect the ratio presences:pseudo-absences to

have affected model accuracy.

Another important aspect of the input data we used is that both sightings and

environmental variables correspond mostly to the ocean surface layer. Three-

dimensional environmental data at an adequate resolution for regional analyses such

as that presented here are currently not available. Even though whale sharks spend

most time at the surface (e.g., Rowat et al., 2007) they also dive frequently, and

therefore, assessment of how they explore the vertical habitat and estimation of its

Page 162: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

164

suitability at a broad scale is also of importance. However, this will only be possible

when new technologies are developed to allow vertical data collection at broad spatial

scales.

Despite the limitations associated with both the opportunistic dataset and

species distribution models, we have demonstrated a powerful way to use such

datasets to identify both spatial and temporal trends of highly migratory species. We

also revealed that longer-term and higher-resolution sightings datasets are still

required to tease apart potentially confounding aspects of inter-ocean migration in

whale sharks. Continued investigation of these connections is necessary to assess

temporal trends in occurrence, and to reveal the suspected impacts of human

modifications to the wider marine realm.

Page 163: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

165

CHAPTER VI. Predicting current and future global distributions of

whale sharks

A. Sequeiraa, C. Mellina,b, D. Fordhama, M. Meekanc, C. Bradshawa,d

aThe Environment Institute and School of Earth and Environmental Sciences, The

University of Adelaide, South Australia 5005, Australia

bAustralian Institute of Marine Science, PMB No. 3, Townsville MC, Townsville,

Queensland 4810, Australia

cAustralian Institute of Marine Science, UWA Oceans Institute (MO96), 35 Stirling

Hwy, Crawley, Western Australia 6009, Australia.

dSouth Australian Research and Development Institute P.O. Box 120, Henley Beach,

South Australia 5022, Australia

(Manuscript in preparation)

Page 164: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

166

Page 165: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

167

Statement of authorship

PREDICTING CURRENT AND FUTURE GLOBAL DISTRIBUTION OF WHALE

SHARKS

Text in manuscript SEQUEIRA, A.M.M. (Candidate) Planned the article, performed the analysis, interpreted data, developed the models, wrote manuscript and acted as corresponding author. I hereby certify that the statement of contribution is accurate.

Signed……………………………………………………………………….Date……………

MELLIN, C. Helped with data interpretation, models development and assisted writting. I hereby certify that the statement of contribution is accurate and I give permission for the inclusion of the manuscript in the thesis.

Signed……………………………………………………………………….Date……………

FORDHAM, D. Assisted with climate change models development and incorporation of results. I hereby certify that the statement of contribution is accurate and I give permission for the inclusion of the manuscript in the thesis.

Signed……………………………………………………………………….Date……………

Page 166: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

168

MEEKAN, M. G.

Supervised development of work and provided evaluation. I hereby certify that the statement of contribution is accurate and I give permission for the inclusion of the manuscript in the thesis.

Signed……………………………………………………………………….Date……………

BRADSHAW, C.J.A. Supervised development of work, helped in model results interpretation and manuscript evaluation. I hereby certify that the statement of contribution is accurate and I give permission for the inclusion of the manuscript in the thesis.

Signed…………………………………………………………..………….Date……………

Page 167: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

169

Predicting current and future global distributions of whale sharks

Abstract

The highly migratory whale shark, spanning the warm and temperate waters around

the globe, is classed as Vulnerable by the IUCN. However, no one has yet predicted

their present-day or possible future global distribution. We used 30 years (1980-2010)

of whale shark observations recorded by tuna purse-seiners fishing in the Atlantic,

Indian and Pacific Oceans, and applied generalised linear mixed-effects models to test

the hypothesis that similar environmental covariates predict whale shark occurrence in

all major ocean basins. We derived global data for potential predictors from satellite

images for chlorophyll a (an index of primary productivity) and sea surface

temperature, and bathymetric charts for depth, bottom slope and distance to shore.

We randomly generated a set of pseudo-absences within the total area covered by the

fisheries, and included fishing effort as an offset to account for potential sampling bias.

We also predicted sea surface temperatures for 2070 using an ensemble of 5 global

climate models under a no-climate-policy reference scenario, and used these

temperatures to predict changes in whale shark distribution. The full model for all

oceans (excluding standard deviation of sea surface temperature) had the highest

relative statistical support (wAICc = 0.99) and explained ~ 60 % of the deviance in the

presence-absence data. Whale shark habitat suitability was mainly driven by spatial

variation in bathymetry and sea surface temperature among oceans, although we

found evidence for the effects of these variables to differ slightly among oceans.

Changes in temperature to 2070 resulted in a slight shift of suitable habitat towards the

poles (~ 5 ºN in the Atlantic Ocean and 3-8 ºS in the Indian Ocean) accompanied by

an overall range contraction (~ 6%). Assuming that whale shark environmental

requirements, and human disturbances (i.e., in a scenario of no-stabilization of

Page 168: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

170

greenhouse gas emissions) remain similar in the near future, we show that changes

derived from warming oceans might promote a net retreat from current aggregation

areas and an overall redistribution of the species.

KEYWORDS Rhincodon typus; Indian Ocean; Atlantic Ocean; Pacific Ocean; remote

sensing; sea surface temperature; tuna purse-seine fisheries; species distribution

models; global warming; climate change; conservation

Introduction

Changes in climate are expected to alter the current distribution of species (Parmesan,

2006; Thomas et al., 2004; Wernberg et al., 2011). Despite a bias of studies towards

terrestrial habitats (Richardson & Poloczanska, 2008), there is mounting evidence for

distributional shifts in the marine environment (e.g., Edwards & Richardson, 2004;

Hiddink & ter Hofstede, 2008), which generally result in pole-ward latitudinal (e.g.,

Perry et al., 2005) or depth (e.g., Dulvy et al., 2008) shifts. These shifts have been

reported in species as divergent as plankton (Beaugrand et al., 2009; Southward et al.,

1995), mussels (Berge et al., 2005), capelin (Rose, 2005), demersal fish (Beaugrand &

Kirby, 2010); (Dulvy et al., 2008; Perry et al., 2005) and macroalgae (Wernberg et al.,

2011). However, assessing how much the distribution of species will be affected by

climate change implies that the current distribution of the species is adequately

identified.

A species’ distribution and its habitat requirements can be estimated by

modelling information on occurrence together with environmental correlates (Kearney

& Porter, 2009). Despite the diverse use of species distribution models in terrestrial

Page 169: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

171

environments, until recently only a handful of these models had been developed for

the marine environment (Elith & Leathwick, 2009; Robinson et al., 2011), but now they

have been applied to many marine species such as corals (e.g., Beger & Possingham,

2008; Garza-Pérez et al., 2004), molluscs (Mellin et al., 2011), reef fish (Mellin et al.,

2010), algae (e.g., Kelly et al., 2001; Sérgio et al., 2007), mammals (e.g., Oviedo &

Solís, 2008; Praca & Gannier, 2007) and sharks (McKinney et al., 2012; Sequeira et

al., in press; Sequeira et al., 2012). However, applying such models to highly migratory

species is particularly challenging (e.g., Kreakie et al., 2012), and the problem is

exacerbated in the marine environment mostly due to poor detection (Elith &

Leathwick, 2009), lack of recorded occurrences in the open ocean (McKinney et al.,

2012), and the large geographical range of the species under consideration (Kaschner

et al., 2006; Sequeira et al., in press; Sequeira et al., 2012).

The whale shark (Rhincodon typus Smith 1828), the largest of all fish (> 18 m;

Borrell et al., 2011; Chen et al., 1997; Compagno, 2001), has a circumglobal

geographical range between 30 ºN and 35 ºS (Compagno, 2001; Last & Stevens,

2009). This range has been defined based on occasional occurrences (Compagno,

2001), but most importantly, the range of temperatures where the species is expected

to occur: tropical/ warm temperate (Compagno, 2001; Last & Stevens, 2009). Being

ectotherms, ambient temperatures directly influence their metabolic processes (Sims,

2003). Whale sharks occur regularly at the ocean surface (Gunn et al., 1999; Rowat et

al., 2007; Thums et al., 2012) in association with sea surface temperatures (Sequeira

et al., 2012) possibly to assist thermoregulation (Thums et al., 2012). They can also

dive deeply, opening speculation on this species’ potential to modify its diving

behaviour with increasing surface water temperatures associated with climate change

(Hoegh-Guldberg & Bruno, 2010), and use deeper, cooler waters more frequently.

Page 170: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

172

However, oxygen limitation might pose a problem for extended time spent in deeper

waters (Graham et al., 2006; Pörtner, 2001; Prince & Goodyear, 2006) so for this

reason, expected warming will likely affect the spatial distribution of whale sharks

possibly by promoting shifts in their present geographical range. Moreover, because

whale sharks are filter-feeders, these shifts might follow the simultaneous

redistribution of plankton (Beaugrand et al., 2009).

Whale sharks have been regularly recorded in locations near shore (Sequeira et

al., 2013) to where they attract substantial tourism interest due to their innocuous

behaviour and large size (e.g., Cárdenas-Torres et al., 2007; Rowat & Engelhardt,

2007). Concerns about population declines, possibly driven by illegal direct catches

elsewhere (Riley et al., 2009) or habitat disturbance, such as derived from tourism

activities (Heyman et al., 2010), have prompted a research focus on documenting

migratory pathways to determine whether partial-range protection is sufficient to

ensure persistence locations (Bradshaw et al., 2008; Rowat, 2007; Sequeira et al.,

2013). In addition to these direct threats, climate change might also already be

affecting whale sharks. However, the species’ high mobility and the lack of extensive

occurrence data mean that teasing apart spatial and temporal patterns is problematic

(Sequeira et al., in press). Thus, determining whether departures from current

distributional limits (Duffy, 2002; Rodrigues et al., 2012; Sa, 2008; Turnbull & Randell,

2006) or temporal patterns in relative abundance (e.g., Bradshaw et al., 2008;

Theberge & Dearden, 2006) are linked to changing temperatures (e.g., Sleeman et al.,

2007) is equally difficult.

Fortunately, statistical tools exist to facilitate inference in this regard. For

example, known occurrences can be used to define the current distribution of a

species, and future distributions can be projected by coupling climate forecasts

Page 171: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

173

derived from global circulation models (e.g., Araújo et al., 2005; Cheung et al., 2010).

Despite the known limitations of these models, such as biases derived from the input

data or from the geographical extent available (Barbet-Massin et al., 2012; Barbet-

Massin et al., 2010; Phillips et al., 2009), and uncertainties derived from the models

used (both species distribution models and global circulation models) (Buisson et al.,

2010), confidence in their application can be improved by using an ensemble result

(Araújo & New, 2006). This ensemble result takes into account variability among

models used (Hijmans & Graham, 2006; Kearney et al., 2010; Knutti et al., 2010;

Lawler et al., 2006). Another important limitation, especially from a conservation

perspective, is the coarse resolution of predictions derived from global circulation

models, and their consequently limited biological relevance (Hannah et al., 2002; Seo

et al., 2009). However, this can be improved by down-scaling the forecasts, coupling

them to ecological processes (Araújo et al., 2005; Fordham et al., 2011; Fordham et

al., 2012; Hannah et al., 2002).

For the first time, we used a global dataset of pelagic whale shark sightings

derived from the tuna purse-seine fishery operating in the Atlantic, Indian and Pacific

Oceans to: (1) predict the spatial distribution of whale shark habitat suitability across

three ocean basins, and (2) extend our projections to a future scenario of ocean

warming based on a no-climate-policy reference (no stabilisation of greenhouse gas

emissions). Following our previous work where we derived ocean-basin scale

predictions for whale shark distribution and temporal trends (Sequeira et al., 2012;

Sequeira et al., 2013), herein our overarching aim was to derive realistic predictions of

the species current geographical range (circumglobal) and provide a modelling tool for

exploring the species’ future distribution under anticipated climate change.

Page 172: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

174

Methods

Data

We obtained whale shark presence data (henceforth, ‘sightings’) from the logbooks of

tuna purse-seine fisheries, where whale sharks associated with net sets (Matsunaga

et al., 2003) were recorded. Data were provided by the Institut de Recherche pour le

Développement (France) for the Atlantic and Indian Oceans, and by the Secretariat of

the Pacific Community for the western Pacific Ocean (Table 11 and Figure 23).

The full dataset spans the three major oceans with a maximum temporal

coverage of 30 years for the Atlantic Ocean (1980 – 2010) from 21 °N to 15 °S and 34

°W to 14 °E, 17 years (1991 – 2007) in the Indian Ocean from 30 °N to 30 °S and 35

to 100 °E and 11 years (2000 – 2010) in the western Pacific Ocean from 15 °N to 15

°S and 130 °E to 150 °W (Table 11 and Figure 23). Data included (i) the number of

whale shark sightings (Table 11 and Figure 23a) and location (longitude and latitude at

0.01° precision) and (ii) fishing effort (days; Figure 23b) per month and per year

pooled for each 5-° grid cell both in the Indian (6618 records) and western Pacific

(2272 records) Oceans, and for both 5-° and 1-° grid cells in the Atlantic Ocean

(maximum of 18277 records). No information was made available on vessel or trip

units for any ocean. The data from the western Pacific Ocean, even though

representative of total tuna purse-seine fisheries in the area (P. Williams pers. comm.),

were incomplete as data access was not granted from all fleets registered with the

Secretariat of the Pacific Community.

Because whale shark numbers fluctuate seasonally (Table 11), but not

necessarily in a spatially consistent manner (Sequeira et al., in press), we used the

season in which the highest number of whale sharks was recorded for each ocean. A

seasonal maximum of 314 whale sharks was recorded in the western Pacific between

Page 173: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

175

January and March, 811 in the Indian Ocean between April and June, and 1153

between July and September in the Atlantic Ocean.

Binary data (presence and absence) are required for binomial estimation in

spatial generalised linear mixed-effect models (GLMM); therefore, we generated

pseudo-absences (10:1 presence ratio, following Barbet-Massin et al., 2012) based on

a random spatial distribution within the area covered by the fisheries (excluding

presence locations).We used the srswor function (simple random sampling without

replacement) from the {Sampling} package in the R programming language (R

Development Core Team, 2012), and generated random pseudo-absences among

non-presence cells in each ocean within the same season for which the presences

were being considered. To assess the influence of the chosen pseudo-absences, we

generated each set 10 times prior to their inclusion in the spatial GLMM (described

below).

We collated an environmental dataset at a 9-km resolution composed of six

physical variables (Physvar): (1) distance to shore (shore; km) calculated with the Near

tool in ArcGIS 9.3.1TM using a world equidistant cylindrical coordinate system, (2)

mean depth (depth; m) and (3) slope (slope; º) derived from the one-minute grid of the

General Bathymetry Chart of the Oceans (GEBCO, 2003), (4) mean and (5) standard

deviation of sea surface temperature (in °C) (SST, SSTsd or together SSTvar) derived

from daytime measures from the Advanced Very High Resolution Radiometer

(AVHRR) PathFinder version 5.0 and (6) mean concentration of chlorophyll a (Chl a in

mg m-3) derived from the Sea-viewing Wide Field-of-view Sensor (SeaWiFS). We

calculated averages and standard deviations of chlorophyll a and sea surface

temperature based on weekly satellite measures available within the total period

covered by the fisheries in each ocean using ArcToolBox functions from ArcGIS

Page 174: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

176

9.3.1TM automated with Python scripts. The environmental dataset for each ocean

matched the ocean-specific season of maximum whale shark abundance.

Page 175: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

177

Tabl

e 11:

Num

ber o

f wha

le sh

ark s

ight

ings

in ea

ch o

cean

per

dec

ade a

s rec

orde

d by

tuna

pur

se-s

eine f

isher

ies.

Sour

ce: I

ndian

Oce

an T

una C

omm

issio

n an

d th

e Ins

titut

de R

eche

rche

pou

r le D

évelo

ppem

ent,

Fran

ce fo

r the

Atla

ntic

[Atl]

and

Indi

an [I

nd],

and

Secr

etar

iat o

f the

Pac

ific C

omm

unity

for t

he w

este

rn

Pacif

ic [P

ac].

See a

lso F

igur

e 23.

Num

bers

of s

ight

ings

are s

plit

by tr

imes

ter:

Jan

– Mar

, Apr

– Ju

n, Ju

l – S

ep an

d Oc

t – D

ec).

Bold

face

indi

cate

s dat

a use

d to

fit m

odels

. Num

bers

for t

he w

este

rn

Pacif

ic ar

e der

ived

from

obs

erve

r dat

a onl

y.

Ja

n - M

ar

Apr –

Jun

Jul -

Sep

Oc

t - D

ec

Tota

ls

Tim

e per

iod

Atl

Ind

Pac

Atl

Ind

Pac

Atl

Ind

Pac

Atl

Ind

Pac

Atl

Ind

Pac

1980

-199

0 2

- -

405

- -

310

- -

17

- -

734

- -

1991

-200

0 2

92

- 51

8 59

9 -

543

40

- 12

36

-

1075

76

7 -

2000

-201

0 3

23*

314

147

212*

16

7 30

0 27

* 18

9 38

15

5*

185

488

417*

85

5

Tota

ls

7 11

5 31

4 10

70

811

167

1153

67

18

9 67

19

1 18

5 22

97

1184

85

5

* Ind

ian O

cean

data

for th

e las

t dec

ade o

nly fo

r the

perio

d 200

0 – 2

007.

Page 176: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

178

Page 177: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

179

Figu

re 23

: Fish

eries

dat

a der

ived

from

tuna

pur

se-s

eine f

isher

y log

book

s in

the A

tlant

ic, In

dian

and

west

ern

Pacif

ic Oc

eans

. (a

) Sig

htin

gs fr

om 19

80 to

2010

: 433

6 wha

le sh

arks

reco

rds,

with

53 %

of s

ight

ings

in th

e Atla

ntic,

27 %

in th

e Ind

ian fr

om 19

91 to

2007

, and

20 %

in th

e wes

tern

Pac

ific f

rom

2000

to 20

10. S

hade

d ar

ea sh

ows w

hale

shar

k geo

grap

hica

l ran

ge (W

S ra

nge)

def

ined

by t

he In

tern

atio

nal U

nion

for C

onse

rvat

ion

of N

atur

e, an

d th

ick, g

rey l

ines

mar

k cur

rent

latit

udin

al ra

nge (

Com

pagn

o, 20

01).

(b

) Tun

a pur

se-s

eine f

ishin

g ef

fort

for t

he A

tlant

ic, In

dian

and

west

ern

Pacif

ic Oc

eans

(the

com

mon

reso

lutio

n of

5 º w

as u

sed

for e

asier

com

paris

on).

Data

from

the e

aste

rn P

acifi

c (wi

thin

inse

t) co

nsist

ed o

f loc

atio

ns w

ithou

t inf

orm

atio

n on

year

or s

easo

n of

sigh

ting

, but

with

no

asso

ciate

d te

mpo

ral in

form

atio

n (N

A in

Map

b) a

nd w

ere t

hus n

ot in

clude

d in

the a

nalys

is.

Page 178: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

180

Generalised linear mixed-effects models

We applied generalised linear mixed-effects models with a binomial error distribution

and a logit link function to the full, three-ocean dataset. To account for spatial bias in

sampling effort without reducing the spatial resolution of the dataset, we included this

variable as an offset, and used the highest resolution available in each ocean. The

mixed-effects models included several possible combinations of the fixed

(environmental predictors – see Table 12) and a spatial random effect (1-° grid cells)

to remove spatial autocorrelation (Sequeira et al., 2012). We also included the

interaction terms ocean x depth, ocean x slope and ocean x temperature, to test the

hypothesis that these predictors had similar predictive power in every ocean. To

account for non-linear dependencies of occurrence with both bathymetric features and

sea surface temperature, we included depth and mean sea surface temperature as a

second-order polynomial function using poly in the {stats} package in R.

We used Akaike’s information criterion corrected for small sample sizes (AICc)

(Burnham & Anderson, 2004) and their weights (wAICc) to compare model

probabilities. We used percentage of deviance explained (%De) to quantify goodness-

of-fit, and Cohen’s Kappa statistics (κ) to assess the model’s predictive power (Cohen,

1960). We calculated mean prediction error for the model with highest support using a

10-fold cross-validation (Davison & Hinkley, 1997).

Page 179: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

181

Table 12: Summary of generalised linear mixed-effects models relating probability of whale shark occurrence to ocean properties in the Atlantic, Indian and western Pacific during the season of peak occurrences for each ocean. Slope, distance to shore (shore), depth and its quadratic term referred to together as physical variables (Physvar); mean sea surface temperature (SST), its quadratic term (SST2) and standard deviation referred to together as SST variables (SSTvar). Average chlorophyll a is represented as Chl a. Shown for each model are biased-corrected model probabilities based on weights of Akaike’s information criterion corrected for small sample sizes (wAICc, only > 0.0001 shown), percentage of deviance explained (%De), 10-fold cross validation error (CVerror) and kappa statistics. Note: All models included an offset term for effort and a spatial random effect (1-º grid cell). An interaction term between ocean and depth, slope or SST was also included whenever these predictors were present.

The GLMM (with a logit link function) can be expressed as: , where Presence is the expected mean probability of

sightings occurrence. and represent the fixed and random covariates used in the models: the environmental predictors and the spatial grid, respectively. and represent the coefficients associated with the fixed and random effects, respectively, and �� is the intercept. The log of fishing effort was included as an offset. The index i corresponds to the number of observations among grid-cells.

Model wAICc %De CVerror kappa Physvar + SST + SST2 0.99 57.9 0.09 0.64 ± 0.03 Physvar + SSTvar - 57.8 0.09 0.64 ± 0.03 Physvar + Chl a - 57.3 0.09 0.64 ± 0.03 Physvar - 56.9 0.09 0.59 ± 0.03 depth + depth2 - 56.8 0.12 0.58 ± 0.03 SST + SST2 - 56.5 0.09 0.57 ± 0.03 SST + SST2 + Chl a - 56.5 0.09 0.55 ± 0.04 SSTvar - 56.5 0.09 0.54 ± 0.04 depth - 55.9 0.09 0.22 ± 0.04 slope - 55.9 0.09 0.30 ± 0.04 Chl a - 50.8 0.11 0.33 ± 0.03 Shore - 48.6 0.09 0.07 ± 0.03

Page 180: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

182

To account for variability in habitat suitability resulting from different modelling

approaches, we have also run some additional analysis using the BIOMOD modelling

platform (Thuiller, 2003). However, being a model-comparison framework (rather than

hypothesis-based) BIOMOD does not allow for direct input on the contribution of each

variable. Details are presented in supplementary information (Appendix D).

Global projection

We compiled a worldwide environmental dataset as described above and including the

same predictors. Here, we used seasonal averages for sea surface temperature and

chlorophyll a for the last decade (2000 – 2012) by using data derived from the

Moderate Imaging Spectroradiometer (MODIS) Aqua available online (seasonal

climatology maps; http://oceancolor.gsfc.nasa.gov/cgi/l3). After fitting the GLMM with

the tuna purse-seine data and running predictions for the entire area covered by the

fisheries, we used a model-averaging approach (based on wAICc) to project habitat

suitability by including the worldwide dataset under current (2000 – 2012) conditions.

As a qualitative validation, we overlaid the currently known locations for whale shark

seasonal occurrence (Sequeira et al., 2013) onto these global habitat suitability maps.

Climate change projections

Using the free software package MAGICC/SCENGEN5.3 (a model for the assessment

of greenhouse-gas induced climate change and a regional climate scenario generator;

for details refer to www.cgd.ucar.edu/cas/wigley/magicc/), we generated monthly

forecasts of expected change in sea surface temperature in 2070 (relative to 1995)

under a no-climate-policy reference scenario (no stabilization of greenhouse gas

Page 181: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

183

emissions) (cf. MiniCAM Ref. in Clarke et al., 2007). We made forecasts using five

atmospheric-ocean general circulation models chosen based on their skill (Fordham et

al., 2011) in predicting global sea surface temperatures for the baseline period 1981 –

2000. The chosen models were CCSM-3, MRI-CGCM2.3.2, ECHAM5/MPI-OM,

MIROC3.2 (hires) and UKMO-HadCM3 (terminology follows CMIP3 Multi-Model

Dataset Archive naming convention; www-pcmdi.llnl.gov/ipcc/about_ipcc.php). We

used bilinear interpolation of the GCM data (2.5-º grid cells) to a finer resolution of 0.5-

º longitude/ latitude to reduce discontinuities in the perturbed climate at the GCM grid-

box boundaries (Fordham et al., 2011). We calculated an average change expected

per grid cell across months for each quarter of the year (Jan – Mar, Apr – Jun, Jul –

Sep and Oct – Dec). We then downscaled the predicted climate anomalies (multi-

model average) using the ‘change factor’ empirical method, where the low-resolution

climate anomaly is added directly to a high resolution (9-km grid cells) observed sea

surface temperature baseline. In our case, this high resolution baseline derived from

the MODIS Aqua satellite composites for 1995 (the mid-year in our time series from

1980 – 2010).

We used these forecasts of sea surface temperature to generate predictions of

worldwide whale shark habitat suitability using our species distribution models. We

considered the range of predictor values used to fit the model, and masked areas

where model predictions occurred outside the environmental space used for the

original model fit.

Page 182: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

184

Results

Spatial patterns of occurrence and sampling effort

Whale shark sightings in the Atlantic and Pacific Oceans occurred mostly in the areas

where sampling effort was highest (Figure 23), whereas most records in the Indian

Ocean fell within the Mozambique Channel, which had lower sampling effort than in

the central-west area (Figure 23b). There was generally higher effort in the Pacific

Ocean (Figure 23b; > 40 000 fishing days in some 5-º cells), even though the Pacific

time series covered only the last decade (i.e., only one third of the duration of the

Atlantic Ocean dataset; Table 11). Within the area covered by the fisheries, mean

depth in whale shark sightings locations was similar in all three oceans (~ 3800 m;

Figure 24), but the Pacific bathymetric range was higher and average distances to

shore were slightly shorter (~ 265 km), and in the Indian Ocean average distance to

shore was greater (~ 435 km). Sea surface temperature varied among the three

oceans (Figure 24) with cooler temperatures in the Atlantic ranging from 19.4 to 28.3

ºC (± 0.31 – 7.2; Jul – Sep), between 25.2 and 31.2 ºC (± 0.39 – 2.57, Apr - Jun) in

the Indian, and between 26.7 and 30.1 ºC (± 0.19 - 1.36; Jan - Mar) in the western

Pacific, with the latter having the lowest standard deviation. Chlorophyll a values

averaged 0.75, 0.18 and 0.11 mg m-3 in the Atlantic, Indian and Pacific, respectively.

Page 183: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

185

Figure 24: Boxplots for major environmental predictors within the area covered by the fisheries in each ocean. Upper and lower limits indicate the maximum and minimum values in each dataset.

Habitat suitability prediction

The percentage of deviance explained was highest (58 %) for the model including all

physical and temperature variables (except standard deviation for sea surface

temperature), a spatial random effect (1-º grid cell) and an offset for effort (Table 12).

Statistical support (wAICc) was highest for the same model (0.99; Table 12). As a

result of the interaction terms (sea surface temperature, slope and depth interactions

with ocean) within the model with the highest statistical support, we found evidence for

a different effect of temperature and depth in each ocean, with both predictors

Page 184: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

186

affecting whale shark occurrence in the Atlantic, only temperature in the Indian, and

only depth in the Pacific Ocean.

The resulting prediction maps (Figure 25) show suitable habitat in different

areas in the Atlantic from July to September, with higher suitability (> 0.8) around

Gabon, Congo and Equatorial Guinea. The model also predicted high suitability (> 0.6)

around Côte d’Ivoire, Ghana and Bénin, Cape Verde and Mauritania. In the Indian

Ocean, the area around the Mozambique Channel and close to shore in the south-

eastern side of the African continent had the highest suitability (~ 0.3) during the

months of April to June. In the western Pacific, the central portion of the total area

covered by the fisheries had the highest suitability for the first trimester of the year, but

predicted suitability there was the lowest of all three oceans examined (~ 0.1). Similar

areas with higher habitat suitability were obtained with BIOMOD (Figure D2,

Spearman �����; logit transformed predcitions).

Figure 25: Predicted whale shark habitat suitability within the area covered by tuna purse-seine fisheries. Representation only for the peak whale shark occurrence seasons: Jul – Sep in the Atlantic, Apr – Jun in the Indian, and Jan – Mar in the western Pacific.

Page 185: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

187

World projection of whale shark habitat suitability

Only the area between ~ 40 ºN and ~ 35 ºS (Figure 26) had environmental variables

within the range used to fit the GLMM. Exceptions occur in some seasons for areas

around the west coast of South America and southwest Africa (Figure 26). In the

Atlantic Ocean, highest habitat suitability was patchily distributed in the north (e.g.,

Gulf of Mexico, Mediterranean Sea, North West coast of Africa) and south (between

South America and Africa, but not farther south than 35 ºS). In the Indian Ocean,

areas with highest habitat suitability were mostly near shore; for example: in the Bay of

Bengal, Arabian Sea, Mozambique Channel and off Ningaloo Reef (Western

Australia). In the western Pacific, highest habitat suitability (albeit lower than in the

other oceans) occurred throughout the year (mostly in the central area covered by the

tuna purse-seine fisheries).

Projections of future whale shark distribution

The average anomaly forecast for each season in 2070 resulted in an increase of

around 0.8 to 2 ºC (Figure D1 in Appendix D), mostly within the region between 30 ºN

and 35 ºS (commonly considered the whale shark’s current latitudinal range).

Forecasting whale shark suitability based on the 2070 sea surface temperature

scenario resulted in a weak but evident pole-ward shift of habitat (Figure 27 and

compare with Figure 26), mostly in the Atlantic and Indian Oceans. We also observed

a contraction in habitat suitability in 2070 resulting in a loss of 5.0 % of suitable area

(in 9-km cells, as per model resolution) in Jan-Mar, 3.3 % in Apr-Jun, 4.7 % in Jul-Sep

and 6.1 % in Oct-Dec (Figure 27). In the Pacific, the projected increase in temperature

resulted in similar habitat suitability (still < 0.1) within roughly the same region,

although a ‘corridor’ of suitable habitat linking the western to the eastern Pacific Ocean

was strengthened for all seasons (Figure 27).

Page 186: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

188

Figure 26: Global predictions of current seasonal habitat suitability for whale sharks. Prediction maps generated from generalised linear mixed-effects models fit with the sightings and effort data collected by tuna purse-seine fisheries. Where environmental inputs fell outside the environmental space used for the original statistical fit results are shown as “out of range” in the map. �� indicates known aggregation locations within the seasons represented in each map (symbol size proportional to relative size of aggregation). Areas where some environmental predictors were not available (e.g., due to cloud cover) are shown in white (no result). To aid visualization, black line delineates areas where habitat suitability > 0.1 was predicted.

Page 187: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

189

Figure 27: Predicted shift in global of whale sharks habitat suitability for 2070 under a no-climate-policy reference (no greenhouse gas emissions) scenario. Black line delineates areas where higher habitat suitability (> 0.1) had been predicted under current environmental conditions (Figure 26), for visual assessment of the habitat contraction (reduction in the number of suitable 9-km cells of 5.0 % in Jan-Mar, 3.3 % in Apr-Jun, 4.7 % in Jul-Sep and 6.1 % in Oct-Dec) and pole-ward shift (~ 5 ºN in Jan-Sep in the Atlantic Ocean, and 3-8 ºS in the Indian Ocean). ��indicates known aggregation locations within the seasons represented in each map, with symbol size proportional to relative size of aggregation (Sequeira et al., 2013)

Page 188: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

190

Discussion

Species distribution models can provide information on suitable habitat for species

occurrence (e.g., Guisan & Thuiller, 2005), and coupling them to climate forecasts

provides a strong approximation of potential future distributional shifts in response to

warming (Araújo et al., 2005; Hannah et al., 2002). Access to a unique global dataset

of at-sea whale shark sightings provided an opportunity to model both current and

future habitat suitability over the entire range of this pan-oceanic species. With sea

surface temperatures predicted to increase by at least 2 º C on average by 2070

(under a no-climate-policy reference scenario), our models forecast both a (weak)

pole-ward shift and contraction of suitable area, with loss of habitat (~ 6%) occurring

mainly in the warmer equatorial region of the Atlantic and Indian Oceans (Figure 27).

Such a contraction is congruent with predictions for other warm-water plankton-feeding

species (e.g., Southward et al., 1995).

The current range of whale sharks is thought to be predominantly between 30

ºN and 35 ºS, which generally accords with our current worldwide habitat suitability

predictions for the austral spring and summer (Figure 26). However, the environmental

envelope predicted during the boreal spring and summer shifts to latitudes as high as

40 ºN – latitude where whale sharks have been reported more recently (e.g., Portugal)

(Rodrigues et al., 2012).

Within the Indian Ocean, our current predictions agree with the distribution of

observations in that basin (Figure 26) (Rowat, 2007; Sequeira et al., 2013). For

example, we predicted high suitability during the first three months of the year around

India, Maldives and Bangladesh, Djibouti, Kenya and Mozambique, which concurs with

seasonal observations (Rowat, 2007). In the second quarter of the year, we also

predicted high suitability at Ningaloo Reef (Australia), the Mozambique Channel and

Page 189: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

191

Gujarat (India), again agreeing with seasonal observations (Pravin, 2000; Rowat,

2007). From July to September, suitability was highest around the Mozambique

Channel (Pierce et al., 2010) and also Seychelles (Rowat, 2007), although we lacked

enough data to predict suitability in the northern part of the Indian Ocean. Finally, we

predicted high suitability in the last three months of the year along the east coast of

Africa, Bangladesh and Thailand, which also reflects observations (Rowat, 2007).

In the western Pacific, the relatively low habitat suitability is partially a function

of the higher sighting (fishing) effort there, which we offset in our models. Regardless,

predicted suitability was slightly higher in the Philippines from April to June (Quiros,

2007), and around Taiwan from July to September (Chang et al., 1997) (Figure 26).

Additionally, our models suggest the existence of a ‘corridor’ of suitable habitat,

especially between July and September, linking the eastern and western Pacific. This

longitudinal pattern is possibly associated with warmer temperatures, because the

predicted habitat suitability within this ‘corridor’ was slightly stronger within our

scenario for 2070. Although whale sharks occur in the Gulf of California (Cárdenas-

Torres et al., 2007) and the Galapagos during most of the year, our model failed to

predict higher suitability there – additional data from the eastern Pacific would

probably help improve model performance.

The highest suitability worldwide was in the Atlantic Ocean, which is also the

region for which the highest-resolution and longest-running (since 1980) data were

available. Apart from the Gulf of Mexico and the Azores, there is little other information

available for whale sharks in this region. In the Azores, occurrences peak at the end of

August and September (Sa, 2008), and in the Gulf of Mexico, they peak in the second

quarter of the year (Heyman et al., 2001). These are well predicted in our global

model. However, our models did not predict high suitability during the following quarter

Page 190: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

192

in the Gulf of Mexico where some sightings are known to occur (Motta et al., 2010).

Other suitability congruent with fisheries observations includes the western coast of

Africa mostly between April and September. However, the high suitability predicted in

the southern Atlantic cannot yet be validated due to a lack of observations. The high

suitability predicted off South Africa in the first quarter is congruent with the hypothesis

that whale sharks can cross from the Indian to the Atlantic Ocean through the Cape of

Good Hope (Sequeira et al., 2013).

The top-ranked model for all three oceans included all physical variables

(excluding standard deviation for sea surface temperature) previously identified

(Sequeira et al., in press; Sequeira et al., 2012), but the interaction terms with ocean

demonstrated some differences among oceans. We found evidence for sea surface

temperature to affect occurrence of whale sharks differently in each ocean, being an

important predictor both in the Atlantic and Indian, but with no discernable effect in the

Pacific Ocean. The range of temperatures for January to March in the Pacific (26.7 –

30 º C) was narrowest compared to other oceans (25.2 – 31.2 º C in the Indian, and

19.4 – 28.3 º C in the Atlantic; Figure 24), and falls entirely within the range reported

for whale sharks surface sightings (26.5 – 30 o C) (Sequeira et al., 2012). This might

explain why our models did not show evidence that sea surface temperature affects

whale shark occurrence in the Pacific Ocean, and might also be part of the reason why

the habitat suitability in the entire Pacific Ocean was generally low.

Whale shark occurrence is often associated with productive areas (Hsu et al.,

2007; Taylor & Pearce, 1999), and chlorophyll a, a proxy for phytoplankton

concentration, has been used successfully to predict occurrence (Kumari & Raman,

2010; McKinney et al., 2012). However, we found no evidence that chlorophyll a

contributed to the suitability predictions as we determined previously for the Indian

Page 191: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

193

Ocean (Sequeira et al., 2012), possibly due to the temporal/ spatial lags associated

with primary production and secondary consumers (dilution and downstream effects)

(Bradshaw et al., 2004; Wafar et al., 1984). The broad spatial scale of investigation

might also have contributed to reducing the importance of primary production (cf.

Bradshaw et al., 2004), especially in the open ocean where variation in chlorophyll a

concentration is low.

According to Thuiller (2004), projections on climate change impacts in species

occurrence should be developed considering the species’ entire range. In our case,

this range is circumglobal, which is the reason behind our extrapolation to the total

area where predictor values are within the range used to calibrate the model. The

climate-induced impact on species occurrence is usually investigated by the coupling

climate forecasts with species distribution models (Araújo et al., 2005; Hannah et al.,

2002). However, this coupling adds extra limitations (as specified in the Introduction),

and results from such modelling frameworks are now considered to provide only a first

approximation of possible changes. In the same way, our predictions of future whale

shark habitat changes are intended only to be informative, providing only a baseline

for temperature-dependent predictions. For a more comprehensive study on the

influence of temperature in the occurrence of this species, not only should more

climate-change scenarios be tested (including multiple global circulation model

outputs), other species distribution models could be generated to assess variability in

habitat suitability results, examine error and determine confidence intervals (Araújo &

New, 2006; Thuiller, 2003). We extended our main results using the BIOMOD

modelling platform (see Appendix D – BIOMOD run). The ensemble habitat suitability

results for the area covered by the fisheries we obtained were similar to those derived

from GLMM (Figure D2 in Appendix D).

Page 192: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

194

Moreover, to improve our model predictions, more sightings data (ideally with

circumglobal coverage) should be used. We have used the largest, most extensive

dataset derived from tuna fisheries logbooks; however, it only covered a portion of

areas managed by Regional Fisheries Management Organisations (New Zealand

Ministry for Primary Industries website at fs.fish.govt.nz/Page.aspx?pk=103&tk=319.).

Therefore, there is scope to develop our models based on data from other areas, such

as the eastern Pacific Ocean (managed by the Inter-American Tropical Tuna

Commission), which could assist improving the overall model results. As suggested in

the previous chapter, a more collaborative effort between researchers and fisheries

management organisations would assist improving scientific research and, in this

specific case, potential improving our model fitting.

The opportunistic dataset we used only contained whale shark presences

relative to the sea surface only. The latter also applies to the set of environmental

predictors we used to build the models. While the first drawback could be addressed

by randomly generating pseudo-absences (Barbet-Massin et al., 2012; Phillips et al.,

2009) and repeating this procedure multiple times to account for any biases derived

from pseudo-absence selection, the second is not simple to address. Environmental

(and sightings) data spanning subsurface dynamics are not currently available for

practical model development, especially where the study area covers most of the

global ocean extent. Further, the whale shark sightings dataset I used included recent

sightings (up to 2010) that might herald future patterns of distribution change.

Recently, whale sharks have been seen in locations where they were not expected to

occur (Duffy, 2002; Rodrigues et al., 2012; Turnbull & Randell, 2006), and so the data

I used could already epitomise shifts in this species’ occurrence. It is therefore

possible that my results show an already biased range of suitable habitat for this

Page 193: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

195

species. However, due to the large area under study and the limited data on

presences, excluding part of the dataset available was not an option.

Despite the current lack of alternatives to fisheries-collected sightings in the

open ocean (Jessup, 2003), combining remotely sensed data with opportunistically

recorded occurrences has, however, proved its worth for estimating whale shark

distributions (Kumari & Raman, 2010; McKinney et al., 2012; Sequeira et al., 2012).

Indeed, most of our understanding of whale shark ecology and biology hails from

opportunistically collected datasets by the eco-tourism (e.g., Wilson et al., 2001) or

fishing industries near shore at aggregation sites (e.g., McKinney et al., 2012). As

such, we should endeavour to analyse all available datasets to conserve this pan-

oceanic species and largest remaining fish – refusal to examine multiple lines of

evidence risks imperilling the species further through lack of ecological understanding.

Page 194: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

196

CHAPTER VII. GENERAL DISCUSSION

Overview of my research

My research provided the first ocean-scale and global distribution models of the

world’s largest fish, the whale shark (Rhincodon typus). Here, I applied a novel and

comprehensive modelling framework that took spatial variation and sub-optimal

sampling effort data into account while teasing out temporal patterns. I showed how

appropriate statistical approaches can be applied to oceanic species for which data

are incomplete, fragmented and uneven, such as for the whale shark, highlighting the

value of unique datasets collected opportunistically. My approach also provides a

mathematical platform for marine climate change predictions that have typically lagged

behind terrestrial systems, and shows the possible impacts of a warming climate on a

pan-oceanic shark species.

In terms of the modelled organism on which I focussed by analytical

framework, my results highlight the need to account for global migratory behaviour

both in whale shark research programs, and conservation management decisions. As

a starting point, I suggest that the whale shark’s current IUCN Red List status of

Vulnerable (Norman, 2005) be revised. According to my spatially explicit trend results,

the number of whale sharks observed by tuna fisheries, which cover a large extent of

the three major oceans, has been reduced to about 50 % in the last decade, both in

the Atlantic and Indian Oceans (there are no data available prior to 2000 for the Pacific

Ocean). The causes for this reduction at such a broad spatial scale (possibly reflecting

a global population trend) are not yet understood and are potentially still occurring and

possibly irreversible. These characteristics are captured within the IUCN Red List

Page 195: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

197

Criterion A for Endangered species. Moreover, considering the life span of whale

sharks, 10 years might correspond to only 10 % of a whale sharks lifetime (or even up

to 25% of a single generation). As I suggested in Chapter V, despite the possible

cyclical occurrence of whale sharks in each ocean basin, the observed reduction in

sightings is occurring simultaneously in a large part of the species known geographical

range. Due to their long life span, late age at maturity, and the relatively little

information available on their reproduction, this species might actually already be

facing a high risk of extinction in the wild from past and ongoing exploitation and

climate warming. Therefore, I recommend that it should be considered at least for the

status of Endangered.

Other criteria from the IUCN Red List class of Endangered might also apply.

For example, Criterion C relates to the total number of mature individuals. Most whale

sharks spotted in specific aggregations are immature (Graham & Roberts, 2007;

Heyman et al., 2001; Rowat et al., 2011; Wilson et al., 2001), and available population

estimates are around 300-500 individuals for a single site at Ningaloo, Australia

(Meekan et al., 2006). Moreover, whale sharks are thought to attain maturity at ~ 9

metres in total length (Colman, 1997), and there is evidence that the average size of

sighted sharks is declining (Bradshaw et al., 2008). Considering that whale sharks are

only rarely sighted, that most of the sightings are immature animals, and viewing

aggregations as possible sub-populations, Criterion C2ai could also apply: “population

size estimated at < 2500 mature individuals, in continuing decline (observed, projected

or inferred) and with < 250 mature individuals per subpopulation.

In addition to the suggestion of status change from Vulnerable (currently

based on the IUCN Red List Criterion A2bd+3d) to Endangered (based on Criterion

A2ad + 4d or C2a1), my research also provides an updated global map reflecting

Page 196: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

198

seasonality in whale shark habitat suitability, which can now be added to the whale

shark section of the IUCN Red List. My thesis answers some of the most important

questions about the ecology of whale sharks to assist with their long-term

conservation. In summary, I proposed how whale sharks are connected at a global

scale, how their temporal patterns of occurrence probably reflect their migratory

behaviour, and how their distribution depends on environmental conditions that shift

seasonally. The latter can, to some extent, explain the known patterns of the species’

seasonal occurrence, and also how a principally warm-water organism can travel

beyond its ideal thermal nice (i.e., to latitudes > 30 °N and 30 °S) (e.g., Rodrigues et

al., 2012). The underlying hypothesis that climate change is already affecting this

species’ distribution underscores an urgent need to measure their global occurrence

and inter-ocean connectivity.

While focusing on the global patterns of whale shark ecology and connectivity,

my research highlighted the following specific findings:

1. Timing and distribution patterns of whale shark occurrence suggest a

connection between several aggregation sites among the three

largest ocean basins (Chapter II);

2. Whale shark habitat suitability in the Indian Ocean is correlated mainly

with spatial variation in sea surface temperature, and shifts seasonally

(Chapter III);

3. The temporal variation in whale shark probability of occurrence could

be representative of inter-decadal cycles (i.e., cycles ≥ 15 years)

(Chapter IV);

Page 197: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

199

4. The probability of whale shark occurrence is asynchronous across

oceans, but even being cyclical, the decreasing number of sightings

reported in the fisheries datasets (at an ocean-basin scale) agrees

with the reported declines from near-shore aggregations (Chapter V)

5. With increasing sea surface temperature from anthropogenic climate

disruption, whale shark habitat suitability is expected to shift pole-

wards, and currently suitable habitat within the warmer equatorial

region of the Atlantic and Indian Oceans will potentially be lost

(Chapter VI).

A worldwide perspective

My thesis’ global perspective indicates that a more comprehensive focus can, and

should be adopted in future research endeavours. This shift of focus from local

aggregations to wider areas (e.g., regional or global) engenders new hypotheses

about the ecology of this species that can be tested as technology improves, more

tagging results are obtained and genetic evidence is refined.

One of the first questions that follows naturally from my global connectivity

approach (Chapter II) is: If individual whale shark movements do span the world’s

oceans, how are their global movement patterns defined? My results show that

environmental conditions, such as temperature and depth, are important determinants

of whale shark occurrence, and could therefore be related to their migratory behaviour.

For example, the seasonal clockwise shift in habitat suitability I found in the Indian

Ocean (Chapter III) could partially explain why re-sightings occur in the same near-

shore locations in following years. My temporal analysis describing possible cyclic

Page 198: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

200

patterns (Chapter IV and V) could help explain the variation recorded in inter-annual

occurrences in several locations (Graham & Roberts, 2007; Rowat et al., 2009a;

Wilson et al., 2001). When expanded from the regional (ocean basins) to a global

scale (Chapter VI), my analysis revealed seasonal suitable habitat in some critical

areas, such as around South Africa. This is the first insight into how migration between

the Indian and Atlantic is possible.

Despite the role environmental conditions play in determining whale shark

occurrence, to clarify movement patterns requires an understanding of population

structure. It has been suggested that whale shark migratory patterns could reflect sex-

and age-specific behaviour (Borrell et al., 2011; Eckert & Stewart, 2001; Ramírez-

Macías et al., 2007; Rowat et al., 2011). In Chapter II where I collated and reviewed all

available whale shark movement data, including published tracks, I highlighted two

points. First, the longest whale shark tracks were mostly reported for females, and

second, most of the larger sharks (> 6 metres in total length) travelled on average

greater distances than smaller sharks (Chapter II). These findings add further support

to the hypothesis that whale shark movement patterns might be associated with

behavioural segregations.

Consequently, we might ask: If movement patterns are associated with

population structure and environmental changes, will different whale shark classes

(length/age or sex) be affected differently by climate change? Although my research

did not focus on defining patterns for different classes (the focus was on occurrences,

with no data available on sex or length/age), my global model suggests that whale

shark habitat suitability will be reduced in locations of the central Indian and Atlantic

Oceans as the climate warms. In the Indian Ocean aggregation sites in particular,

mostly immature males are observed in the area which I projected as having the

Page 199: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

201

highest probability of warming – such as the Seychelles (Rowat & Gore, 2007) and

Maldives (Riley et al., 2010). If this immature male class is constrained to using these

areas, then climate change will possibly have a stronger effect on them than other

components of the population. In the Atlantic Ocean, the predicted reduction in habitat

suitability in the central area might impose a separation (north and south) of the

population in the future. If the longer migrations, such as the one between the

Caribbean Sea to near the Saint Peter Saint Paul archipelago reported by Hueter

(2008), are limited to females, and are associated with specific behaviours such as

reproduction, then climate change might affect this behaviour (e.g., by imposing a

physical barrier to these migrations), and therefore lead to severe impacts on the

global whale shark population.

A related question is whether fishing and tourism are also impacting only

specific classes. If sex and age segregation is indeed happening, with juvenile males

occurring near shore and females being more oceanic, then any habitat disturbance

caused by coastal development, shipping and/or tourism could have a greater impact

on juvenile males, while broad-scale tuna fisheries might affect mostly females. If true,

this notion highlights the need to measure whale shark population structure, and to

define the extent of disturbance, damage and survival rates derived from human-

related activities (Bradshaw et al., 2007).

Better management

My results provide a strong incentive for current whale shark management policies to

be revised. There are currently no global measures in place to limit the exploitation of

live whale sharks (Rowat & Brooks, 2012). Whale sharks were recently included in

Page 200: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

202

Appendix II of the Convention on Migratory Species (CMS, 2010), which could lead to

the development of international conservation measures (Rowat & Brooks, 2012).

However, current management of whale sharks occurs mostly within confined tourist

locations through the imposition of a maximum number of tourism operator licenses,

and by implementing codes of conduct to minimise disturbance in some countries

(Cárdenas-Torres et al., 2007; Pierce et al., 2010; Quiros, 2007). In the 30 years after

the first initiative for international management of marine resources (UNCLOS, 1982),

and despite whale sharks being classed as Vulnerable in the IUCN Red List

(www.iucn.org), a global effort to protect whale sharks is only indirectly supported

through control of international trade of specimens (Appendix II of the Convention of

Trade in Endangered Species of Wild Fauna and Flora;

www.cites.org/eng/app/appendices.php). Also, national bans have been imposed on

targeted commercial fisheries of whale sharks, but they lack enforcement (Riley et al.,

2009), and the global-scale fisheries interaction with whale sharks is currently not

regulated in any way.

My global connectivity model implies that whale sharks cross multiple national

and high-seas fisheries jurisdictions (whales sharks occur near the shore of > 100

countries; Figure 5 in Chapter II), so if long-term conservation of this species is to be

achieved, active global conservation measures need to be implemented and enforced.

The large extent at which interactions between tuna fisheries and whale sharks

occurrences manifest (see Figure 23 in Chapter VI), and the evidence that whale

shark numbers are declining in both touristic locations (Bradshaw et al., 2008;

Bradshaw et al., 2007) and in fisheries interactions (Chapter V) further highlight the

need for global conservation measures. For example, in the Western and Central

Pacific there is on-going discussion on the implementation of a ban on using the whale

Page 201: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

203

shark as tuna aggregation device. However, the implementation of this ban has been

hindered mostly by a lack of support from Japan (see reports of regular sessions of the

Western and Central Pacific Fisheries Commission; www.wcpfc.int/meetings/1). Given

the relatively lower sighting probability in the three oceans in the last decade, this

measure should be not only enforced in the Western and Central Pacific, but also

embraced by all the Regional Fisheries Management Organizations.

Future improvements1

Input data

As I mention throughout the thesis, the datasets I used were collected

opportunistically, and therefore they bring with them a series of inherent complications.

I addressed and accounted for some of these imperfections during model

development. For example, I addressed the temporal and spatial bias of sampling

effort within each grid-cell by including fishing effort in the models as a ‘weight’

(Chapter III) or as an ‘offset’ (Chapters IV, V and VI). The latter accounts for the fact

that more whale shark sightings are expected in areas where fishing effort was higher,

and seems more appropriate to reflect the different exposure of each area to the event

of interest (i.e., a sighting), while the former weights the model’s response according to

the amount of effort observed in each grid cell. However, other problems such as the

coarse resolution of the effort data are also problematic, and can only be addressed by

gaining access to higher-resolution data. Access to better datasets can be achieved if

collaborations are established between researchers and fisheries management

1 Following examiners’ comments, parts of this section were included in the chapters and

therefore the reader will find some repetition.

Page 202: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

204

organisations. Such collaborations would allow researchers to give greater insight on

how a better dataset could be collated for specific objectives using the resources

currently in place. Commercial-in-confidence restrictions on access to fisheries data

(especially sensitive data such as fish catch) delay scientific research from providing

more accurate models on fisheries impacts.

Another important aspect of the data I used, including both whale shark

sightings and environmental variables, is that they pertain mostly to the sea surface.

Environmental data collection at depth, and at an adequate resolution for a regional or

global analysis such as presented here, are currently not available. Therefore, even

though whale sharks dive frequently, assessment of how they explore the vertical

habitat and estimation of its suitability at a broad scale will only be possible when new

technologies are developed to allow vertical data collection at a broad scale.

The datasets I used contained only recorded presences, and therefore two

disadvantages are associated with such data. The first is related to the probability of

detecting a whale shark. There is a high possibility that in any given area, whale

sharks are present but not observed, and so an occurrence is not recorded (e.g., if

shark was below the surface). The second disadvantage relates to the lack of

information on absences. To account for this drawback, I used different methods to

select absence points within the area covered by the fisheries and include them in the

models as pseudo-absences; this is known to affect model results, which I discuss

below. However, even datasets containing absence locations can be problematic

when using them as inputs for species distribution models (Elith et al., 2006; Zaniewski

et al., 2002). This is the case when the reason for recorded absences is not correctly

identified (i.e., they might be associated with poor detection, habitat disturbance,

competition or inappropriate timing rather than with an unsuitable habitat). Therefore,

Page 203: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

205

both the inclusion of ‘real’ absences or pseudo-absences can lead to biased results.

With no current alternatives to fisheries-collected sightings in the open ocean (Jessup,

2003), the regional and global analysis presented here could not have been done

without using these uniquely available datasets.

In the context of climate change, there is yet another point to consider. The

whale shark sightings dataset I used included recent sightings (up to 2010). Recently,

whale sharks have been seen in locations where they were not expected to occur

(Duffy, 2002; Rodrigues et al., 2012; Turnbull & Randell, 2006), and so the data I used

in my models could already epitomise shifts in this species occurrence. It is therefore

possible that my results show an already biased range of suitable habitat for this

species. However, due to the large area under study and the limited data on

presences, excluding part of the dataset available was not an option.

Models

Despite being partially limited by data collected opportunistically, the species

distribution models themselves have several limitations. Although I used presence and

pseudo-absence data in my linear models, I also developed distribution models in

MaxEnt2 (Elith et al., 2011; Phillips & Dudík, 2008) using presence-only data. Even

though the resulting maps of whale shark probability of occurrence were relatively

similar using both methods (Chapter III), the linear models generally provide more

realistic projections (Brotons et al., 2004; Elith et al., 2006; Zaniewski et al., 2002).

2 MaxEnt also requires background data to estimate the species distribution; therefore, it

cannot be considered a ‘true’ presence-only model such as rectilinear envelopes.

Page 204: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

206

Also, Tsoar et al. (2007) concluded that presence-only models are more accurate for

species with narrow niches (not expected to be the case for whale sharks).

To allow binomial estimation in the generalised linear models, I generated

pseudo-absences, but these can influence results and predictions (Barbet-Massin et

al., 2012; VanDerWal et al., 2009; Wisz & Guisan, 2009). It is therefore important to

acknowledge the assumptions made during the generation of pseudo-absences.

Following published suggestions on how to select pseudo-absences accounting for the

same sampling bias in presences (Phillips et al., 2009), or by using pre-defined areas

within some distance from the recorded presences (Engler et al., 2004; Lobo et al.,

2010; Zaniewski et al., 2002), I tried different methods for pseudo-absence selection.

These included a selection of grid cells with a probability weighted by the inverse

distance to the presences, weighted by the total tuna catch, and based on a random

selection of non-presence grid cells (Chapter III). In subsequent analysis, I used only

random pseudo-absence generation, which, being simple to apply, did not produce

prediction maps dissimilar to the ones obtained by the other methods, and it did not

decrease model performance (Chapter III). The last point accords with the results from

a recent publication specifically addressing pseudo-absence selection issues (Barbet-

Massin et al., 2012). Also, the only method that showed better performance (in some

cases only) was dependent on data from the tuna catch, which was only available at a

much coarser resolution.

The number of pseudo-absences chosen can also affect models results. I

made this choice based on the number of presences available for each model. I used

presence:pseudo-absence ratios (i.e., prevalence) of 1:1 and 1:10 in the spatial

analysis, and 1:100 in the temporal models. I also averaged results of multiple model

runs with different sets of pseudo-absences in each model. Barbet-Massin et al.

Page 205: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

207

(2012) tested the influence of prevalence in models accuracy, and concluded that for

generalised linear models, accuracy increased until prevalence was 0.1 (i.e., 1

presence : 10 pseudo-absences – as used in my spatial analysis; Chapter IV and V

and VI), and is constant when a higher number of pseudo-absences are used.

Therefore, using a large number of pseudo-absences is not expected to decrease

models accuracy, but allows enough points for the spatially explicit temporal analysis.

Another related issue is the decision regarding which probability value should

be used to convert the model results into presences or absences (Jiménez-Valverde &

Lobo, 2007; Liu et al., 2005). This value is commonly assumed to be 0.5, but for

presence: absence ratios different to 1:1, this assumption results in incorrect

classification of absences (if prevalence is low, i.e., more absences than presences) or

presences (if prevalence is high) (Jiménez-Valverde & Lobo, 2007). Several methods

can be used to specify this threshold (see Liu et al., 2005), such as averaging the

predicted probabilities, using sensitivity and specificity, the prevalence value, or the

value that maximises �. I considered this threshold equal to the prevalence used,

which is considered a good approach (Liu et al., 2005).

In the final chapter (Chapter VI), I coupled my distribution models with the

results of models projecting future climate (derived from global circulation models)

(e.g., Araújo et al., 2005; Cheung et al., 2010) to produce worldwide predictions of

future whale shark occurrence. According to Thuiller et al. (2004b), projection of

climate change impacts should be developed when considering the entire range for

the species. In the case of whale sharks, I have shown that this can only be achieved

when considering the worldwide distribution. Although the coupling of climate forecast

and species distribution models is considered to be a powerful way to investigate

possible climate-induced distributional shifts in species occurrence (Araújo et al.,

Page 206: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

208

2005; Hannah et al., 2002), it incurs additional limitations, and is now considered to

provide only a first approximation of possible changes. For example, there is

considerable variation in the results of different climate-forecasting models (e.g.,

Araújo et al., 2005; Thuiller, 2004), and the spatial scale of the resulting predictions is

usually too coarse to have any biological relevance (Hannah et al., 2002). Following

recent published suggestions (Fordham et al., 2011; Fordham et al., 2012), I used

averaged results of five climate-forecasting models to minimise uncertainty related to

the underlying generation of the climate prediction. I also down-scaled the climate

predictions before coupling them to my distribution models. However, my predictions

of future habitat changes are intended only to be informative, showing a possible way

by which whale shark distribution might be affected by changes in temperature. For a

more coherent study on the influence of temperature in the occurrence of this species,

more climate change scenarios should be tested along with the inclusion of other

global circulation model results in the ensemble approach.

Results

Validation of model performance is an important component of any modelling

procedure (Franklin, 2010). I validated my models through the use of the 10-fold cross

validation technique (Davison & Hinkley, 1997), leaving a tenth of the dataset out while

running the models, and then using it to validate the results. As a qualitative validation,

I also compared the predicted probabilities of occurrence both to the original data (to

match higher probability areas with number of sightings), and with current knowledge

on whale shark seasonal occurrences in near-shore locations. Independent validation

by using independently collected data could be useful. This will be possible as more

data are collected, and then used to validate my prediction maps. However, this sort of

Page 207: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

209

validation can only be used for predictions of current habitat suitability. Validation of

future projections is, of course, not possible; however, validation of model performance

can be achieved by hind-casting and validating the results with already collected data

(Araújo et al., 2005). As pointed out above, independent model validation in my whale

shark case study was not possible due to data limitations.

Future directions

My research has identified some key aspects of whale shark ecology that should be

analysed further to advance our knowledge of whale sharks. Below, I describe these

aspects presenting some of the gaps that still remain, and I give some suggestions on

where I think further research should be taken to allow timely conservation of this

impressive species.

In Chapter II, where I formulated my hypothesis about global connectivity of

the whale shark meta-population, I suggested that some locations could be whale

shark ‘thoroughfares’. Relatively little research has been done in these locations (e.g.,

Christmas Island and Saint Peter Saint Paul archipelago), and deployment of tags and

tissue collections near potential thoroughfares would likely increase the probability of

capturing pan- or trans-oceanic movements. Throughout my thesis I also highlighted

the value of using opportunistic data, considering the logistic challenges of surveying

widely distributed species over their entire range. Opportunistically collected data,

especially wide-ranging and long-term data such as the fisheries datasets I used, are

essential sources of information to assist with the understanding of the ecology of such

a highly migratory species. The data I used covered only sections of the area

managed by Regional Fisheries Management Organisations (Figure 28a), and so

Page 208: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

210

there is still scope to develop similar models based on data from other areas, such as

the eastern Pacific Ocean (managed by the Inter-American Tropical Tuna

Commission). Also, some of the other Regional Fisheries bodies (see Figure 28b)

might possess data of importance for whale shark studies. A way to collect additional

data that can be useful for similar studies is simply by adding records of whale shark

sightings within catch logbooks also recording the search effort. A more elaborate

approach is also to include sightings of other species. This could be used as a proxy

for absences when a shark is not recorded.

a)

b)

Figure 28: Map of world organisations managing fisheries. a) Regional Fisheries Management Organisations for highly migratory fish stocks (Tuna and Tuna-like). Map obtained from the New Zealand Ministry for Primary Industries website at

A NOTE:

This figure/table/image has been removed to comply with copyright regulations. It is included in the print copy of the thesis held by the University of Adelaide Library.

Page 209: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

211

fs.fish.govt.nz/Page.aspx?pk=103&tk=319. b) Regional Fishery Bodies (RFB) – Map obtained from FAO Fisheries and Aquaculture Department [online]. Rome. Updated 24 September 2012. [Cited 15 October 2012]. www.fao.org/fishery/rfb/en (FI Institutional Websites). Abbreviations: ICCAT – International Commission for Conservation of Atlantic Tuna; IATTC - Inter-American Tropical Tuna Commission; IOTC - Indian Ocean Tuna Commission; WCPFC - Western and Central Pacific Fisheries Commission; CCSBT – Commission for the Conservation of Southern Bluefin Tuna; GFCM – General Fisheries Commission of the Mediterranean; IPHC – International Pacific Halibut Commission; PFC – Pacific Salmon Commission; NPAFC – North Pacific Anadromous Fish Commission; PICES - North Pacific Marine Science Organization; FFA – Pacific Islands Forum Fisheries Agency; SPC – Secretariat of the Pacific Community; OLDPESCA – Latin-American Organization for Fisheries Development; OSPESCA – Central America Aquaculture and Fisheries Organisation; CPPS – Permanent Commission for the South Pacific; SPRFMO – South Pacific Regional Fisheries Management Organisation; CCAMLR – Commission for the Conservation of Antarctic Marine Living Resources; NAFO – Northwest Atlantic Fisheries Organisation; NEAFC – North East Atlantic Fisheries Commission; NASCO – North Atlantic Salmon Conservation Organization; NAMMCO – The North Atlantic Marine Mammal Commission; ICES – International Council for the Exploration of the Sea; WECAFC – Western Central Atlantic Fishery Commission; CRFM – Caribbean Regional Fisheries Mechanism; CECAF – Fisheries Committee for the Eastern Central Atlantic; SRFC – Subregional Fisheries Commission; FCWC – Fisheries Committee of the West Central Gulf of Guinea; COREP – Regional Fisheries Committee for the Gulf of Guinea; COMHAFAT – Ministerial Conference on Fisheries Cooperation among African States Bordering the Atlantic Ocean; SEAFO – Southeast Atlantic Fisheries Organization.

Despite finer-scale evidence that chlorophyll a (used as a proxy for food

availability) is an important predictor of whale shark occurrence (Kumari & Raman,

2010; McKinney et al., 2012), my results revealed that at least at broad spatial scales,

its importance does not persist. Availability of data describing the distribution of whale

shark prey, such as zooplankton abundance (Clark & Nelson, 1997; Jarman & Wilson,

2004), might improve the results of the distribution models. The lack of these data at

spatial scales relevant to oceanic patterns is an important gap that hinders further

insight on how whale shark movement is related to food abundance. Although

availability of such data depends on technological advances, data collection on

zooplankton concentration in limited areas (such as aggregation locations) could be

included in future research programs.

Another gap in our knowledge of whale sharks relates to population structure.

The construction of broad-scale (meta-) population models to connect whale shark

sub-populations should follow the global conceptual movement model presented in

Chapter II. Such population models could be based on inferred demographic

Page 210: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

212

(Bradshaw et al., 2007) and movement rate data, but key aspects of whale shark life

history still need to be measured. More tagging and genetic studies will assist in

describing those aspects, and in quantifying if there are any stage-specific movement

trends. Studies using ‘proxy’ organisms, such as region-specific parasites (Braicovich

& Timi, 2008; Lester, 1990), could also assist in describing movement for this species.

One way to test the validity of my connectivity model is by focusing tagging and

photographic-identification studies on ‘connectivity regions’ - regions where evidence

for connection already exists (Chapter II). Examples of these regions include

Madagascar, the Seychelles and Tanzania, or the region including the Philippines,

Malaysia and Gulf of Thailand.

Tagging approaches can provide insight into the movement behaviour of

migratory species, however they bring their own problems (e.g., Brunnschweiler et al.,

2009). For example, despite the long tracks obtained for some of the species in the

successful TOPP program (tagging of Pacific predators; Block et al., 2011), most of

the temporal coverage of the individual tracks was less than one year (maximum

duration for different species ranged from 19 to 358 days) (Block et al., 2011

supplementary material). This multinational, collaborative program (part of the Census

of Marine Life) received funding from several organisations to allow the deployment of

several thousands of tags, of which only less than half resulted in individual tracks

(Block et al., 2011). However, some of the problems associated with tagging

procedures can be improved as technology progresses. Some suggestions to

maximise the duration of tracking have already been made, and should be adopted

when planning new tagging studies. These suggestions include using fast-loc GPS

technology (Hays et al., 2007), applying anti-fouling paint (Hammerschlag et al., 2011),

and use copper salt-water switches (Hays et al., 2007). But improvements to prevent

Page 211: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

213

premature detachment (Brunnschweiler et al., 2009) are also needed, and new

attachment options should be considered. To maximise the potential for measuring

long-distance migrations, I suggest that tagging studies target multiple sharks

simultaneously within the same aggregation location (cf. Eckert & Stewart, 2001;

Wilson et al., 2006), and that deployments are made close to the end of the peak

aggregation season. Additional tagging studies on whale sharks released from pelagic

fisheries nets would also provide information on oceanic movements to complement

results obtained from known aggregations.

In the context of future distribution shifts, my results provide only a baseline

for temperature-dependent predictions, and to provide an initial picture of the most

plausible and/or extreme predictions under warmer conditions, multiple climate change

scenarios should be tested along with multiple global circulation models outputs.

However, other factors such as changes in zooplankton abundance might also

influence the future distribution of whale sharks. Climate change is currently affecting

plankton distributions (Edwards & Richardson, 2004; Hinder et al., 2012), and so my

prediction of future suitable habitat might be further restricted by food limitations. The

long-term management of whale sharks will therefore require a better understanding of

their basic ecology and demography, but also of how these will be altered under a

warming climate.

So, for timely conservation of this iconic species, I suggest that the following

measures should be taken:

� Develop tagging and photo-identification studies covering ‘thoroughfares’ and

‘connectivity regions’. This can be done by establishing international

collaborations. To avoid divergence of methods used, guidelines for analysis

should be shared or similar data should be analysed by the same group.

Page 212: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

214

� Establish links with fisheries management organisations to:

o give insight into data collection and gain access to better data

o participate in observer and/or research programmes including tagging

and identification of whale sharks used in fisheries interactions

� Include sampling of zooplankton abundance in forthcoming research

programs

� Consider the development of models including data on whale shark prey and

population structure

� Develop active global management strategies on the use of whale sharks by

commercial fisheries.

The last measure would have the most immediate positive impact on whale

shark populations and would, in turn, allow more time to collect new data to assist with

the conservation of this species. Safe Release Guidelines have been suggested at the

regular sessions of the Western and Central Pacific Fisheries Commission, as well as

the inclusion of 100 % observer programmes and tagging studies. Such data are

fundamental to generate appropriate management strategies. Also, the IUCN status

for the species should be revised and changed from Vulnerable to Endangered.

In conclusion, I have demonstrated how opportunistic data can be used in

species distribution modelling by addressing specific problems such as effort bias and

spatial autocorrelation. By using similar modelling tools with new input data, including

food availability and information on population structure, the timing of whale shark

appearances at specific sites and their movements could be predicted, and

subsequently used to examine the drivers of population trends. As a conservation

Page 213: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

215

ecologist, I believe we now have a last (and time-constrained) opportunity to preserve

this unique species, and action is needed immediately.

Page 214: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

216

AFTERWORD

On a logistic note, it is worth referring that while doing my Ph.D., I have participated in

four whale shark field trips where I collected data to identify individuals (photo-

identification, stereo-measurement and gender), collected biopsy samples (for genetic

analyses) and deployed tags (to track movement). Due to delays in recovering tags,

movement data could not be included in this thesis. Other field work components, as

derived from the individuals’ data collection (identification, measurements and

genetics) are still being analysed (and compared with data collected in other study

sites), and results will soon be evaluated in the context of my global approach.

Page 215: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

217

APPE

NDIX

A –

Infe

rred

glob

al co

nnec

tivity

of w

hale

shar

k pop

ulat

ions

Tabl

e A1:

Sum

mar

y of r

epre

sent

ative

wha

le sh

ark s

tudi

es (s

ince

1996

) inc

lude

d in

Fig

ure 4

. Gr

oupe

d by

(1) o

ccur

renc

e, po

pulat

ion

and

corre

lates

for o

ccur

renc

e, (2

) mov

emen

t, (3

) tou

rism

and

cons

erva

tion,

(4) m

orta

lity

and

(5) f

orag

ing,

in d

iffer

ent l

ocat

ions

wor

ldwi

de.

Lo

cal

Occu

rrenc

e Po

pulat

ion

Bio-

phys

ical c

orre

lates

Move

men

ts

Eco-

tour

ism /

cons

erva

tion

Morta

lity

Fora

ging

Atlantic Ocean

Cana

da

(Tur

nbull

& R

ande

ll, 20

06)

- -

- -

Gulf o

f Mex

ico

(Hoff

maye

r et a

l., 20

06)

(Bur

ks e

t al.,

2006

) (C

árde

nas-P

alomo

et a

l., 20

10)

(McK

inney

et a

l., 20

12)

- (S

uáre

z et a

l., 20

07)

- (H

offma

yer e

t al.,

2007

) (H

ueter

, 200

7)

(Mott

a et a

l., 20

10)

Carib

bean

Sea

(G

raha

m &

Robe

rts, 2

007)

(d

e la P

arra

Ven

egas

et a

l., 20

11)

(Gra

ham

et a

l., 20

06)

(Giffo

rd e

t al.,

2007

b)

(Gra

ham

& Ro

berts

, 200

7)

(Quir

os, 2

005)

(C

árde

nas-T

orre

s et a

l., 20

07)

(Hey

man e

t al.,

2010

) (Z

iegler

et a

l., 20

11)

(Diaz

, 200

7)

(Hey

man e

t al.,

2001

)

Braz

il (G

adig,

2007

) (H

azin

et a

l., 20

08)

- (G

adig,

2007

) -

-

Ango

la/Ni

geria

(W

eir, 2

010)

-

- -

-

Indian Ocean

KwaZ

ulu-N

atal

(Sou

th Af

rica)

(C

liff e

t al.,

2007

) (G

ifford

et a

l., 20

07a)

(G

ifford

et a

l., 20

07b)

-

(Bec

kley e

t al.,

1997

) -

Moza

mbiqu

e -

(Bru

nnsc

hweil

er e

t al.,

2009

) (B

runn

schw

eiler

& S

ims,

2012

) (P

ierce

et a

l., 20

10)

- -

Keny

a -

- (B

asse

n, 20

07)

(Njon

jo, 20

07)

- -

Mada

gasc

ar

(Jona

hson

& H

ardin

g, 20

07)

- -

- -

Seyc

helle

s (R

owat

et a

l., 20

09a)

(R

owat

et a

l., 20

09b)

(R

owat

& Go

re, 2

007)

(R

owat

et a

l., 20

11)

(Row

at, 20

07)

(Row

at &

Enge

lhard

t, 200

7)

- -

India

(Kum

ari &

Ram

an, 2

007)

(K

umar

i & R

aman

, 201

0)

- (P

ravin

, 200

0)

(Josh

i et a

l., 20

07)

(Han

fee, 1

997)

(H

anfee

, 200

1)

(Bor

rell e

t al.,

2011

)

Page 216: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

218

Indian Ocean

KwaZ

ulu-N

atal

(Sou

th Af

rica)

(R

iley e

t al.,

2010

) -

- (R

iley e

t al.,

2009

) -

Djibo

uti

(Row

at et

al.,

2007

) (R

owat

et a

l., 20

11)

- -

- Pa

kistan

and

Ba

nglad

esh

(Row

at et

al.,

2008

) -

- -

-

Thail

and

(T

hebe

rge &

Dea

rden

, 200

6)

- -

(Hae

traku

l et a

l., 20

07)

-

Chris

tmas

Islan

d (A

ustra

lia)

(Hob

bs e

t al.,

2009

) -

(Dav

ies, 2

007)

-

-

Ning

aloo

(W

ester

n Aus

tralia

)

(Wils

on e

t al.,

2001

) (M

eeka

n et a

l., 20

06)

(Bra

dsha

w, 20

07)

(Bra

dsha

w et

al.,

2007

) (N

orma

n & S

teven

s, 20

07)

(Bra

dsha

w et

al.,

2008

) (H

olmbe

rg e

t al.,

2008

) (H

olmbe

rg e

t al.,

2009

) (S

leema

n et a

l., 20

10a)

(S

leema

n et a

l., 20

10b)

(Gun

n et a

l., 19

99)

(Wils

on e

t al.,

2006

) (W

ilson

et a

l., 20

07)

(Dav

is et

al.,

1997

) (B

rads

haw,

2007

) (M

au &

Wils

on, 2

007)

(C

atlin

& Jo

nes,

2009

) (C

atlin

et a

l., 20

10)

(Fitz

patric

k et a

l., 20

06)

(Colo

mer,

2007

) (S

peed

et a

l., 20

08)

(Spe

ed e

t al.,

2009

)

(Wils

on &

New

boun

d, 20

01)

(Wils

on, 2

002)

(Ja

rman

& W

ilson

, 200

4)

(Tay

lor, 2

007)

(M

eeka

n et a

l., 20

09)

Indo-Pacific

Indon

esia

(Stac

ey e

t al.,

2008

) (E

cker

t et a

l., 20

02)

- (W

hite &

Cav

anag

h, 20

07)

-

Philip

pines

(A

ca &

Sch

midt,

2011

) (E

cker

t et a

l., 20

02)

(Quir

os, 2

005)

(Q

uiros

, 200

7)

(Pine

et a

l., 20

07)

(Alav

a et a

l., 19

97)

-

Malay

sia

- (E

cker

t et a

l., 20

02)

- -

-

Taiw

an

- (H

su e

t al.,

2007

) (H

su e

t al.,

2012

) (C

hen e

t al.,

1997

) (C

hen &

Phip

ps, 2

002)

-

Pacific Ocean Ch

ina

- -

- (L

i et a

l., 20

12)

- Ja

pan

- (M

atsun

aga e

t al.,

2003

) -

- -

New

Zeala

nd

(Duff

y, 20

02)

- -

- (D

uffy,

2002

)

Gulf o

f Cali

fornia

-

(Eck

ert &

Stew

art, 2

001)

(R

odríg

uez-D

owde

ll et a

l., 20

07)

- (C

lark &

Nels

on, 1

997)

(N

elson

& E

cker

t, 200

7)

Regi

onal

/ Glo

bal

(Gra

ham,

2007

) (B

rook

s et a

l., 20

10)

(Seq

ueira

et a

l., 20

12)

(Nor

man,

2007

) (F

ordh

am, 2

007)

(R

owat,

2007

) (R

oman

ov, 2

002)

-

Page 217: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

219

APPE

NDIX

B –

Ocea

n-sc

ale p

redi

ctio

n of

wha

le sh

ark d

istrib

utio

n

Tabl

e B1:

Sum

mar

y of t

he M

axEn

t mod

els re

latin

g pr

obab

ility o

f wha

le sh

ark (

Rhin

codo

n ty

pus,

Smith

1828

) occ

urre

nce t

o in

divid

ual o

cean

pro

perti

es.

- slo

pe, d

epth

, dist

ance

to sh

ore (

shor

e), m

ean

sea s

urfa

ce te

mpe

ratu

re (S

ST m

ean)

and

its q

uadr

atic

term

(SST

mea

n2),

sea s

urfa

ce te

mpe

ratu

re st

anda

rd d

eviat

ion

(SST

SD)

and

chlo

roph

yll a

(Chl

a). T

hree

diff

eren

t met

hods

wer

e use

d fo

r gen

erat

ing

pseu

do-a

bsen

ces:

rand

om, in

vers

ely d

istan

t to

whale

shar

k sig

htin

g lo

catio

ns (I

DW) a

nd b

ased

on

tota

l tu

na ca

tch

(tuna

), an

d re

sults

are s

hown

in ro

ws 1

to 3

when

onl

y lin

ear a

nd q

uadr

atic

feat

ures

wer

e use

d. T

he tw

o las

t row

s sho

w re

sults

whe

n Ma

xEnt

mod

el wa

s give

n th

e ful

l ba

ckgr

ound

(bac

kgro

und)

with

cova

riate

dat

a ava

ilabl

e and

vary

ing

the f

eatu

re ty

pe u

sed

in th

e mod

el. T

he Ja

ck-k

nife

test

resu

lts sh

owin

g wh

ich en

viron

men

tal v

ariab

le ha

d th

e hi

ghes

t gain

whe

n us

ed in

isol

atio

n (b

est a

lone

) and

whi

ch en

viron

men

tal v

ariab

les d

ecre

ased

the g

ain th

e mos

t whe

n om

itted

(wor

st w

ithou

t) ar

e also

show

n to

geth

er w

ith th

e va

lue o

btain

ed fo

r the

area

und

er th

e cur

ve (A

UC) t

est.

Show

n fo

r eac

h m

odel

are t

he p

erce

nt o

f con

tribu

tion

(con

t(%) a

nd th

e per

mut

atio

n im

porta

nce (

Imp)

- re

sults

= 0

are s

hown

as

‘-’, v

alues

�� 5

in lig

ht g

rey,

and

high

est v

alues

in b

old.

Se

ason

au

tum

n wi

nter

sp

ring

sum

mer

Mod

el Co

nt(%

) Im

p C

ont(%

) Im

p Co

nt(%

) Im

p Co

nt(%

) Im

p ra

ndom

– lin

ear a

nd q

uadr

atic

feat

ures

SS

T me

an2

4.9

53.4

48.9

40.4

2.3

32

- -

SS

T me

an

3.2

22.7

- -

1.5

42.5

13.7

-

SST

SD

10.2

3.6

9.4

3.2

35.9

11

3.6

8

Chl a

5.6

1.7

37

.5 42

22

.5 3.2

21

.2 28

.8

depth

12

.9 3.8

0.5

2.6

15

.5 2.8

37

.4 39

.3

shor

e 62

.2 14

.7 2.4

9.9

21

.9 8.4

24

22

.2

slope

1.2

-

1.3

1.9

0.3

- 0.2

1.7

Best

alon

e sh

ore

Chl a

Ch

l a

Depth

Wor

st w

ithou

t sh

ore

Chl a

SS

T SD

De

pth

AU

C 0.7

21

0.701

0.6

91

0.668

Page 218: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

220

IDW

– lin

ear a

nd q

uadr

atic

feat

ures

SS

T me

an2

1.1

42.6

8.3

- 2.1

-

- 4.3

SS

T me

an

1.9

29.8

- -

17

3.8

11.6

-

SST

SD

3 0.4

-

- 30

18

.1 2

5.2

Ch

l a

12.6

3.5

52.3

27.2

15.8

34.8

35

35.3

de

pth

34.2

20.8

34.7

49.7

15.9

- 28

.9 45

.6

shor

e 45

.6 2.8

1

9.1

15.4

26.5

21.7

8.1

slo

pe

1.5

0.1

3.6

14

3.8

16.7

0.7

1.4

Be

st al

one

Depth

Ch

l a

Chl a

De

pth

W

orst

with

out

Depth

Ch

l a

Chl a

De

pth

AU

C 0.6

25

0.574

0.6

27

0.607

tu

na –

linea

r and

qua

drati

c fea

ture

s

SST

mean

2 7.6

59

.2 55

.8 61

.6 1.3

24

.8 -

-

SST

mean

1.7

18

.8 -

- 0.6

30

.9 -

-

SST

SD

1.1

- 8.9

1.6

36

.2 16

.7 3.8

8.1

Chl a

9.3

3.6

29

.8 34

.2 22

.5 14

.7 30

.1 22

.5

depth

13

.9 5.2

5.4

2.6

19

2.3

51

.8 58

.4

shor

e 65

.4 12

.7 0.1

-

20.4

10.7

14.1

11

slo

pe

1 0.5

-

- -

- 0.2

-

Be

st al

one

shor

e Ch

l a

Chl a

De

pth

W

orst

with

out

shor

e Ch

l a

SST

SD

Depth

AUC

0.710

0.6

73

0.672

0.6

76

back

grou

nd –

linea

r and

quad

ratic

featu

res

SS

T me

an2

8.9

51.8

48.3

47.6

7.5

28.4

1.7

26.7

SS

T me

an

3.6

17.9

- -

16.3

45

11.1

39

SS

T SD

4.8

5.2

3.8

8.4

35

14

.8 5.4

4.5

Page 219: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

221

Ch

l a

10.8

0.6

33.2

41.6

2.3

1.3

6.2

2.4

depth

28

.3 9.4

11

.8 0.9

19

.3 1.5

54

.7 21

.2

shor

e 39

.8 14

.6 2.3

-

19.6

9 20

.4 6.3

slope

3.7

0.5

0.5

1.6

-

- 0.6

-

Be

st al

one

Depth

Ch

l a

SST

mean

2 De

pth

W

orst

with

out

Depth

Ch

l a

Shor

e De

pth

AU

C 0.9

28

0.837

0.8

58

0.843

ba

ckgr

ound

– au

to fe

atur

es

SS

T me

an2

4.8

11

4.6

12.6

35.8

35.9

8.8

8

SST

mean

14

4.2

2.3

-

0.6

0.3

2.6

0.5

SS

T SD

20

.6 21

.9 7.8

8.7

14

.9 17

.1 6

7.9

Ch

l a

20

25.3

65

65.5

23.7

32.1

47.4

41.3

de

pth

23.7

17.8

16.9

9.8

6.7

2.4

27.8

29.8

sh

ore

12.1

13.9

2.8

3.1

15.7

10.5

7 11

.3

slope

4.9

5.8

0.5

0.3

2.6

1.7

0.4

1.2

Best

alone

Ch

l a

Chl a

SS

T me

an

Chl a

Wor

st wi

thout

Chl a

Ch

l a

Chl a

Ch

l a

AU

C 0.9

61

0.920

0.9

56

0.928

Page 220: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

222

Figure B1: Bathymetry of the Indian Ocean as per the General Bathymetric Chart of the Oceans (1 minute grid ~ 1.8 km).

Figure B2: Seasonal chlorophyll a concentration (Chl a) obtained from averaged of weekly SeaWiFS satellite composites at a 9 km spatial resolution from 1997 to the end of 2007.

20˚

20˚

-20˚

20˚

-20˚

Page 221: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

223

Figure B3: Seasonal sea surface temperature (SST) obtained from averaged weekly MODIS-Aqua satellite composites at a 9-km spatial resolution from 2002 to the end of 2007.

Figure B4: Probabilities (top panel) and resulting pseudo-absences locations (bottom panel) generated by three different techniques. (Left: randomly, middle: inversely proportional to distance from shark locations, and right: directly proportional to tuna catch).

20˚

-20˚

20˚

-20˚

Page 222: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

224

Tuna

Cat

ch -

AUTU

MN

Tuna

Cat

ch -

SPRI

NG

Figure B5: Moran’s I plots showing the reduction in spatial autocorrelation in the GLM residuals. When a random effect was included to the models. Moran’s I is shown in the y axis and the lags ni the x axis. Results for only one technique (tuna) in autumn and spring (when more presence data were available) are shown due to space constraints. Model 1: Presence ~ Slope; Model 2: Presence ~ Depth, Model 3: Presence ~ Shore, Model 4: Presence ~ SST variables”, Model 5: Presence ~ Chl a, Model 6: Presence ~ Physical variables + SST variables", Model 7: Presence ~ Physical variables + SST variables + Chl a.

Page 223: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

225

Figure B6: Whale shark (Rhincodon typus, Smith 1828) habitat suitability for each season as predicted by MaxEnt when using the full background with only linear and quadratic features.

Page 224: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

226

GLM

/ GLM

M Ma

xEnt

Figure B7: Seasonal habitat suitability of whale sharks (Rhincodon typus, Smith 1828) in the Indian Ocean. Maps show the GLM/GLMM ensemble result of three pseudo-absences generation techniques used per season in the top row and the best-performing (according to AUC) MaxEnt model in the bottom row. The 0.4 estimated probability of whale shark occurrence contour is shown for the ensemble maps for clarity. Black dots represent real whale shark locations as summarised by Rowat (2007); dot size is proportional to frequency of sightings.

Page 225: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

227

Receiving operator characteristic curve / area under the cure (ROC/AUC)

To measure models performance we also calculated the receiving operator

characteristic curve / area under the curve (ROC/AUC) and results are shown in Table

B2. The trend arising when comparing the AUC results between each MaxEnt and

GLM model is similar to the one I obtained previously when comparing � – i.e., when

using the random technique to generate pseudo-absences, GLM was scored higher by

both measures (AUC and �).

Table B2: Comparison of the Area Under the Curve (AUC) and Kappa results for the two modelling approaches used in chapter III: MAxEnt and GLM/GLMM for each of the pseudo-absence dataset generated method (P/A) and for each season: Autumn (Aut), Winter (Win), Spring (Spr) and Summer (Sum).

P/A Approach Aut Win Spr Sum

Rand

om

MaxE

nt Kappa 0.54 0.57 0.50 0.49

AUC 0.72 0.70 0.69 0.67

GLM

Kappa 0.79 0.51 0.59 0.51

AUC 0.87 0.61 0.79 0.63

IDW

MaxE

nt Kappa 0.34 0.10 0.25 0.31

AUC 0.63 0.57 0.63 0.61

GLM

Kappa 0.35 0 0.33 0

AUC 0.82 0.72 0.80 0.74

Tuna

MaxE

nt Kappa 0.64 0.37 0.48 0.43

AUC 0.71 0.67 0.67 0.68

GLM

Kappa 0.68 0.47 0.51 0.56

AUC 0.90 0.75 0.79 0.77

Page 226: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

228

Boyce index comparison between MaxEnt and GLM/GLMM models

We assessed the monotonic relationship (Spearman’s rank correlation) between

habitat suitability classes estimated by the models and the proportion of evaluation

points adjusted by the area covered (Boyce et al., 2002) by calculating the Boyce

index (autumn only). According to this index (Table B3) the results among pseudo-

absence selection techniques and among modelling tools are similar, showing that the

predictions from all models considered are consistent with the presences distribution in

the sampled dataset. Results for the GLM predictions were generally slightly higher.

Table B3: Boyce’s index for the three pseudo-absence generation techniques (P/A) for each modelling technique (MaxEnt and GLM) for autumn. Standard deviation was always < 0.001.

P/A Tool Boyce index

Random MaxEnt 0.827

GLM 0.845

IDW MaxEnt 0.854

GLM 0.911

Tuna MaxEnt 0.829

GLM 0.952

Page 227: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

229

BIOMOD, GBM and BRT models for the Indian Ocean

BIOMOD

We set the models according to the modelling summary shown in Table B4, and

details on the best model selected according to True Skill Statistics and Receiver

Operator Curve are shown in Table B5. The variable importance ranking for each

model considered in BIOMOD is detailed in Table B6.

Table B4: BIOMOD modeling summary. Modeling whale shark distribution in the Indian Ocean, using the same dataset used in Chapter III. ANN –Artificial Neural Networks; CTA – Classification Tree Analysis; GAM – Generalised Additive Models; GBM – Generalised Boosted Models (or BRT – Boosted Regression Trees); GLM – Generalised Linear Models; MARS – Multivariate Adaptive Regression Splines; FDA – Flexible Discriminate Analysis; RF – Random Forest.

Number of species modelled 1 (whale shark)

Numerical variables Chla_av.ln Dist2Shore.ln Depth.c Slope.ln SST_av SST_sd logEff (note: logEff could not be included as an offset within BIOMOD)

Number of evaluation repetitions 3

Number of pseudo-absences runs 2

Models selected ANN, CTA, GAM, GBM, GLM, MARS, FDA, RF, SRE

Total number of model runs 72

Page 228: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

230

Table B5: Best model selection in BIOMOD According to True Skill Statistic (TSS) and Receiver Operator Curve (ROC) for each pseudo-absence dataset (PA), and each repetition considered (Rep). Random Forests were not considered for this classification. Bold italic letters indicate when GLM (Generalised Linear Models) was ranked as “best model” within the model runs considered (i.e., only 2 PA datasets). Generally GBM (Generalised Boosted Models) were ranked higher. CTA (Classification Tree Analysis) was also ranked higher in the first repetition of the run with the second set of pseudo-absences.

Best Model

Cross-validation

Independent data

Total score

Cut off

Sensitivity Specificity

PA1 TSS GBM 0.859 0.864 0.888 436.5 96.1 92.7

ROC GBM 0.977 0.973 0.982 532.9 94.3 94.3

PA1 Rep 1

TSS GBM 0.863 0.860 0.887 494.2 95.2 93.5

ROC GBM 0.975 0.973 0.981 539.3 94.0 94.0

PA1 Rep 2

TSS GLM 0.841 0.840 0.853 658.7 91.5 93.8

ROC GBM 0.974 0.972 0.981 496.4 93.3 93.4

PA1 Rep 3

TSS GBM 0.876 0.857 0.886 502.9 94.6 94.0

ROC GBM 0.981 0.973 0.982 510.3 94.1 94.2

PA2 TSS GBM 0.832 0.865 0.883 482.0 96.3 92.0

ROC GBM 0.961 0.975 0.980 547.8 93.8 93.9

PA2 Rep 1

TSS CTA 0.830 0.865 0.887 320.0 94.7 94.0

ROC GBM 0.951 0.973 0.977 570.5 92.7 92.8

PA2 Rep 2

TSS GBM 0.848 0.863 0.881 531.6 95.0 93.1

ROC GBM 0.964 0.974 0.979 560.3 93.3 93.3

PA3 Rep 3

TSS GBM 0.851 0.867 0.879 546.4 95.0 92.9

ROC GBM 0.968 0.975 0.979 580.8 93.5 93.6

The GLM model formula considered within the BIOMOD platform for each of

the pseudo-absence dataset were as follow:

1) Presence ~ logEff + SST_av3 + Depth.c + Depth.c3 + Chla_av.ln3 +

Chla_av.ln + SST_sd3 + Slope.ln + SST_av + SST_av2

2) Presence ~ logEff + SST_av3 + SST_av + Depth.c3 + Depth.c +

Chla_av.ln3 + Chla_av.ln + Slope.ln + SST_av2 + Dist2Shore.ln3 +

Dist2Shore.ln2 + logEff2 + logEff3

Page 229: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

231

Note that the model formulas differ in each run with different pseudo-absences

considered.

The BIOMOD ensemble result for probability of whale shark distribution during

autumn in the Indian Ocean according to the models considered is depicted in Figure

B8, and it shows good agreement with the maps obtained both in Chapter III and

Chapter IV (after re-fitting the model for autumn).

Table B6: Ranking of variable importance for each model considered within the BIOMOD platform. Effort resulted in the variables with highest importance in all models, as expected. Note that this variable could not have been included as an offset in BIOMOD. The models developed throughout the thesis aim at understanding which variables (other than fishing effort) can be used to explain the occurrence of whale sharks. Following effort, depth, SST_av and Chl a, were the variables with the highest importance. ANN –Artificial Neural Networks; CTA – Classification Tree Analysis; GAM – Generalised Additive Models; GBM – Generalised Boosted Models (or BRT – Boosted Regression Trees); GLM – Generalised Linear Models; MARS – Multivariate Adaptive Regression Splines; FDA – Flexible Discriminate Analysis; RF – Random Forest. Chl a – mean concentration of chlorophyll a; Shore – distance to shore (logarithm); Depth – centred depth; Slope – logarithm of slope; SST_av – mean sea surface temperature; SST_sd – sea surface temperature standard deviation.

Models Chl a Shore Depth Slope SST_av SST_sd Effort ANN 0.010 0.011 0.376 0.016 0.003 0.008 0.594

CTA 0.000 0.295 0.275 0.020 0.024 0.014 0.573

GAM 0.079 0.018 0.043 0.013 0.024 0.010 0.509

GBM 0.016 0.010 0.024 0.003 0.057 0.002 0.614

GLM 0.090 0.019 0.062 0.012 0.015 0.000 0.531

MARS 0.189 0.000 0.111 0.001 0.133 0.007 0.589

FDA 0.021 0.010 0.067 0.013 0.066 0.000 0.660

RF 0.106 0.035 0.114 0.013 0.128 0.044 0.485

SRE 0.048 0.023 0.024 0.015 0.021 0.023 0.031

Page 230: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

232

Figure B8: Whale shark autumn distribution in the Indian Ocean as a result of the ensemble of models in BIOMOD. Colour gradient reflects probability of whale shark occurrence with red indicating higher probability values. Only the area covered by the fisheries is shown in the maps. The ensemble result is calculated by averaging across all the models for each run. Shown are: (top left to right order) the presence locations used, and prob.mean and prob.mean.weigth – which use the mean and weighted mean of thresholds (respectively) to convert probabilities into presence/absence data; (bottom left to right order) median – used the median of the thresholds, and Kappa.mean and TSS.mean – use 0.5 as the threshold to distinguish between presence and absence.

Page 231: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

233

BRT / GBM

We adapted the code from Elith et al. (2008) to run the BRT (Boosted Regression

Trees) analysis with our data from the Indian Ocean (autumn season only). The tree

complexity was set to 5, the learning rate was 0.005, and the bag fraction was

considered at 0.5. Effort was included as an offset (logEff).

The maximum number of trees fitted was 1800, with the best number of trees

defined at 1350. The evaluation procedure resulted in a ROC (Remote Operated

Curve) of 0.956. The function “gbm.perspec” was used to check for pairwise

interactions, and the results are shown in Table B7. 3D figures that assist interpreting

the interactions found are depicted in Figure B9. The interaction values represent the

contribution of the pairwise interaction to the predictive performance of the model

(Williams et al., 2010).

The map for distribution of whale sharks obtained by BRT is similar to the

maps obtained by the other modelling techniques used (Figure B10).

Table B7: Analysis of pairwise interactions between variables using BRT Bold font indicates the top ranked two interactions. A 3D interpretation of these two interactions in shown in Figure B9.

Chl a Shore Depth Slope SST_av SST_sd

Chl a 0 5.07 13.35 2.85 12.12 40.44

Shore 0 0 124.73 1.04 14.57 44.15

Depth 0 0 0 2.04 71.99 43.49

Slope 0 0 0 0 13.24 10.48

SST_av 0 0 0 0 0 39.45

SST_sd 0 0 0 0 0 0

Page 232: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

234

When running the model using the GBM package, we used the model formula

including all the predictors as linear terms, quadratic terms for both temperature and

depth, and effort as an offset. This package allows for computation of the Friedman’s

H-statistic through the function interact.gbm. The Friedman’s H statistic, which is

returned by that function, provides an assessment of the relative strength of interaction

effects in the models. The value returned by interact.gbm was 2.9x10-14, confirming the

low strength of interaction effects in the model.

Figure B9: 3D interpretation of the two top ranked pairwise interactions found with BRT. Left: Interaction between distance to shore (logarithm; Dist2Shore.ln) and centred depth (Depth.c). Right: Interaction between sea surface temperature (SST_av) and centred depth (Depth.c). The maximum values for the interactions are: 0.02 (left) and 0.06 (right).

Page 233: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

235

Figure B10: Map with probability of whale shark occurrence in the Indian Ocean obtained by BRT. Colour gradient reflects probability of whale shark occurrence, with red indicating higher probabilities. Only the area covered by the fisheries is shown in the map. Higher probabilities of occurrence are similar to the ones described previously in chapter 3.

Page 234: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

236

APPENDIX C – Spatial and temporal predictions of decadal trends in

Indian Ocean whale sharks

Model predictors (Step 1)

Here we re-fitted the whale shark distribution model for autumn (Sequeira et al., 2012),

using generalised linear mixed-effects models estimated with a Laplace approximation

to the likelihood (lmer function; Bates et al., 2011) rather than using a penalized quasi-

likelihood approach (glmmPQL function in R; Venables & Ripley, 2002b) as we used in

Sequeira et al. (2012). Using the lmer approach we can now obtain information

criteria output to rank models. When re-running the autumn spatial model, we also

used a different response variable for these models by increasing the pseudo-

absence:presence ratio from 1:1 (Sequeira et al., 2012) to 1:10 to ensure that the

variable coded as a random effect (spatially aggregated cells) contained at least one

point for the analysis. We included an additional quadratic term for depth (bathymetry)

because although whale sharks spend most of their time at the surface (Sleeman et

al., 2010b) they frequently make sub-surface ‘dives’ (Wilson et al., 2006). Including

this quadratic term tests the hypothesis that sightings occur mostly in locations where

depth is within a specific range. Moreover, depth can also be viewed as a proxy for a

set of environmental conditions differentiating the conditions encountered in shallow

versus deep habitats (e.g., upwelling and primary productivity, hydrodynamics), which

could play a role in influencing whale shark occurrence at the surface. We also

standardized some explanatory variables used in Sequeira et al. (2012) to stabilize

parameter estimation within the lmer function: we centred depth and mean sea

surface temperature, and log-transformed mean chlorophyll a, distance to shore, slope

Page 235: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

237

and effort (i.e., autumn mean value) (Table C1). In this model update, we have

included effort (autumn mean value) as an offset term to account directly for the

proportional increase in sightings with proportional increases in spatial fishing effort.

Table C1: Treatment given to each explanatory variable (standardization) used to obtained the spatial predictor (SpatialP) (Sequeira et al., 2012) for the first step of our temporal models; SST – sea surface temperature; Chl a – chlorophyll a.

Variable Treatment Range Mean depth centred -3754.8 – 2444.4 0.00 slope log -5.21 – 3.43 -1.25 distance to shore

log 0.10 – 14.17 12.61

mean SST* centred -3.97 – 2.40 0.00 meean Chl a log -2.63 – 2.91 -1.93 effort log (used as an

offset) -0.18 – 7.56 3.30

Page 236: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

238

Figu

re C

1: P

urse

-sein

e fish

ing

effo

rt (in

day

s) d

urin

g au

tum

n fo

r eac

h ye

ar fr

om 19

91 to

2007

.

Page 237: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

239

Figure C2: Habitat suitability of whale sharks (Rhincodon typus Smith 1828) in the Indian Ocean during autumn. Updated model from Sequeira et al. 2012 (see supplementary information in Appendix C for details). Areas with higher probability of whale shark sightings are similar to the ones previously described and show that mostly the western Indian Ocean – specifically the Mozambique Channel – is more suitable for whale sharks in autumn, although lower probabilities (around 0.3) were obtained when compared to Sequeira et al. (2012). Black dots represent known occurrences (Rowat, 2007), with size representing strength of occurrence (i.e., larger dots correspond to a higher number of expected sightings).

Page 238: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

240

Figure C3: Correlation between climatic predictor variables. Top panel: Scatter plot of matrices, showing: below the diagonal – the bivariate scatter plots with linear fits and correlation ellipses, and above the diagonal – the Pearson’s correlation results (top) and Spearman’s �� (bottom).

Page 239: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

241

Figure C4: Partial effect of the Indian Ocean Dipole (IOD) on the probability of whale shark sightings.

Log-odds Probability

IOD

Ti

me

(with

IOD

)

Top row: Partial effect of the Indian Ocean Dipole (IOD) (Saaji et al. 1999) on the probability of whale shark sightings. IOD reflects anomalies in the sea surface temperature in the equatorial area of the Indian Ocean between 50 – 70 °E / 10 °N – 10 °S and 90 – 110 °E / 0 °N – 10 °S). Bottom row: Partial effect of Time after accounting for climatic contributions to whale shark sighting probability. Dotted lines represent 95 % confidence interval.

Page 240: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

242

Prevalence test (down-weighting pseudo-absences)

We have re-run the temporal models after down-weighting the 100 absences (from 1

to 0.01) to get a 1:1 ratio of presences to absences (i.e., a presence:pseudo-absence

ratio of 0.5), as suggested. Because the partial effect of time (Figure 16, in Chapter IV)

is highly dependent on prevalence, the change in the presence to absence ratio

resulted in a much more pronounced curve reflecting the dependency of occurrence

with time. This is depicted in Figure C5.

a) b)

Figure C5: Partial effect of time on whale shark presence, using a presence:pseudo-absence ratio of 0.5 (prevalence). (a) on the log-odds scale showing the rate of change and (b) on the probability scale, showing the effect of time on the probability of whale shark presence. Dotted lines represent 95 % confidence interval.

Although this change in prevalence increased the estimated percentage of

deviance explained (71.1 % for the best ranked model), it did not affect the model

ranking based on wAICc (Table 7, in Chapter IV).

Page 241: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

243

APPENDIX D – Predicting current and future global distributions of

whale sharks

Figure D1: Global sea surface temperature averages from April to June. Top panel: current temperatures as seasonal averages for 2002 – 2012 derived from MODIS-Aqua satellite. Bottom panel: increase in sea surface temperatures predicted by 2070. The climate change scenario for 2070 was obtained by adding the predicted increase calculated by an ensemble forecast of climate change derived from five global circulation models: CCSM-3, MRI-CGCM2.3.2, ECHAM5/MPI-OM, MIROC3.2 (hires) and UKMO-HadCM3. These GCM were chosen based on their global skill in predicting recent SST temperatures (1981 – 2000). Model terminology follows CMIP3 Multi-Model Dataset Archive naming convention (www-pcmdi.llnl.gov/ipcc/about_ipcc.php), and baseline observed temperatures used were from 1995 (AVHRR Pathfinder).

Page 242: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

244

BIOMOD – Global model

In the preliminary run in the BIOMOD platform, the models were set according to the

summary shown in Table D1, and details on the top-ranked model according to true

skill statistics (TSS) and receiver operator curve (ROC) are shown in Table D2. The

variable importance ranking for each model considered in BIOMOD is detailed in Table

D3.

Table D1: BIOMOD modelling summary for the runs with the global dataset. Modeling whale shark distribution in the three major oceans using the dataset described in Chapter VI. Due to the complexity of the dataset and the computing time associated when running multiple models with large datasets, the preliminary run was made with only one evaluation run and one repetition of pseudo-absences selection. Fishing effort (logEff) was included in this run as a ‘weight’. ANN – artificial neural networks; CTA – classification tree analysis; GAM – generalised additive models; GBM – generalised boosted models (or BRT – boosted regression trees); GLM – generalised linear models; MARS – multivariate adaptive regression splines; FDA – flexible discriminate analysis; RF – random forest.

Number of species modelled 1 (whale shark) Numerical variables Chla_av.ln

Dist2Shore.ln Depth.c Slope.ln SST_av SST_sd

Factorial variables Ocean Number of evaluation repetitions 1 Number of pseudo-absences runs 1 Models selected ANN, CTA, GAM, GBM, GLM, FDA, RF (MARS and

SRE were automatically excluded due to the inclusion of a factorial variable)

Total number of model runs 14

Page 243: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

245

Table D2: Model selection in BIOMOD According to true skill statistic (TSS) and receiver operator curve (ROC) for the pseudo-absence dataset generated and the one replicate considered (Rep). Random forests were not considered for this classification, and GBM (generalised boosted models) were ranked higher.

Best Model

Cross-validation

Independent data

Total score

Cut off

Sensitivity Specificity

PA1 TSS GBM 0.732 - 0.742 559.3 85.2 89.0

ROC GBM 0.925 - 0.939 519.9 86.7 86.8

PA1 Rep

TSS GBM 0.732 - 0.739 535.1 85.7 88.2

ROC GBM 0.925 - 0.938 506.5 86.7 86.7

Table D3: Ranking of variable importance for the global model considered within the BIOMOD platform. Effort resulted in the variables with highest importance in all models, as expected. Note that this variable could not have been included as an offset in BIOMOD. The models developed throughout the thesis aim at understanding which variables (other than fishing effort) can be used to explain the occurrence of whale sharks. Following effort, depth, SST_av and Chl a, were the variables with the highest importance. ANN – artificial neural networks; CTA – classification tree analysis; GAM – generalised additive models; GBM – generalised boosted models (or BRT – boosted regression trees); GLM – generalised linear models; MARS – multivariate adaptive regression splines; FDA – flexible discriminate analysis; RF – random forest. Chl a – mean concentration of chlorophyll a; Shore – distance to shore (logarithm); Depth – centred depth; Slope – logarithm of slope; SST_av – mean sea surface temperature; SST_sd – sea surface temperature standard deviation.

Models Ocean Chl a Shore Depth Slope SST_av SST_sd ANN 0.364 0.704 0.363 0.187 0.045 0.682 0.284

CTA 0.000 0.472 0.425 0.193 0.005 0.491 0.000

GAM 0.101 0.053 0.424 0.093 0.047 0.455 0.002

GBM 0.001 0.251 0.376 0.076 0.001 0.244 0.000

GLM 0.161 0.121 0.377 0.141 0.059 0.620 1.313

FDA 0.000 0.178 0.394 0.141 0.059 0.620 1.313

RF 0.012 0.427 0.331 0.191 0.046 0.131 0.135

The GLM formula considered within the BIOMOD platform was:

Presence ~ Dist2Shore.ln3 + Dist2Shore.ln + Chla_av.ln3 + Depth.c +

Depth.c3 + SST_sd3 + Ocean + Slope.ln + Slope.ln3 + Slope.ln2 +

Page 244: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

246

Chla_av.ln2 + Depth.c2 + Chla_av.ln + SST_sd2 + SST_av.c +

Dist2Shore.ln2

where: Chla_av.ln – logarith of mean concentration of chlorophyll a; Dist2Shore –

distance to shore (logarithm); Depth.c – centred depth; Slope.ln – logarithm of slope;

SST_av.c – centred mean sea surface temperature; SST_sd – sea surface

temperature standard deviation. Effort (logEff) was included as a weight in this

preliminary run.

The BIOMOD ensemble result for the preliminary model run is depicted in Figure D2,

showing good agreement with the map obtained in Chapter VI (for the area covered by

the tuna-purse seine fisheries in the three oceans), with similar areas with habitat

suitability.

Figure D2: Comparison of GLMM and BIOMOD ensemble prediction for the area covered by the tuna fisheries (mean probability). The ensemble result is calculated by averaging across all the models used in the BIO)MOD run.

Page 245: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

247

ADDENDUM – Thesis amendments

Outcome of thesis examination

Letter from Examiner 1

Page 246: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

248

Page 247: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

249

Page 248: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

250

Page 249: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

251

Letter from Examiner 2

Page 250: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

252

Page 251: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

253

Page 252: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

254

Page 253: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

255

Reply to examiners

GENERAL COMMENTS:

Examiner 1

… This thesis focuses on tackling various aspects of this challenge; here the

candidate does so masterfully, employing multiple modelling approaches and

carefully comparing outcomes. I strongly recommend the candidate adopt this

approach for Chapters 4 to 6, and provide more detailed comments to this effect

below.

Conceptually, however, the thesis is rather weak. I would have liked to see the

candidate link her research to the relevant ecological theory and clearly

explicate her hypotheses relating to each model covariate. Much of the thesis is

predicated on: 1) the idea that the whale shark may be comprised of a single

global meta-population, with a sub-population or multiple subpopulations in

each ocean basin, yet the candidate does not draw upon (or demonstrate

knowledge) of the vast and relevant ecological theory pertaining to

metapopulations; 2) the idea that climate change will significantly impact whale

sharks by altering their distributions. Here, I would have liked to see a deeper

discussion of the potential implications of climate change, including the

potential for this species to utilize different depths, the potential for altered

predator-prey dynamics, and consideration of the possibility of beneficial

effects of climate change (not all effects will be negative!). The thesis chapters

which are not yet published would benefit from these additions.

Page 254: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

256

I suggest changing the overall thesis title to:

'Regional and Global Distribution Models for Whale Sharks

(Assessing Occurrence Trends of a Highly Migratory Marine Species'

or

'Ocean-basin Scale and Global Distribution Models for Whale Sharks

(Assessing Occurrence Trends of a Highly Migratory Marine Species)'

to more accurately reflect the thesis contents; only Chapters 2 and 6 are global

in scale.”

The thesis is reasonably well-written, but would benefit from a careful final

proofreading and grammatical check throughout by a native English speaker.

Mistakes are minor but persistent. I have noted a few of these below in my

detailed comments.

The title reflects the main aim of this PhD, which is conceptualised in Chapter II, tested

throughout Chapters III-V in different oceans, and achieved in Chapter VI. For this

reason I think the current title (below) is more adequate.

'Global distribution models for whale sharks: assessing occurrence trends of

highly migratory marine species'.

I address the other suggested amendments within each chapter as specified below.

Page 255: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

257

SUMMARY

Examiner 1

remove 'temporal' from first line, (or add 'spatial),

Changed.

suggest to reword 'life cycle' here and throughout; sounds awkward. E.g. 'life

stages'

Changed.

Page 256: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

258

CHAPTER I

Examiner 1

The first paragraph is unnecessarily broad. lnstead, you might lead with some of

the relevant ecological theory. I would definitely like to see some discussion of

the ecological theory, questions and hypotheses before diving into the methods

(even in a methods driven thesis);

I have now included some of the relevant ecological theories – namely spatial ecology

and metapopulation theories in the General Introduction section. However, I kept the

first paragraph because it relates to the general background (i.e., general importance

of the oceans) putting the relevance of my thesis into perspective. This paragraph is

important to set the stage for my thesis within the major role of the oceans and ocean

life, and for bringing to evidence the lack of knowledge about most marine species

driven mainly from sampling limitations. This assists the reader in understanding the

difficulties of studying a pan-oceanic shark species, and the limitations associated with

the variables that can be used while designing distribution models for such species.

I highlight the changes to this section in the new text below:

“...However, this technology applies only to the sea surface, and so most of the

ocean is still largely unsampled. The latter also applies to sampling marine

species from higher trophic levels, which still relies on expedition surveys. In

spatial ecology (Tilman & Kareiva, 1997), the distribution of species can be

interpreted as a match between their ecological requirements and the

environmental conditions available. Therefore, the difficulties of sampling the

marine environment (through spatial surveys) hamper detection of the relevant

spatial coverage to interpret these associations correctly. Modelling approaches

Page 257: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

259

are used to assist the understanding of the relationships between fundamental

biological conditions (or processes) and species occurrence - as I discuss

below.

The distribution of marine populations is also strongly associated with

movement, whether through migration (e.g., herring; Dickey-Collas et al.,

2009)or through larval dispersal (e.g., reef fishes; Jones et al., 2009). This leads

many marine ecology studies to embrace and test metapopulation concepts and

hypotheses (Hanski, 1998; Roughgarden et al., 1985), even if not specifically

acknowledged. Metapopulation dynamics are the processes associated with

population turnover, i.e., extinction versus colonisation (based on Levin’s model

from 1969), between patchily distributed subpopulations with some degree of

connectivity between them (Hanski & Gilpin, 1991). In terrestrial systems, this

theory has been widely applied to a range of species including butterflies

(Hanski, 1994), rabbits (Nielsen et al., 2008), rats (O’Brien et al., 2008) and

bears (Hellgren et al., 2005). Indeed, fragmentation of terrestrial habitats makes

the application of applied metapopulation theory highly relevant. In the marine

environment, metapopulation concepts have been applied mostly to

invertebrates and reef fishes (Sale & Kritzer, 2003); however, testing

hypotheses at broader spatial scales might not be suitable or even practical for

many marine species that are either highly migratory or widely dispersing

(Kritzer & Sale, 2006).”

and:

“...Despite the importance of this taxon, the basic biology and ecology of most

sharks are still poorly quantified. For example, the economically important

(tourism, fishing) whale shark – is still largely unknown. There is some evidence

Page 258: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

260

that whale sharks can travel across oceans, which raises the possibility of

applying concepts and testing hypotheses associated with metapopulation

theory. Depending on true connectivity (currently unquantified) of whale shark

sub-populations worldwide, they might actually be part of a unique global meta-

population (Sequeira et al., 2013), such that localised perturbations might have

detrimental flow-on effects to other sub-populations in distant waters.”

p.28, 2nd line -'and might change' should be 'how it might change'

Changed.

p. 31 - tourism industry - lf you mean whale sharks, then be specific. Otherwise

this is incorrect i.e. do not generalize about 400+ shark species, when only a

handful are relevant to any tourist industry;

Agreed and changed to:

“Appropriate management of shark populations is also a commercially lucrative

ambition because of the high revenue some shark species bring to the tourism

industry.”

p. 31 - References here seem a bit dated (e.g. see recent report by Jessica

Meeuwig's group on the value of sharks to tourism (from Palau?), as well as a

key one on shark life histories by Luis Lucifora, Veronica Garcia and Jeffrey

Hutchings from 2012 Ecology;

The suggested references were now added.

Page 259: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

261

“Due to their slow life histories (delayed maturity, long life spans, few offspring),

sharks are particularly vulnerable to exploitation (Baum et al., 2003; Field et al.,

2009; García et al., 2007; Hutchings et al., 2012; Musick et al., 2000; Schindler

et al., 2002). Consequently, many recent management strategies, increasingly

reliant on good biological information and model projections, have been devised

to maximise the long-term persistence of shark populations in the face of these

pressures (Barker & Schluessel, 2005).

Appropriate management of shark populations is also a commercially lucrative

ambition because of the high revenue some shark species bring to the tourism

industry (Vianna et al., 2012). Sharks are often the subject of major tourism

ventures, such as shark watching and diving (Anderson & Waheed, 2001;

Brunnschweiler, 2010; Catlin & Jones, 2009), and some have argued that

certain species are worth more alive than dead – a recurrent source of profit

instead of a one-off income (Topelko & Dearden, 2009; Vianna et al., 2012).”

p.31/32 - Does Essington et al 2002 really demonstrates that sharks play a

keystone role (this is not what I recall from the pair of Essington / Schindler CNP

Ecopath models). Other more recent refs you could cite here?

Agreed. Essington et al. (2002) report that sharks play a key role, but do not describe

anything about this process. References from studies on cascading fishing effects

directly describe how the food web structure and ecosystem dynamics are affected by

removal of predatory fish. I have now included the references Scheffer et al. (2005) on

the consequences of the elimination of large predatory fish, and Myers et al. (2007) on

the consequences of apex predatory sharks loss. I have also included another

Page 260: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

262

reference about modelling possible indirect effects of shark removal from ecosystems

(Frid et al., 2008). I changed the references used in the sentence the examiner

highlighted to:

“Moreover, sharks are also thought to play a key role in controlling the food web

structure and the dynamics of marine ecosystems (Frid et al., 2008; Myers et

al., 2007; Scheffer et al., 2005).”

p.32 'study case' should be 'case study' throughout thesis (e.g also on p. 38 etc)

Changed.

p.32 - The statement that poikilotherms are a challenging case study is weakly

argued. Many such species are the subject of climate modeling studies. Could

you not draw upon some knowledge from the terrestrial literature to help inform

this statement?

This text now reads:

“For example, the economically important (tourism, fishing) whale shark – the

largest fish species – is still largely unstudied. There is some evidence that

whale sharks can travel across oceans, which raises the possibility of applying

concepts and testing hypotheses associated with metapopulation theory.

Depending on true connectivity (currently unquantified) of whale shark sub-

populations worldwide, they might actually be part of a unique global meta-

population, such that localised perturbations might have detrimental flow-on

effects to other sub-populations in distant waters. Whale shark distribution is

also strongly associated with temperature; they are generally found in warm and

Page 261: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

263

temperate seas (Last & Stevens, 2009) and spend most of their time at the

surface (Gunn et al., 1999; Sleeman et al., 2010a; Sleeman et al., 2010b;

Wilson et al., 2006) within a narrow range of temperatures (Sequeira et al.,

2012). This limited temperature range suggests that their time in deeper waters

is limited by thermoregulatory constraints (Thums et al., 2012) or minimises the

time spent in deep, oxygen-poor water (Graham et al., 2006). Changes in water

temperature associated with climate change (Hoegh-Guldberg & Bruno, 2010),

are thus likely to change whale shark distribution or abundance as they re-

equilibrate to new temperature regimes and adapt to changing prey distribution

(Gutierrez et al., 2008).

Being filter feeders, whale sharks are also likely to be affected by the expected

climate-driven changes in ocean productivity and food-web dynamics (Hoegh-

Guldberg & Bruno, 2010). Plankton distribution is expected to shift and its

abundance will likely change in response to the thermal properties of changing

oceans (Reygondeau & Beaugrand, 2011). This logically suggest that such

changes will propagate through to higher trophic levels (Kirby & Beaugrand,

2009). This led Hayes et al. (2005) to refer to plankton monitoring programmes

as “sentinels” to identify changes in marine ecosystems. Whale sharks have the

capacity to travel large distances, so altered plankton distribution will repeatedly

affect whale shark migration patterns (Chin et al., 2010). Moreover, whale

sharks aggregations are usually associated with specific productivity events,

such as coral spawn in Ningaloo, Western Australia (Taylor, 1996) that

promotes an increase in zooplankton production. Changes in the timing of these

events, as well as other impacts of climate change on coral reefs (e.g., habitat

Page 262: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

264

destruction within whale shark feeding areas), might therefore also impact

whale shark occurrence (Chin et al., 2010).

Despite whale sharks spending the majority of their time at the surface (Gunn

et al., 1999; Rowat et al., 2007; Thums et al., 2012), they can also dive deeply.

This opens speculation to the potential for this species to use different depths

(where water is cooler) as surface water temperatures increase. However,

oxygen limitation might pose a problem for extended time spent in deeper

waters (Pörtner, 2001; Prince & Goodyear, 2006). For most of the reasons

given above, whale sharks have been referred to “as potentially the most

vulnerable species [to climate change] in the pelagic ecological group” (Chin et

al., 2010; in their assessment of the vulnerability of sharks and rays in the Great

Barrier Reef), and present a challenging case study in the context of climate

change and expected shifts in marine species distribution.”

p. 35 top paragraph: Please clarify here why it is that tuna aggregate near whale

sharks - because they share prey or because the tuna prey upon species with

which the whale sharks share prey?

I have changed this section to:

“Aggregations of whale sharks usually coincide with higher concentrations of

plankton in surrounding waters (Heyman et al., 2001; Sleeman et al., 2007).

Being filter feeders, they generally only consume small secondary grazers (e.g.,

krill or crab larvae; small fish) (Meekan et al., 2009; Wilson & Newbound, 2001),

and therefore other species are commonly associated with them, including

commercially valuable anchovy (Duffy, 2002) and herring (Wilson, 2002). This

Page 263: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

265

characteristic has been particularly helpful in oceanic tuna fisheries where

whale sharks are used to spot tuna (Matsunaga et al., 2003) commonly

aggregating under the shark (possibly foraging on similar prey) – in other words,

whale sharks are used as fish aggregating devices.”

p. 35 'economic importance worldwide' This is an overstatement. I think you

mean in 'select coastal locations around the world'

It now reads:

“Due to their predictable aggregation behaviour, together with their harmless

characteristics, whale sharks have become economically important in several

coastal locations around the world.”

p. 36 'in the near future' is this specifically stated in the IUCN Red List

Guidelines? lf so, cite them. lf not, and you infer it, then please remove.

It now reads:

“Since 1990, whale sharks have been listed on the IUCN Red List of

Threatened Species (IUCN, 2010), and since 2000 are classified as Vulnerable

(i.e., at risk of extinction in the wild).”

p. 38 Strikes me as odd to refer to the world's largest fish as 'elusive'

Despite being the largest fish, whale shark patterns of occurrence near shore are not

defined, and are still largely unknown in the open ocean, i.e., they are elusive. This

term has been used in reference to whale sharks in several publications: Taylor, 2007;

Page 264: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

266

Riley et al., 2010; Rowat & Brooks, 2012; Sequeira et al., 2012; Eckert and Stewart,

2001. Therefore, I believe the reference to whale shark as ‘elusive’ is appropriate in

the following paragraph:

“Besides presenting a challenging case study in the context of climate change

and likely shifts in marine species distribution, the elusive whale shark also

represents a good model to understand biotic responses to oceanographic

patterns; being a filter feeder, they should track primary production more closely

than higher-order predators (Grémillet et al., 2008).”

Page 265: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

267

CHAPTER II

Examiner 1

Note: This chapter is now published in the Journal of Fish Biology, 2013 (82:2, 595-

617), and therefore any changes made in the thesis based on examiner’s reports

cannot be included in the published manuscript. Also, due to the journal requirements,

the format of the text presented in the article (e.g., use of passive voice) differs from

the contents in the thesis chapter.

Abstract, p. 45 -half way down: "generates a model describing... are part of a

single, global meta-population', change to "....might be part of ...."

Changed.

lntroduction p. 46 -'Being large and docile....' Sentence does not follow logically

from this beginning.

Should now read:

“The large and docile whale shark (Rhincodon typus, Smith 1828) is a filter-

feeding chondrichthyan that can reach over 18 m in total length (Borrell et al.,

2011; Chen et al., 1997; Compagno, 2001), making it the largest extant fish

species. Its geographic range is more less known across all tropical and warm

temperate waters...”

This chapter provides an interesting overview of whale shark research. I don't,

however, feel that here (e.g. pages 71-72) or elsewhere in the thesis (Chapter VI)

Page 266: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

268

you have made a convincing argument as to why we should be particularly

concerned about the potential impacts of climate change on this wide-ranging,

highly mobile marine species.

I have made changes both in the General Introduction (see answer above to the

examiner’s similar comment re. p.32) and in the (not-yet-published) Chapter VI

sections of the thesis to include a more convincing argument about how whale sharks

will be impacted by climate change, and why this is of major concern if we are to

conserve this species in the long term. Chapter I is currently published (Journal of Fish

Biology, 82:2, 597-617), and because the emphasis of this chapter is on the potential

of whale sharks to form a unique global population, and not on the importance of

whale sharks as a case study for impacts of climate change, these changes were not

included in this chapter.

p. 48 - life cycle (Do you mean e.g. juveniles vs. adults) - think you mean 'life

stages'

Changed.

p.48/49 - I'm also surprised by the argument that tagging data couldn't be used

to estimate 'broad-scale migratory patterns' in this species. This is exactly what

the TOPP (www.topp.org) program achieved over the past decade + for many

large, highly mobile marine species.

I state, in the pages referred by the examiner, that tracking methods can provide

insight into the movement behaviour of whale sharks; therefore, they can be used to

estimate ‘broad-scale migratory patterns’. However, in the context of the work being

Page 267: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

269

presented in this chapter (entitled “Inferred global connectivity of whale shark

populations”), I detail that due to the difficulties associated with collecting long-term

and wide-coverage tagging data for whale sharks, these methods are unlikely to

gather data encompassing the full (global) range of the species’ distribution and

greatest individual movements:

“Examining whale sharks at the scale of single aggregations cannot adequately

describe the species’ life history because it encapsulates only a small proportion

of the life stages. Collecting data outside aggregation areas is therefore

essential. Continual developments in tagging technology to improve estimates

of home range size, movement patterns and habitat use (Hammerschlag et al.,

2011) have been partially successful in this regard, although despite > 3000

whale shark tagged to date (e.g., www.whaleshark.org), trajectories have not

revealed reliable evidence for inter-aggregation connection. This is not entirely

surprising given the low probability of resighting migratory individuals in widely

spaced aggregations. The lack of bio-logging efficiency is mostly due to

premature detachment and limited spatial coverage of acquired data (e.g.,

Brunnschweiler et al., 2009), tag removal and/or damage (Fitzpatrick et al.,

2006; Hays et al., 2007) which might result in part from attacks from other

sharks or killer whales (Fitzpatrick et al., 2006; Speed et al., 2008), and the

accumulation of bio-fouling organisms causing tag malfunction (Hays et al.,

2007). Moreover, even though tracking methods can provide some insight into

the movement behaviour of whale sharks, they are still unlikely to encompass

the full range of the population’s distribution.”

Page 268: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

270

The TOPP program was a multinational collaborative effort aiming at explaining the

dynamics of marine predators occurring in the North West coast of America, as part of

the Census of Marine Life. It received funding from several organisations what allowed

the deployment of several thousands of tags, of which less than half resulted in

individual tracks (Block et al., 2011). Also, despite the long tracks obtained for some of

the species, most of the temporal coverage of the individual tracks was less than one

year (maximum duration for different species ranged from 19 to 358 days) (Block et

al., 2011 supplementary material). So, the same difficulties I highlighted in this chapter

regarding collection/recovery of tag data were also highlighted in the study by Block et

al. (2011). It was the amount of both collaborative effort and funding (together with the

duration of the project itself – 10 years), that allowed TOPP to become a successful

program. I have now included a reference to this program in the thesis General

Discussion:

“Tagging approaches can provide insight into the movement behaviour of

migratory species, however they bring their own problems (e.g., Brunnschweiler

et al., 2009). Even in the successful TOPP program (tagging of Pacific

predators; Block et al., 2011), despite the long tracks obtained for some of the

species, most of the temporal coverage of the individual tracks was less than

one year (maximum duration for different species ranged from 19 to 358 days)

(Block et al., 2011 supplementary material). This multinational, collaborative

program (part of the Census of Marine Life) received funding from several

organisations to allow the deployment of several thousands of tags, of which

only less than half resulted in individual tracks (Block et al., 2011). However,

some of the problems associated with tagging procedures can be improved as

technology progresses. Some suggestions to maximise the duration of tracking

Page 269: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

271

have already been made, and should be adopted when planning new tagging

studies. These suggestions include using fast-loc GPS technology (Hays et al.,

2007), applying anti-fouling paint (Hammerschlag et al., 2011), and use copper

salt-water switches (Hays et al., 2007). But improvements to prevent premature

detachment (Brunnschweiler et al., 2009) are also needed, and new attachment

options should be considered. To maximise the potential for measuring long-

distance migrations, I suggest that tagging studies target multiple sharks

simultaneously within the same aggregation location (cf. Eckert & Stewart,

2001; Wilson et al., 2006), and that deployments are made close to the end of

the peak aggregation season. Additional tagging studies on whale sharks

released from pelagic fisheries nets would also provide information on oceanic

movements to complement results obtained from known aggregations.”

Despite whale sharks occurring mostly near the shore of developing nations (i.e.,

funding mainly provided via investment from other countries), > 3000 tags have been

deployed on whale sharks, but only a few resulted in usable track data (Sequeira et

al., 2013; i.e., this chapter). This confirms that using the same approach to study

whale sharks (and the same success rate in recovering data) would make it unlikely

that tracking data will provide insight into the movement behaviour of whale sharks

encompassing the full range of their global distribution.

Page 270: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

272

CHAPTER III

Examiner2

Note: This chapter is published in Diversity and Distributions, 2012 (18, 504-519), and

therefore any changes made in the thesis based on examiner’s reports were not

included in the published manuscript.

I do not have much to say given the paper is already published in a respectful

journal. The literature cover is relatively good although some key papers are

ignored.

The methods are not very well-explained to me. There are even some

misconceptions. MaxEnt is NOT a presence-only algorithm, not more than any

GLM for instance. I am not entirely convinced by the classification of models as

well. MaxEnt and BRT are not a machine-learning method. BRT are only

classification trees in a boosting framework.

Even though I referred to MaxEnt as a presence-only model in the Introduction of this

chapter, in the Methods I specified the inclusion of background data, and have

discussed different ways to include background data in the discussion section. In

Chapter IV, I have also corrected the MaxEnt description by including the need for

background data in its definition:

“MaxEnt is a tool for generating species distribution models from presences-

only data. This modelling tool uses covariate data from species presence

locations and background sampling to estimate habitat suitability for the species

occurrence (for a detailed statistical explanation of MaxEnt, see Elith et al.,

2011).”

Page 271: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

273

As for MaxEnt being a machine-learning method, my reference to this was in Chapter I

(General Introduction), and was followed with a reference to Phillips et al. (2004) -

where they present MaxEnt’s value for species distribution modelling contextualising it

within the machine-learning community. I could not find my reference to BRT as a

machine-learning method or as classification trees.

Why using Kappa for GLMs? The use of Kappa should be stopped now because

it is strongly biased with prevalence, which is definitely the case here given the

use of pseudo-absence. True Skill Statistics or Boyce Index should be preferred.

To account for the fact that Kappa is strongly biased with prevalence, I used the

observed prevalence (presence:pseudo-absence ratio) as the value to optimise the

threshold used in the cmx function (from the package {PresenceAbsence}Bo) to

estimate Kappa. This I discussed in Chapter VII (General Discussion) as shown below:

“Another related issue is the decision regarding which probability value should

be used to convert the model results into presences or absences (Jiménez-

Valverde & Lobo, 2007; Liu et al., 2005). This value is commonly assumed to be

0.5, but for presence: absence ratios different to 1:1, this assumption results in

incorrect classification of absences (if prevalence is low, i.e., more absences

than presences) or presences (if prevalence is high) (Jiménez-Valverde & Lobo,

2007). Several methods can be used to specify this threshold (see Liu et al.,

2005), such as averaging the predicted probabilities, using sensitivity and

specificity, the prevalence value, or the value that maximises �. I considered

this threshold equal to the prevalence used, which is considered a good

approach (Liu et al., 2005).”

Page 272: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

274

Also, within the function optimal.thresholds() from the R package {PresenceAbsence},

the threshold that maximises the sum of sensitivity and specificity (equivalent to True

Skill Statistics; Freeman, 2012) could have been chosen (instead of observed

prevalence). However, while being independent from prevalence, this index can over-

predict the species’ distribution (Freeman & Moisen, 2008). I have now calculated the

Boyce index (Boyce et al., 2002) – this is shown below (answer to comment on page

104). Results are similar for all pseudo-absences techniques (especially for MaxEnt),

and are generally higher (0.83 - 0.95) than the ones obtained with kappa (0.34 - 0.79).

There is no justification on the use of MaxEnt or GLM? Why these two

approaches? Only because the authors like GLM and because MaxEnt is over-

used? Why not including an autoregressive component to MaxEnt?

The choice of GLM derived from their ability to analyse non-Gaussian data, their

flexibility and the relative ease with which relationships can be identified and

hypotheses tested using the linear imposition. GLM can also be used in a hypothesis-

testing framework using AIC and BIC as indices for model support (Burnham &

Anderson, 2004). Moreover, this type of regression model generally results in more

realistic projections (Brotons et al., 2004; Zaniewski et al., 2002). But no modelling

approach is infallible, so we compared GLM results to MaxEnt, which deals with

presence-only data (and background information) (Elith et al., 2011), and has been

increasingly applied to model species distribution. Following the examiner’s

comments, I have re-run the models within a BIOMOD framework (Thuiller, 2003) to

check which modelling approach performed better – I give some of these details

below, answering the examiner’s comment on pages 105-106.

Page 273: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

275

Inclusion of an autoregressive component which describes time-varying processes

could possibly assist to describe lags between whale shark occurrence and the

predictors used. In relation to my study, this could be particularly useful in relation to

chlorophyll a. However, due to the little variation in the concentration of the latter

predictor (in open ocean) I believe it would be difficult to obtain a result. In the sense

that autoregressive components represent random processes, some of these were

included in the MaxEnt models in two ways. First, by using different pseudo-absences

sets generated from the background data using each of the three methods tested

(random, IDW and tuna), and by including a random 10000 background cells

(maximum allowed in MaxEnt) from the entire area considered in the models. Second,

the type of variables considered for modelling was also varied by specifying the

“features” allowed in MaxEnt runs – i.e., through the selection of “linear and quadratic

features” or the use of “auto-features”. Moreover, the work presented in this chapter

was further developed in the Chapter IV to include an analysis of temporal trends in

whale shark occurrence.

I do not understand why the ROC curve is not estimated for GLM/GLMM instead

of keeping only with the deviance. It would have been easier to compare with

MaxEnt.

Following the examiner’s suggestion, I have now calculated the ROC/AUC value for

the GLM/GLMM models used in Chapter III, and I have included a table with results in

Appendix B (Table B2, copied below) together with the following text:

“Receiving operator characteristic curve / area under the cure (ROC/AUC)

Page 274: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

276

To measure models performance we also calculated the receiving operator

characteristic curve / area under the curve (ROC/AUC) and results are shown in

Table B2. Thetrend arising when comparing the AUC results between each

MaxEnt and GLM model is similar to the one I obtained previously when

comparing � – i.e., when using the random technique to generate pseudo-

absences, GLM was scored higher by both measures (AUC and �).

Using ROC/AUC to measure performance of species distribution models has been

criticised for being influenced by the total extent to which the models are applied (Lobo

et al., 2007), and especially when the objective of the models is to determine the

potential distribution of species (Jiménez-Valverde, 2011). Therefore, to compare

MaxEnt, GLM and GLMM results, I have used the ��statistic (accounting for the fact

that this measure depends on prevalence), which some authors suggest (Manel et al.,

2001). The percentage of deviance explained was only used for GLM and GLMM.

Comparison of the Area Under the Curve (AUC) and Kappa results for the two modelling approaches used in chapter III: MAxEnt and GLM/GLMM for each of the pseudo-absence dataset generated method (P/A) and for each season: Autumn (Aut), Winter (Win), Spring (Spr) and Summer (Sum).

P/A Approach Aut Win Spr Sum

Rand

om

MaxE

nt Kappa 0.54 0.57 0.50 0.49

AUC 0.72 0.70 0.69 0.67

GLM

Kappa 0.79 0.51 0.59 0.51

AUC 0.87 0.61 0.79 0.63

IDW

MaxE

nt Kappa 0.34 0.10 0.25 0.31

AUC 0.63 0.57 0.63 0.61

GLM

Kappa 0.35 0 0.33 0

AUC 0.82 0.72 0.80 0.74

Page 275: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

277

Tuna

MaxE

nt Kappa 0.64 0.37 0.48 0.43

AUC 0.71 0.67 0.67 0.68

GLM

Kappa 0.68 0.47 0.51 0.56

AUC 0.90 0.75 0.79 0.77

Explained deviance measure the explanatory power of a model, not is predictive

accuracy. This is a quite different.

This was now changed to:

“We used the percentage of deviance explained (De) to quantify each model’s

explanatory power.”

There is a problem of rows (the third row) in Table 5

I have now corrected the problem with the rows in Table 5.

Discussion: end of Page 104: The Boyce Index allows to discount for pseudo-

absence and to compare different models or repetitions with the same set of

presence.

I have now calculated the Boyce index for the three pseudo-absence generation

methods for each modelling tool – MaxEnt and GLM (Autumn runs only; Table B3).The

Boyce index assesses the monotonic relationship (Spearman’s rank correlation)

between habitat suitability classes and the proportion of evaluation points adjusted by

the area covered (Boyce et al., 2002). At the end of page 104, I discuss that “because

pseudo-absences locations vary within each model, the resulting deviance explained

Page 276: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

278

and AUC are not directly comparable across models”, and that the three methods

used to choose pseudo-absences are most useful to determine how absence locations

influence model results. According to the Boyce index calculated for autumn, the

results among pseudo-absence selection techniques and among modelling tools are

similar, showing that the predictions from all models considered are consistent with the

presences distribution in the sampled dataset. Results for the GLM predictions were

generally slightly higher. I have included a table with results in Appendix B (Table B2,

copied below) together with the following text:

“Boyce index comparison between MaxEnt and GLM/GLMM models

We assessed the monotonic relationship (Spearman’s rank correlation) between

habitat suitability classes estimated by the models and the proportion of

evaluation points adjusted by the area covered (Boyce et al., 2002) by

calculating the Boyce index (autumn only). According to this index (Table B3)

the results among pseudo-absence selection techniques and among modelling

tools are similar, showing that the predictions from all models considered are

consistent with the presences distribution in the sampled dataset. Results for

the GLM predictions were generally slightly higher.”

Boyce’s index for the three pseudo-absence generation techniques (P/A) for each modelling technique (MaxEnt and GLM) for autumn. Standard deviation was always < 0.001.

P/A Tool Boyce index

Random MaxEnt 0.827

GLM 0.845

IDW MaxEnt 0.854

GLM 0.911

Tuna MaxEnt 0.829

GLM 0.952

Page 277: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

279

Page 105: Something is missing in the discussion about a missing factor. I do

not buy the fact that spatial autocorrelation bias the models. Given most

environmental variables are spatially autocorrelated, the fact that the residuals

are also spatially auto-correlated strongly suggest a missing element or a

missing variable. What about dispersal, what about another variable? It could

also be because of interactions between factors that are not taking into account

here.

The models I developed in Chapter III (and throughout the thesis) could possibly be

improved by the addition of other predictor variables. This was discussed in the

General Discussion, where I suggest that a predictor directly describing whale shark

prey (zooplankton abundance) could improve the results of the distribution models.

There are also other important caveats; applied only to the sea surface, my models do

not consider any variables describing the vertical distribution of this species, nor do

they account for whale shark population dynamics and/or structure. However, data on

the key aspects of whale shark life history (including dispersal) are still to be

measured, and therefore no further data are currently available to add to the predictor

set.

I have, however, included an additional term (depth2) when re-fitting the

models developed in Chapter III – detailed in the supplementary information for

Chapter IV (autumn model only). This term, accounting for whale shark occurrence

mostly in locations where depth is within a specific range, did improve the model fit;

however, the residuals’ spatial autocorrelation was reduced only after the inclusion of

the spatial random effect, as in Chapter III.

Moreover, I found no support for any interactions between the predictor

variables used based on model runs made:

Page 278: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

280

i. in BIOMOD (Thuiller, 2003), none of the GLM model equations (one for

each pseudo-absence repetition) included interactions between predictors

ii. using generalised boosting models (GBM), the interact.gbm object

returned a Friedman’s statistic = 2.9 × 10-14, which indicates weak

interaction effects (Ridgeway, 2013; GBM package)

iii. using boosted regression trees (BRT; Elith et al., 2008), where even

though I found some support for an interaction between depth (centred)

and distance to shore, and between depth (centred) and mean sea

surface temperature, the coefficients for these were low (0.02 and 0.06,

respectively) and did not contribute much to the models’ explanatory

power.

I have now included a synthesis of the results obtained for these modelling

tests (with BIOMOD and GBM/BRT) in Appendix B (“BIOMOD, GBM and BRT models

for the Indian Ocean”).

Page 106: MaxEnt can use interactions between variables, this is why the set of

important variables are not the same. It comes back to my previous comment

that GLM should also include some interactions between crucial variables.

MaxEnt can use interactions between variables only when “auto-features” is selected.

In Chapter III, I ran MaxEnt models mostly using the “linear and quadratic” features to

make their results more comparable with those derived using GLM. I only selected

“auto-features” in an extra test run. I discussed the results from each MaxEnt output

within that chapter. Moreover, as discussed in response to the previous examiner’s

comment, I found no support for interactions between predictor variables.

Page 279: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

281

CHAPTER IV

Examiner2

Note: This chapter is now published in Marine Ecology Progress Series (doi:

10.3354/meps10166), and therefore, any changes made in the thesis based on

examiner’s reports were not included in the published manuscript.

Page 116-117: To artificially increase the prevalence of the species and thus the

generated probability of presence, it is generally recommended to down-weight

the absence to have the same weight than the presence.

Following the examiner’s comment, I have now re-run the temporal models with down-

weighted pseudo-absences and I have included the results in Appendix C

(“Prevalence test – down-weighting pseudo-absences”) as follows:

“We have now re-run the models after down-weighting the 100 absences (from

1 to 0.01) to get a 1:1 ratio of presences to absences (i.e., a presence:pseudo-

absence ratio of 0.5), as suggested. Because the partial effect of time (Figure

16, in Chapter IV) is highly dependent on prevalence, the change in the

presence to absence ratio resulted in a much more pronounced curve reflecting

the dependency of occurrence with time. This is depicted in Figure C5. Although

this change in prevalence increased the estimated percentage of deviance

explained (92.3 % for the best ranked model), it did not affect the model ranking

based on wAICc (Table 7, in Chapter IV).”

Page 120: there is a mistake. Year is the random factor, and not time. (modify

the order in the parenthesis). Else this is contradictory with Table 1.

Page 280: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

282

Changed. It now reads:

“We included both a fixed and a random effect for time (Time and year,

respectively) to ...”

Page 281: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

283

CHAPTER V

Examiner 1

Note: Here, I address Examiner 1’s comments made on Chapter V only. I will address

comments to Chapter VI in the following section. Several changes were made to

Chapter V – as specified below, and this chapter is now being prepared for submission

to the Journal of Experimental Marine Biology and Ecology.

These two chapters would benefit from implementation of an additional (and

different) modeling approach (e.g. models with a truncated distribution or

boosted regression trees) and/or some sensitivity analyses (e.g. re: generation

of the pseudo-absences in for the Atlantic and Pacific Ocean data sets as in

Chapter III).

Chapter V consolidates the methods used both in Chapters III and IV, but applies them

to the Atlantic and Pacific Oceans instead. Due to the complexity of the analyses, and

because the main objective was to obtain results comparable to those described in the

previous chapter (mostly in terms of the partial effect of time), I used the same

modelling approaches.

Similar to what I did in Chapter III, I included a ‘sensitivity’ analysis in each

chapter, as multiple pseudo-absence datasets were always generated prior to the

modelling to assess the influence of pseudo-absences location choice in the models

results. I have now added the following text to make this clear:

“To assess the influence of the date and location associated with the selected

pseudo-absences, we generated each set of pseudo-absences 100 times prior

Page 282: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

284

to their use in the spatial models, and 10 times for the temporal models (due to

computing time).”

In Chapter III, I have made an assessment of the influence of different pseudo-

absences generation methods (such as using random selection, distance to whale

sharks or based on tuna catch). Because the results from each pseudo-absence

generation method used in Chapter III were similar, I decided to keep using the

simpler method (random selection; as discussed in Chapter VII – see next answer) in

the following chapters. This is in accordance with the results from a recent study on

selection of pseudo-absences (Barbet-Massin et al., 2012), where the authors

recommend a random selection of pseudo-absences.

At minimum, I would like the candidate to include some discussion of the

potential biases and limitations arising from the data used and/or the modelling

approaches undertaken in these two chapters. Such admissions and

discussions would strengthen (not weaken) these chapters and their related

manuscripts intended for publication in the peer-reviewed literature. The

candidate does an excellent job of reviewing these in Chapter VII, her General

Discussion, but this chapter will not be published in the primary literature, and

this material is critical to each of Chapters 5 and 6 in order for readers to be able

to evaluate the research.

Following the examiner’s advice, I have now included in Chapter V further discussion

of the potential biases and limitations arising from the data and modelling approaches I

used, as previously discussed in Chapter VII:

Page 283: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

285

“The datasets we used were opportunistically collected and therefore lead to

inherent complications. Differential sampling effort is a major issue for which we

accounted at least partially in both modelling steps to reduce potential bias in

spatial and temporal effort. However, an underlying assumption of our models is

that failure to report a whale shark presence (whether due to a failure in

detecting a shark or in reporting a detection) is evenly distributed across the

sampling area and study period. Access to higher-resolution data would likely

improve the results (exemplified by the Atlantic Ocean models), and can be

achieved through establishing collaboration between researchers and fisheries-

management organisations. Commercial-in-confidence restrictions on access to

fisheries data (especially sensitive data such as fish catch) delay scientific

research from providing more accurate models on fisheries impacts. Such

collaborations would also allow researchers to give greater insight on how better

datasets could be collated for specific objectives using the resources currently

in place. However, while data collected by fisheries might not be as precise as

those derived from scientifically designed surveys, they are still essential

sources of information given the logistic challenges of surveying widely

distributed species over their entire range. Fisheries data therefore present

unique opportunities to advance our understanding of such species.

Because the datasets we used contained only recorded presences, the

generation of pseudo-absences was necessary to allow for the binomial

estimation in the models. Differences associated with the method to generate

pseudo-absences was assessed in a previous study using similar datasets

(Sequeira et al., 2012) where we found that model performance was not

reduced by using random pseudo-absences selection, and was therefore the

Page 284: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

286

method used in this analysis. This accords with the results from another recent

publication specifically addressing pseudo-absence selection issues (Barbet-

Massin et al., 2012). The number of pseudo-absences selected is another issue

to consider when generated. Barbet-Massin et al. (2012) reported that model

accuracy increases until the presence to pseudo-absence ratio (i.e., prevalence)

equals 0.1, and remains constant for lower ratios. In the models we developed,

prevalence was 0.1 in the spatial assessment, and 0.01 in the temporal

assessment (to allow enough points for the spatially explicit temporal analysis);

therefore in both cases, we do not expect the ratio presences:pseudo-absences

to have affected model accuracy.

Another important aspect of the input data we used is that both sightings and

environmental variables correspond mostly to the ocean surface layer. Three-

dimensional environmental data at an adequate resolution for regional analyses

such as that presented here are currently not available. Even though whale

sharks spend most time at the surface (e.g., Rowat et al., 2007) they also dive

frequently, and therefore, assessment of how they explore the vertical habitat

and estimation of its suitability at a broad scale is also of importance. However,

this will only be possible when new technologies are developed to allow vertical

data collection at broad spatial scales.

Despite the limitations associated with both the opportunistic dataset and

species distribution models, we have demonstrated a powerful way to use such

datasets to identify both spatial and temporal trends of highly migratory species.

We also revealed that longer-term and higher-resolution sightings datasets are

still required to tease apart potentially confounding aspects of inter-ocean

migration in whale sharks. Continued investigation of these connections is

Page 285: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

287

necessary to assess temporal trends in occurrence, and to reveal the suspected

impacts of human modifications to the wider marine realm.”

In each of these chapters, please also:

show the equations for the models being used, so that the reader can more

clearly see what is being done;

The equations for Chapter V, in terms of variables used, are detailed in Table 8 and

Table 9. I have now also included a generic description of the equations in the

Methods:

“A generic way to write the GLMM (with a logit link function) is:

where Presence is the expected mean probability of sighting occurrence. and

represent the fixed and random covariates used in the models: the

environmental predictors and the spatial grid in the spatial models, and habitat

suitability, zero-centred effort and time, and year in the temporal models. The

index i corresponds to the number of observations among grid-cells in the

spatial models, and years in the temporal models. and represent the

parameters of the fixed and random effects, respectively, and � is the intercept.

The log of fishing effort was included as an offset only in the spatial models.”

provide some more detail on the exact nature of the logbook records of whale

shark occurrences, including quantitative information or comment on the per

cent of logbooks in which whale shark occurrences were consistently, or ever,

Page 286: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

288

recorded (i.e. does the entire fleet report whale sharks?). How might this alter

your interpretation of the zeroes (i.e. do zeros only reflect true absences and

failure to detect whale sharks, or do zeros reflect these two cases as well as

'failure to report whale sharks)? How might this alter how you treat effort' in

your models? OR Are these data not provided on a trip by trip, or vessel by

vessel basis, but merely aggregated as per the spatial precision described (e.g.

p. 139). These types of detail need to be made clear for readers.

The logbook records made available were merely aggregated as per the spatial

precision described. No information was available on vessel or trip units, and

therefore, we could not estimate quantitative information on the percent of logbooks in

which whale shark occurrences were consistent. To clarify this point, I have changed

the text to:

“The dataset made available by the Institut de Recherche pour le

Développement (France) and the Secretariat of the Pacific Community

comprises most of the central area of the Atlantic (21 °N to 15 °S and 34 °W to

14 °E) and central western Pacific (15 °N to 15 °S and 130 °E to 150 °W)

(Figure 18). It includes date (month and year), longitude and latitude information

for whale shark sightings (0.01° precision), and spatially aggregated information

on sampling effort (number of days spent fishing) per month and per grid cell of

1-° resolution in the Atlantic, and 5-° resolution in the Pacific (Figure 18). No

information on individual vessel or trip units was available.”

Regarding the interpretation of zeros in the data, they can reflect both situations

highlighted by the examiner, i.e., they can be either a real absence or simply a failure

to report a real presence (whether due to a failure in detecting a whale shark or not

reporting detection). An underlying assumption of my models is that this potential bias

Page 287: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

289

(failure to report presence) is evenly distributed across the sampling area and study

period. Because there is no way to distinguish these types of information for zeros, I

generated pseudo-absences, and accounted for biases arising from “this choice of

points as absences” by generating several presence / pseudo-absence datasets (see

previous responses). I could have only treated effort differently if detailed information

had been available on the number of fleets, and trips recording whale shark sightings.

Due to commercial-in-confidence agreements, the datasets I used in this

chapter cannot be displayed as supplementary material for the benefit of readers.

However, public-domain data (similarly aggregated, but at coarser spatial grids) for

both the Western Pacific and Atlantic Oceans can be obtained online:

www.wcpfc.int/science-and-scientific-data-functions/public-domain-data and

www.iccat.int/en/accesingdb.htm, respectively.

Following the examiner’s comment, I have now added the following sentence

to the Discussion:

“Differential sampling effort is a major issue for which we accounted at least

partially in both modelling steps to reduce potential bias in spatial and temporal

effort. However, an underlying assumption of our models is that failure to report

a whale shark presence (where due to a failure in detecting a shark or in

reporting detection) is evenly distributed across the sampling area and study

period.”

clearly explicate the relationship between the sightings and effort data (i.e. the

fact that the two are not paired, and hence candidate cannot implement

Page 288: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

290

traditional 'CPUE'or'SPUE' models, in which the two are available at the same

scale);

The reason why I did not implement traditional ‘SPUE’ models was because the spatial

resolution at which data were made available differed for sightings and effort, with

spatial information for sightings available at 0.01 º precision, and effort data

aggregated at 1 º or 5 º. Normalising the sightings-per-unit-effort (SPUE) would

therefore decrease the resolution of my input (sightings) data to 100 or 500 times

lower (1 º or 5 º instead of 0.01º). Results from models applied to such coarse dataset,

where environmental data would also need to be aggregated to those lower

resolutions, would not result in meaningful information. For example, the relationship

between whale shark occurrences and the (narrow) range of sea surface temperatures

would not have been detected if the temperature values had been averaged within

each 5-º grid cells (i.e., ~ 555 × 555 km grid cells at the Equator).

To keep the highest resolution possible in my dataset, I used the resolution of

the sightings data (0.01-º resolution), and included effort as an offset. This allows

accounting for higher sighting probabilities to be expected when more days were spent

fishing, thereby reflecting how sightings and fishing effort are paired. To clarify this

point, I have now added the following sentence to the Methods:

“To account for the sampling bias associated with effort (more whale shark

sightings expected in areas where fishing effort was higher), and to keep the

highest resolution possible in the input data, we have included effort as an offset

in the models.”

Page 289: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

291

Rather than modeling only certain components of the data (i.e. April to June), I

would prefer the candidate to model all of the data, including seasonal

covariate(s) in the model (e.g. modeling season as continuous variable e.g. as in

Baum & Blanchard 201 0 Fisheries Research);

The main objective of Chapter V is to obtain results comparable to those from the

Indian Ocean described in the two previous chapters – i.e, spatial distribution of whale

sharks at the ocean-basin scale and estimation of temporal trends. In the conclusions

of Chapter III (published in Diversity and Distributions), I discussed that because 70 %

of the sightings occurred during the months of April to June, our spatial predictions are

more robust for that season. For this reason, I did the temporal analysis only for these

months in Chapter IV.

As I state in the abstract: “[whale shark] population trends can only be

understood by examining patterns of occurrence synchrony among locations”.

Therefore, to obtain results comparable to the trend observed in the Indian Ocean, I

used data for that same season (i.e., April to June) to calculate trends in the Atlantic

and Pacific Oceans.

I did, however, include models for some other seasons in Chapter VI, where I

modelled global whale shark habitat suitability using data from the season where more

sightings occurred within each ocean. These were austral winter in the Atlantic (with

1153 total sightings), austral autumn in the Indian (with 811 total sightings), and

austral summer in the Pacific (with 314 total sightings). Due to the generally low

number of sightings reported in seasons not considered in my thesis (e.g., only 7

sightings were reported in the Atlantic during the months of January to March),

including all data in the models would not add important information to improve their

Page 290: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

292

predictive performance. To make this point clearer in the manuscript, I have now

added the following:

“To compare our results for the Atlantic and Pacific with previous results for the

Indian Ocean (Sequeira et al., in press), and due to the generally low number of

sightings in other seasons for each ocean, we used data only for the months of

April to June (austral autumn).”

Furthermore, all likelihood-based binomial models are based on the concept that grid

cells must have unique information regarding the presence or absence of the species

to enable discrimination of a covariate’s effects on presence. Including data for more

than one season would mean repeating grid cell information, because the same grid

cell that has been assigned a presence in one season could be selected as either a

presence or absence in another.

Using season as a continuous variable applying sines and cosines with

periodicity of 0.5 and 1 year (Baum & Blanchard, 2010) seems like an interesting

approach for studies considering year-long variation. However, the method used by

Baum and Blanchart (2010) did not include a likelihood-based method, and therefore

did not allow for estimation of the AIC and BIC ranks to compare models: My study is

entirely designed around the observation that whale shark occurrence is seasonal

(Heyman et al., 2001; Taylor, 1996).

explain more clearly the justification for modeling time/year as both a fixed and

a random effect;

The random effect for time (year) accounts for the fact that same-year observations

are not independent (i.e., year effect). Such analysis of the conditional modes (i.e.,

Page 291: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

293

random effects) can be useful when generating hypotheses for future studies or for

qualitative assessment of any pattern (Gelman & Hill, 2007; MacNeil et al., 2009).

However, for making statistical inferences, any trend observed in these conditional

modes should be included in the fixed part of the models (following Zuur et al., 2009).

This ensures that the random structure contains only information that could not have

been modelled with fixed effects. Therefore, to describe correctly a possible linear or

parabolic-like trend in the whale shark probability of occurrence, the variable Time was

included as a fixed effect through the polynomial function poly() assuming a

polynomial degree of 2 (to allow the detection of non-linear temporal trends). By

accounting for this fixed effect, the resulting plot of the conditional modes of the year

random effect (i.e., conditional on the observations and predictor values) only show

random variation remaining over years (e.g., Figure 29).

Figure 29: Conditional modes of the random year intercept term for models including time as fixed effect (quadratic term; left), and without the fixed temporal trend (right).

To clarify the justification for inclusion of time, both as fixed (Time) and random (year)

effects, I added the following sentence:

“...and time, both as a fixed (time) and random effect (year) to ensure that the

random structure contains only information that could not have been modelled

with fixed effects (following Zuur et al., 2009). To account for a parabolic-like

Page 292: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

294

dependence of occurrences with time, we added the quadratic term for time

using the poly function (with degree 2).”

Discussion p. 150 'we hypothesi[z]e that whale shark occurrence might be

asynchronously cyclical across oceans.'

Typo hypothesi[z]e is now corrected.

This statement is speculative and unconvincing. There is a huge difference

between a small number of individuals moving between ocean basins to ensure

'low genetic differentiation of whale sharks sampled in different oceans' versus

entire (sub)-populations moving across ocean basins to produce asynchronous

cycles across oceans.

The statement is now discussed as:

“Given these opposing results, the evidence for an inter-decadal occurrence

pattern in the Indian Ocean (Sequeira et al., in press), the low genetic

differentiation of whale sharks sampled in different oceans (Castro et al., 2007;

Schmidt et al., 2009), and the implicit notion that whale sharks travel across

oceans (Hueter et al., 2008; Rowat & Gore, 2007), we hypothesise that whale

shark occurrence might be asynchronously cyclical across oceans. Such a

pattern would be consistent with the inter-ocean migratory behaviour we have

previously hypothesised (Sequeira et al., 2013).

Despite genetic clues (Castro et al., 2007; Schmidt et al., 2009) being the only

evidence to date corroborating inter-ocean migration, these results could also

have arisen even if only a small number of individuals were moving between

Page 293: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

295

ocean basins (Hartl & Clark, 1989). However, the number of whale sharks

occurring in aggregation locations within the predictable season is highly

variable among years (Rowat et al., 2009a; Wilson et al., 2001), and this

variability holds for the number of whale sharks sighted at the ocean-basin scale

(Figure 20a and b). The opposing trends we detected between the Indian and

the western Pacific Oceans during the last decade might therefore reflect a

migration of sharks between oceans. Similar distributional shifts, promoting

asynchrony in occurrences, have also been suggested for other filter-feeding

sharks (basking sharks Cetorhinus maximus) between the west coast of

Republic of Ireland and the Norwegian Sea (Sims & Reid, 2002), and were

potentially associated with spatial changes in foraging conditions. If whale

sharks are migrating between oceans (Sequeira et al., 2013), a cyclic pattern

such as the one observed in the Atlantic Ocean would explain both the

interannual variation in whale sharks numbers at the ocean-basin scale (Figure

20a and b) and the asynchrony in their occurrences in the Indian and Pacific

Oceans. Moreover, such large-group and multi-year migrations would still be

consistent with the low differentiation found in genetic studies (Castro et al.,

2007; Schmidt et al., 2009). These whale shark migrations could be associated

with changes in environmental conditions (Sleeman et al., 2010b; Wilson et al.,

2001), or with sex/age or reproduction-related behaviour (Ramírez-Macías et

al., 2007; Sequeira et al., 2013). Reproduction-associated multi-year cycles

have been observed in anadromous fish (e.g., sockeye salmon Oncorhynchus

nerka) (Dingle, 1996). Multi-year cycles have also been reported for fish catches

in the western Indian Ocean (Jury et al., 2010), and for other marine

Page 294: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

296

megafauna, such as the decadal cycle in cetacean strandings in southeast

Australia (Evans et al., 2005).”

Page 295: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

297

Examiner 2

This is an interesting complementary paper to the previous one. It was

somehow difficult to pinpoint the differences between the two (except the two

oceans).

This chapter consolidates the spatial and temporal analysis presented in the two

previous chapters (Chapters III and IV; Indian Ocean) applying the same methods to

the Atlantic and Pacific Ocean. The main objective was to get results comparable to

the ones obtained in the previous chapters (mostly in terms of time partial effect), with

the general aim of getting an overview of the trend in whale sharks occurrence in the

three major ocean-basins (as opposed to the ‘traditional’ analyses limited to

aggregation areas).

My main concern relates to the Method part, which is not always understandable

for non-aficionados. For instance, the Fourier transformation is absolutely not

explained. You cannot guess that all readers will know what is it about and how

it works. More explanations are needed here. Explain also why you chose this

framework.

I have now updated the Methods detailing the Fourier Transforms (including the

reasons to choose this approach):

“Fourier transforms allow for the decomposition of signals (i.e., time series) into

the sum of sinusoidal curves with different frequencies – a highly useful

approach for ecological time series (Platt & Denman, 1975). To analyse

possible cyclic variation in whale shark occurrences, we applied the fast Fourier

transform function (FFT; following Moler, 2004) to the sightings time series after

Page 296: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

298

correcting for effort bias (i.e., standardizing sightings per unit effort [SPUE], with

an effort unit = 1 fishing day). We used MatLab version 7.12.0.635 (R2011a)

(The Mathworks Inc., Natick, MA), and interpreted the results of the FFT by

plotting the periodogram for the function: power against the inverse of frequency

(time). A cycle is then defined by identifying the strongest frequency.”

I am a bit puzzled by the use of AICw without further attention (the same applied

to the previous chapter as well). What is nice with this framework is also to

extract a weight of evidence of each variable from the list of models. Indeed,

although the weight of evidence in favour of one given model is of interest (it

allows to highlight the potential problem of stepwise if several models have

similar Ai (close to 0), the general interest for ecologists is a step further. By

analogy to weight of evidence in favour of model, the weight of evidence of each

predictor (wpi) can be simply estimated as the sum of the model AICs weights

(wpi) over all models in which the selected predictor appeared. The predictor

with the highest wpi (the closest to 1) gets the highest weight of evidence to

explain the response variable (the highest relative importance). This approach

based on set of multiple models is far more robust than inferring variable

importance based on a single selected model.

I agree with the examiner, and the weight of evidence for each temporal predictor was

now calculated, and I have included them in the Results as a new table (copied

below). I have also included the following text in the:

Methods

Page 297: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

299

“We also calculated the weight of evidence for each predictor used in the

temporal models by summing the wAIC weights over all models in which each

predictor appeared. This allows estimation of the predictor with the highest

relative importance to explain the response variable.”

Results:

“The weight of evidence we estimated for each temporal predictor is shown in

Table 10. Habitat suitability (Hsuit) and effort received the highest weight (1 and

0.999, respectively) in the Atlantic Oceans, while in the Pacific we found the

highest weight for effort (weight of evidence = 1) followed by the linear term for

time (time: 0.953).”

Discussion:

“We found a lack of support for the linear or quadratic time predictor in the

Atlantic Ocean from April to June (Table 9) where habitat suitability (Hsuit) and

effort received the highest weight of evidence (Table 10). At least for the last

decade in the western Pacific, there was evidence for a linear, positive trend in

the probability of whale shark occurrence (Table 9; Figure 21). In the latter

ocean, even though habitat suitability was also ranked as an important variable

(weight of evidence = 0.761), we found the highest weight of evidence for the

predictors effort and time (Table 10).”

Estimated weight of evidence for each temporal predictor used in the generalised linear mixed-effects models: Hsuit - spatial predictor derived from the spatial distribution models, effort - temporal variation in fishing effort and a linear and quadratic term for time (years).

Ocean weight of evidence Hsuit effort time time2

Atlantic 1 0.999 0.367 0.133 Pacific 0.761 1 0.953 0.328

Page 298: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

300

Similarly, instead of keeping one model at the end, I would strongly recommend

to use a weighted-average. These predictions are simply the weighted-averaged

predictions of all submodels. Sub-models with low weight of evidence will have

basically not power, where submodels with similar AIC will gave even weight.

Where wiPi is the prediction from sub-model l, weighted by the weight of

evidence in favour of this sub-model. Associated with this weighted mean, a

weighted variance or standard deviation can be easily estimated which allows a

useful estimation of the uncertainty associated with the modelling approaches.

Unconditional interval of confidence can then be drawn.

The predictions resulting from the spatial models were always calculated using model-

averaging – in all chapters. To clarify this point, I have added the following sentence to

the Methods:

“We compared the models relative strength of evidence by weighting each

model’s Akaike’s information criterion corrected for small sample sizes (wAICc)

(Burnham & Anderson, 2004), and assessed the goodness-of-fit by calculating

the percentage of deviance explained (%De) for each model. We calculated the

10-fold cross-validation error for the model with highest wAICc support and

assessed the model’s predictive power with the � statistic (Cohen, 1960). To

build the habitat suitability maps (as result of the first spatial modelling step), we

weighted each model’s predictions according to its weight of evidence (wAICc),

and used the result of the weighted mean prediction for all models as input to

the temporal models.”

Note: The references indicated by the examiner were already included in the thesis.

Page 299: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

301

CHAPTER VI

Examiner 1

Note: Here I address the examiner 1’s comments made in the previous section, which

also applied to this chapter. Those comments are repeated here under “[Repetition of

comments]”, and identified with the preceding word: “[Repetition]”.

p.162 'penury' Reword.

Now reads:

“However, the species’ high mobility and the lack of expansive occurrence data

mean that teasing apart spatial and temporal patterns is problematic (Sequeira

et al., in press).”

lntroduction pgs. 162-163 This chapter is tightly linked to the previous modelling

chapters in the thesis, and the candidate should be clear about this in the

lntroduction. i.e. When building upon previous research, it is important to

denote what was done (and published) previously and what is new and distinct

in the current piece of work.

I have now changed the end of the Introduction to include a reference to my previous

work:

“For the first time, we used a global dataset of pelagic whale shark sightings

derived from the tuna purse-seine fishery operating in the Atlantic, Indian and

Pacific Oceans to: (1) predict the spatial distribution of whale shark habitat

suitability across three ocean basins, and (2) extend our projections to a future

Page 300: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

302

scenario of ocean warming based on a no-climate-policy reference (no

stabilisation of greenhouse gas emissions). Following our previous work where

we derived ocean-basin scale predictions for whale shark distribution and

temporal trends (Sequeira et al., 2012; Sequeira et al., 2013), herein our

overarching aim was to derive realistic predictions of the species current

geographical range (circumglobal), and provide a modelling tool for exploring

the species’ future distribution under anticipated climate change.”

[Repetition of comments]

[Repetition] This chapter [II] provides an interesting overview of whale shark

research. I don't, however, feel that here (e.g. pages 71-72) or elsewhere in the

thesis (Chapter VI) you have made a convincing argument as to why we should

be particularly concerned about the potential impacts of climate change on this

wide-ranging, highly mobile marine species.

The Introduction of Chapter VI where the examiner’s comment applies now reads:

“... between 30 ºN and 35 ºS (Compagno, 2001; Last & Stevens, 2009). This

range has been defined based on occasional occurrences (Compagno, 2001),

but most importantly, the range of temperatures where the species is expected

to occur: tropical/ warm temperate (Compagno, 2001; Last & Stevens, 2009).

Being ectotherms, ambient temperatures directly influence their metabolic

processes (Sims, 2003). Whale sharks occur regularly at the ocean surface

(Gunn et al., 1999; Rowat et al., 2007; Thums et al., 2012) in association with

sea surface temperatures (Sequeira et al., 2012) possibly to assist

thermoregulation (Thums et al., 2012). They can also dive deeply, opening

Page 301: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

303

speculation on this species’ potential to modify its diving behaviour with

increasing surface water temperatures associated with climate change (Hoegh-

Guldberg & Bruno, 2010), and use deeper, cooler waters more frequently.

However, oxygen limitation might pose a problem for extended time spent in

deeper waters (Graham et al., 2006; Pörtner, 2001; Prince & Goodyear, 2006)

so for this reason, expected warming will likely affect the spatial distribution of

whale sharks possibly by promoting shifts in their present geographical range.

Moreover, because whale sharks are filter-feeders, these shifts might follow the

simultaneous redistribution of plankton (Beaugrand et al., 2009).”

[Repetition] These two chapters would benefit from implementation of an

additional (and different) modeling approach (e.g. models with a truncated

distribution or boosted regression trees) and/or some sensitivity analyses (e.g.

re: generation of the pseudo-absences in for the Atlantic and Pacific Ocean data

sets as in Chapter III).

Following the comments of both examiners, I will include an additional modelling

approach in this chapter by using the BIOMOD modelling framework. This framework

includes multiple habitat suitability modelling techniques, such as generalised linear,

additive and boosting models (Thuiller et al., 2009 for details), allowing the explicit

treatment of model uncertainties. Results obtained so far with this modelling

framework are similar to the predictions obtained with my previous models (see details

in Appendix D). However, my models were developed in a hypothesis-based

framework, while BIOMOD is instead a model-comparison framework not allowing for

direct input on the contribution of each variable (i.e., variables can be included as 2nd-

or 3rd-order polynomials). Therefore, I chose to keep GLMMs as the main modelling

Page 302: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

304

technique. I have now added the following sentence to the Methods (and in the

Results I point the reader to the results presented in supplementary material):

“To account for variability in habitat suitability resulting from different modelling

approaches, we have also run some additional analysis using the BIOMOD

modelling platform (Thuiller, 2003). However, being a model-comparison

framework (rather than hypothesis-based) BIOMOD does not allow for direct

input on the contribution of each variable. Details are presented in

supplementary information (Appendix D).”

Regarding the sensitivity analysis on pseudo-absence datasets, as answered

previously (comments to Chapter V), to assess the influence of pseudo-absences

location choice in the models results, I always generated multiple pseudo-absence

datasets prior to modelling. I have now clarified this in the Methods:

“To assess the influence of the chosen pseudo-absences, we generated each

set 10 times prior to their inclusion in the spatial GLMM.”

[Repetition] At minimum, I would like the candidate to include some discussion

of the potential biases and limitations arising from the data used and/or the

modelling approaches undertaken in these two chapters. Such admissions and

discussions would strengthen (not weaken) these chapters and their related

manuscripts intended for publication in the peer-reviewed literature. The

candidate does an excellent job of reviewing these in Chapter VII, her General

Discussion, but this chapter will not be published in the primary literature, and

this material is critical to each of Chapters 5 and 6 in order for readers to be able

to evaluate the research.

Page 303: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

305

I have now included a discussion of the potential biases and limitations arising from

the data and approach used. Text added to the discussion section is copied below:

“According to Thuiller (2004), projections on climate change impacts in species

occurrence should be developed considering the species’ entire range. In our

case, this range is circumglobal, which is the reason behind our extrapolation to

the total area where predictor values are within the range used to calibrate the

model. The climate-induced impact on species occurrence is usually

investigated by the coupling climate forecasts with species distribution models

(Araújo et al., 2005; Hannah et al., 2002). However, this coupling adds extra

limitations (as specified in the Introduction), and results from such modelling

frameworks are now considered to provide only a first approximation of possible

changes. In the same way, our predictions of future whale shark habitat

changes are intended only to be informative, providing only a baseline for

temperature-dependent predictions. For a more comprehensive study on the

influence of temperature in the occurrence of this species, not only should more

climate-change scenarios be tested (including multiple global circulation model

outputs), other species distribution models could be generated to assess

variability in habitat suitability results, examine error and determine confidence

intervals (Araújo & New, 2006; Thuiller, 2003). We extended our main results

using the BIOMOD modelling platform (see Appendix D – BIOMOD run). The

ensemble habitat suitability results we obtained were similar to those derived

from GLMM (Figure D2 in Appendix D).

Moreover, to improve our model predictions, more sightings data (ideally with

circumglobal coverage) should be used. We have used the largest, most

extensive dataset derived from tuna fisheries logbooks; however, it only covered

Page 304: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

306

a portion of areas managed by Regional Fisheries Management Organisations

(New Zealand Ministry for Primary Industries website at

fs.fish.govt.nz/Page.aspx?pk=103&tk=319.). Therefore, there is scope to

develop our models based on data from other areas, such as the eastern Pacific

Ocean (managed by the Inter-American Tropical Tuna Commission), which

could assist improving the overall model results. As suggested in the previous

chapter, a more collaborative effort between researchers and fisheries

management organisations would assist improving scientific research and, in

this specific case, potential improving our model fitting.

The opportunistic dataset we used only contained whale shark presences

relative to the sea surface only. The latter also applies to the set of

environmental predictors we used to build the models. While the first drawback

could be addressed by randomly generating pseudo-absences (Barbet-Massin

et al., 2012; Phillips et al., 2009) and repeating this procedure multiple times to

account for any biases derived from pseudo-absence selection, the second is

not simple to address. Environmental (and sightings) data spanning subsurface

dynamics are not currently available for practical model development, especially

where the study area covers most of the global ocean extent. Further, the whale

shark sightings dataset I used included recent sightings (up to 2010) that might

herald future patterns of distribution change. Recently, whale sharks have been

seen in locations where they were not expected to occur (Duffy, 2002;

Rodrigues et al., 2012; Turnbull & Randell, 2006), and so the data I used could

already epitomise shifts in this species’ occurrence. It is therefore possible that

my results show an already biased range of suitable habitat for this species.

Page 305: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

307

However, due to the large area under study and the limited data on presences,

excluding part of the dataset available was not an option.

Despite the current lack of alternatives to fisheries-collected sightings in the

open ocean (Jessup, 2003), combining remotely sensed data with

opportunistically recorded occurrences has, however, proved its worth for

estimating whale shark distributions (Kumari & Raman, 2010; McKinney et al.,

2012; Sequeira et al., 2012). Indeed, most of our understanding of whale shark

ecology and biology hails from opportunistically collected datasets by the eco-

tourism (e.g., Wilson et al., 2001) or fishing industries near shore at aggregation

sites (e.g., McKinney et al., 2012). As such, we should endeavour to analyse all

available datasets to conserve this pan-oceanic species and largest remaining

fish – refusal to examine multiple lines of evidence risks imperilling the species

further through lack of ecological understanding.”

[Repetition] -show the equations for the models being used, so that the reader

can more clearly see what is being done;

I have now added the following text to the caption for Table 11:

“The GLMM (with a logit link function) can be expressed as:

where Presence is the expected mean probability of sightings occurrence.

and represent the fixed and random covariates used in the models: the

environmental predictors and the spatial grid, respectively. and represent

the coefficients associated with the fixed and random effects, respectively, and

Page 306: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

308

� is the intercept. The log of fishing effort was included as an offset. The index i

corresponds to the number of observations among grid-cells.”

[Repetition] -provide some more detail on the exact nature of the logbook

records of whale shark occurrences, including quantitative information or

comment on the per cent of logbooks in which whale shark occurrences were

consistently, or ever, recorded (i.e. does the entire fleet report whale sharks?).

How might this alter your interpretation of the zeroes (i.e. do zeros only reflect

true absences and failure to detect whale sharks, or do zeros reflect these two

cases as well as 'failure to report whale sharks)? How might this alter how you

treat effort' in your models? OR Are these data not provided on a trip by trip, or

vessel by vessel basis, but merely aggregated as per the spatial precision

described (e.g. p. 139). These types of detail need to be made clear for readers.

The previous answer to the same comment in Chapter V applies here, and because

the dataset was already described in more detail in that chapter, I have only included

the following sentence in the data description section:

“No information was made available on vessel or trip units for any ocean.”

[Repetition] -clearly explicate the relationship between the sightings and effort

data (i.e. the fact that the two are not paired, and hence candidate cannot

implement traditional 'CPUE'or'SPUE' models, in which the two are available at

the same scale);

The previous answer to the same comment in Chapter V also applies here. The

following change was made in the text:

Page 307: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

309

“To account for spatial bias in sampling effort without reducing the spatial

resolution of the dataset, we included this variable as an offset, and used the

highest resolution available in each ocean.”

Page 308: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

310

Examiner2

The beginning of the Introduction is more or less a copy-paste from the other

articles submitted or published. I understand it is not always easy to find

inspiration but please try to change the sentence and the associated references

(This whole sentence "Species distribution models have been used widely to

estimate habitat requirements and range sizes (Guisan & Thuiller, 2005; Guisan

&, Zimmermann, 2000; Hirzel et al., 2002; Phillips et al., 2009).

This sentence was now changed to:

“...However, assessing how much the distribution of species will be affected by

climate change implies that the current distribution of the species is adequately

identified.

A species’ distribution and its habitat requirements can be estimated by

modelling information on occurrence together with environmental correlates

(Kearney & Porter, 2009). Despite the diverse use of species distribution

models in terrestrial environments, until recently only a handful of these models

had been developed for the marine environment...”

Despite their diverse use, until recently only a handful of these models had been

developed for the marine environment (Elith &...)" appears several times in the

PhD thesis.

Changed as outlined above. However, each chapter needs to stand alone as an

independent publication, and therefore, general content reflecting the same

information (e.g., the current and previous use of species distribution models) are

Page 309: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

311

necessarily repeated within the thesis. This was highlighted in the General

Introduction:

“All my data chapters were written as ‘stand-alone’ manuscripts, and therefore

some of the material presented in the Introduction and Methods sections are

repeated in each chapter.”

This whole part is not clear: "Despite the known limitations of these models,

such as broad spatial scales of prediction (Hannah et al., 2002; Seo et al., 2009)

and considerable variation among climate forecasts (Araújo et a1.,2005; Hannah

et al., 2002; I(nuTti et al.,2010), their biological relevance can be improved by

down-scaling and coupling to ecological processes (Araújo et a1.,2005;

Fordham et al, 2011; Fordham et al.,2012; Hannah et al,2002)."

I have now changed the text to:

“Fortunately, statistical tools exist to facilitate inference in this regard. For

example, known occurrences can be used to define the current distribution of a

species, and future distributions can be projected by coupling climate forecasts

derived from global circulation models (e.g., Araújo et al., 2005; Cheung et al.,

2010). Despite the known limitations of these models, such as biases derived

from the input data or from the geographical extent available (Barbet-Massin et

al., 2012; Barbet-Massin et al., 2010; Phillips et al., 2009), and uncertainties

derived from the models used (both species distribution models and global

circulation models) (Buisson et al., 2010), confidence in their application can be

improved by using an ensemble result (Araújo & New, 2006). This ensemble

result takes into account variability among models used (Hijmans & Graham,

Page 310: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

312

2006; Kearney et al., 2010; Knutti et al., 2010; Lawler et al., 2006). Another

important limitation, especially from a conservation perspective, is the coarse

resolution of predictions derived from global circulation models, and their

consequently limited biological relevance (Hannah et al., 2002; Seo et al.,

2009). However, this can be improved by down-scaling the forecasts, coupling

them to ecological processes (Araújo et al., 2005; Fordham et al., 2011;

Fordham et al., 2012; Hannah et al., 2002).”

There are plenty of more relevant references concerning uncertainty, lack of

mechanisms etc. Here is a non-exhaustive list:

I have now included most of these references in the Introduction of this chapter (as

detailed above).

It is also not very clear what the authors want to explain. I do not understand

"such as broad spatial scales of prediction (Hannah et a1.,2002; Seo et a1.,

2009)". I do not see what is the limitation here. Try to be clearer and more

precise.

I have now changed this sentence to:

“Another important limitation, especially from a conservation perspective, is the

coarse resolution of predictions derived from global circulation models, and their

consequently limited biological relevance (Hannah et al., 2002; Seo et al.,

2009).”

Page 311: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

313

I have a general problem with the overall methodology here. The authors state

that there are known limitations and especially variations when projecting into

the future, but they use only one approach without controlling whether the

results are strongly biased toward the selected approach. Although I like GLMM,

I do not think they are the most suitable tools for such a risk assessment and

that I suggest a careful examination of the uncertainty if such results have to be

used. I would recommend the authors to have a look at packages such as dismo

or biomod2 in R to allow running a set of distribution models and then having a

careful examination of the error, confidence intervals. The biomod2 package,

which I know most, allows such an approach. In a climate change context, this

is something we should always do given the potential use of outputs. Ecologists

find logical to use a set of climate change models, but still doubt they have to do

the same with ecological models. This is out of sense.

I agree that using a set of ecological models allows for examination of error between

different approaches, and I am now including results obtained from the BIOMOD

package (see Appendix D). However my study was driven by a hypothesis-testing

idea, where the aim was to assess the extent to which the predictor variables were

useful to define global suitable habitat for whale sharks by understanding which

contribution derives from each predictor and how they might relate to the biology of the

species in study. The modelling framework used in BIOMOD is not tailored for testing

hypotheses on the predictors used. It instead uses the best fit to the data in each

model, which might include multi-order polynomial functions for some of the predictors.

Such functions are not easy to interpret in a biological context.

Page 312: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

314

I am also concerned about the use of spatial model in a climate change context.

The spatial autocorrelation is certainly due to either dispersal or a missing

variable. Both of them are not explicitly modelled here and there is no way to

know how they will change into the future. It is impossible to "project" a spatial

component and can only lead to spurious projection. I would strongly

reconsider the spatial component here, and rather try to build a simplistic

diffusion model to account for dispersal.

As stated in my previous answer (Examiners’ comment on pp.105, Chapter III),

because data on the key aspects of whale shark life history (including dispersal) are

still to be measured, no more data are currently available to use as predictor, nor to

develop even a simplistic diffusion model (which would at least require a diffusion

coefficient). While this has been done at a fine scale (Sleeman et al., 2010b), it is

beyond reasonable biological inference to do this on a global scale. As the examiner

states, even if these data were available, they could not be explicitly modelled in a

climate change context. The random effects added to my models represent

unobserved random variation.

However, I disagree with the impossibility to use the spatial component in the

models within a forecast framework. The spatial random effect used in my models is

simply a code for each 1-º grid cell in the area covered by the fisheries. By including

this random effect, I am accounting for the correlation between the environmental

predictors within each 1-º grid cell, and excluding its contribution from the models.

When using the forecast results from the global circulation models – where resolution

was increased from 2.5-º to 0.5-º grid cells using bilinear interpolation – the correlation

between points within each 1-º grid cell will still be present. Including this random

Page 313: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

315

effect in the model set in a forecast framework continues reducing possible bias

derived from within 1-º grid cells correlation.

Fig 26: It is surprising to see some red crosses (around Japan and Americas)

that are supposedly in habitats suitability close to 0. Does it mean that the

models fail to detect those zones? How would you explain the high suitability in

Atlantic whereas very few known aggregations have been spotted?

The models simply did not predict high habitat suitability for those areas, as I state in

the Discussion:

“Although whale sharks occur in the Gulf of California (Cárdenas-Torres et al.,

2007) and the Galapagos during most of the year, our model did not predict

higher suitability there – additional data from the eastern Pacific would probably

help improve model performance.”

As I further discussed in that chapter, the relatively low habitat suitability found

in the Pacific reflects the much higher effort in the area, and relatively few whale

sharks sighted. Because sea surface temperature variation was small within the

calibrated area for that ocean (mostly within the range of sea surface temperatures

where whale sharks occur) (Sequeira et al., 2012), this predictor had a lower effect in

estimating habitat suitability in the Pacific Ocean, which might explain the lower habitat

suitability revealed by the models. This was now added to the Discussion as shown

below:

“This might explain why our models did not show evidence that sea surface

temperature affects whale shark occurrence in the Pacific Ocean, and might

Page 314: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

316

also be part of the reason why the habitat suitability in the entire Pacific Ocean

was generally low.”

Moreover, because areas off the coast of America (and Japan) are outside the

area for which our model was fitted, and even though we restricted the prediction area

to the same environmental envelope used in the model, predictions outside the study

area were geographically extrapolated and should be considered carefully. With this

chapter, I aimed to provide a modelling tool for exploring the future distribution of

whale sharks under anticipated climate change. The final maps of habitat suitability

show that I have developed that tool; however, better input data will be needed to

improve the realism of the predictions. I have now made a reference to this point in the

Discussion (please refer to my answer to the Examiner 1’s comment regarding

inclusion of a discussion on the potential biases and limitations of the approach used).

The discussion misses a critical assessment on the quality and reliability of the

models in respect to uncertainty (what about plotting the confidence interval or

coefficient of variation?), robustness and potential use of such projections.

I have now included an assessment of the potential uses and abuses of the model

projections in the Discussion as mentioned previously. However, because I am

currently revising the manuscript based on this chapter , this work is still in progress

and additional results (e.g., from BIOMOD) will be added to the published version -

including an assessment of uncertainty and robustness of the models used in

comparison to other modelling tools, and plots of confidence intervals in the outputs

generated.

Page 315: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

317

CHAPTER VII

Examiner 1

p. 183 'partialled' -Correct.

Now reads:

“According to my spatially explicit trend results, the number of whale sharks

observed by tuna fisheries,...”

p. 183-184: I urge you to carefully reconsider this recommendation. Small

fluctuations in relative abundance, of less than the span of 1 generation, do not

fit the IUCN Red List Criterion (41 or A2). Such changes could be spurious or

cyclical, but clearly should not be extrapolated to the required period of three

generations. We, in the conservation community, need to be vigilant about

not'crying wolf about every observed or inferred population fluctuation.

Following the examiners’ comment, I have now developed on my recommendation

(which I keep, and detail below). For the first time and based on direct observations, I

have shown evidence that whale shark sightings are declining, at a scale that largely

encompasses the species’ known geographical range. Total sightings in both the

Atlantic and Indian Oceans have declined by about 50% in the last decade, and in

some seasons the reduction reached ~ 70% (e.g., sightings during the second quarter

of the year decreased to about 28% in the Atlantic and 35% in the Indian Oceans in

the last few decades).

Page 316: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

318

The current whale shark IUCN status (VU A2bd + 3d) reflects that this taxon is

vulnerable (i.e., faces a high risk of extinction in the wild) based on the (best) evidence

available showing a:

- (A) reduction in population size derived from

- (2) observed, estimated, inferred or suspected population size reduction of ≥

30% over the last 10 years or three generations (whichever is the longer),

where the reduction or its causes may not have ceased OR may not be

understood OR may not be reversible, based both on

o (b) an index of abundance appropriate to the taxon, and on the

o (d) actual or potential levels of exploitation

together with a

- (3) population size reduction of ≥ 30%, projected or suspected to be met

within the next 10 years or three generations, whichever is the longer (up to a

maximum of 100 years), based on the

o (d) actual or potential levels of exploitation.

I have developed on my recommendation that whale sharks should be considered for

the status of Endangered, specifying that the criteria used to recommend this status

are:

- (A) reduction in population size derived from:

o (2) observed, estimated, inferred or suspected population size

reduction of ≥ 50% over the last 10 years or three generations

(whichever is the longer), where the reduction or its causes may not

have ceased OR may not be understood OR may not be reversible,

based on

Page 317: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

319

� (a) direct observation and

� (d) actual or potential levels of exploitation

and from:

o (4) observed, estimated, inferred, projected or suspected population

size reduction of ≥ 50% over any 10 year or three generation period,

whichever is longer (up to a maximum of 100 years in the future),

where the time period must include both the past and the future, and

where the reduction or its causes may not have ceased OR may not

be understood OR may not be reversible, based on

� (d) actual or potential levels of exploitation

Criterion C might also apply:

o (C) population size estimated to number fewer than 2500 mature

individuals and

� (2) continuing decline, observed, projected, or inferred, in

numbers of mature individuals and

� (a) Population structure in the form of:

� (i) no subpopulation estimated to contain more than 250

mature individuals, or

I suggest therefore a change of status from VU A2bd + 3d to EN A2ad + 4d

or possibly, EN C2ai. Below I detail the changes made to the General Discussion:

“According to my spatially explicit trend results, the number of whale sharks

observed by tuna fisheries, which cover a large extent of the three major

oceans, has been reduced to about 50 % in the last decade, both in the Atlantic

Page 318: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

320

and Indian Oceans (there are no data available prior to 2000 for the Pacific

Ocean). The causes for this reduction at such a broad spatial scale (possibly

reflecting a global population trend) are not yet understood and are potentially

still occurring and possibly irreversible. These characteristics are captured

within the IUCN Red List Criterion A for Endangered species. Moreover,

considering the life span of whale sharks, 10 years might correspond to only 10

% of a whale sharks lifetime (or even up to 25% of a single generation). As I

suggested in Chapter V, despite the possible cyclical occurrence of whale

sharks in each ocean basin, the observed reduction in sightings is occurring

simultaneously in a large part of the species known geographical range. Due to

their long life span, late age at maturity, and the relatively little information

available on their reproduction, this species might actually already be facing a

high risk of extinction in the wild from past and ongoing exploitation and climate

warming. Therefore, I recommend that it should be considered at least for the

status of Endangered.

Other criteria from the IUCN Red List class of Endangered might also apply.

For example, Criterion C relates to the total number of mature individuals. Most

whale sharks spotted in specific aggregations are immature (Graham &

Roberts, 2007; Heyman et al., 2001; Rowat et al., 2011; Wilson et al., 2001),

and available population estimates are around 300-500 individuals for a single

site at Ningaloo, Australia (Meekan et al., 2006). Moreover, whale sharks are

thought to attain maturity at ~ 9 metres in total length (Colman, 1997), and there

is evidence that the average size of sighted sharks is declining (Bradshaw et al.,

2008). Considering that whale sharks are only rarely sighted, that most of the

sightings are immature animals, and viewing aggregations as possible sub-

Page 319: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

321

populations, Criterion C2ai could also apply: “population size estimated at <

2500 mature individuals, in continuing decline (observed, projected or inferred)

and with < 250 mature individuals per subpopulation.

In addition to the suggestion of status change from Vulnerable (currently

based on the IUCN Red List Criterion A2bd+3d) to Endangered (based on

Criterion A2ad + 4d or C2a1), my research also provides an updated global map

reflecting seasonality in whale shark habitat suitability, which can now be added

to the whale shark section of the IUCN Red List.”

p. 184 'strongly suggest a connection' This overstates what your evidence

shows. Please reword accordingly.

Deleted “strongly”. Now reads:

“While focusing on the global patterns of whale shark ecology and connectivity,

my research highlighted the following specific findings:

1. Timing and distribution patterns of whale shark occurrence suggest a

connection between several aggregation sites among the three largest

ocean basins (Chapter II)”

p. 188 'preserve' Please distinguish between 'preserve' and 'conserve'. Do you

really mean preserve?

In the text below, I used the word ‘preserve’ to mean the intention of keeping whale

sharks safe (i.e., protected from danger). Although the word ‘conserve’ would also

Page 320: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

322

apply (meaning that the whale shark stock should be managed wisely, i.e., preventing

depletion), I have replaced ‘preserve’ with ‘protect’, as shown below:

“...and despite whale sharks being classed as Vulnerable in the IUCN Red List

(www.iucn.org), a global effort to protect whale sharks is only indirectly

supported through control of international trade of specimens (Appendix II of the

Convention of Trade in Endangered Species of Wild Fauna and Flora;

www.cites.org/eng/app/appendices.php).”

p. 188. What about the Convention on Migratory Species?

I have now included a reference to the Convention on Migratory Species. Text now

reads:

“My results provide a strong incentive for current whale shark management

policies to be revised. There are currently no global measures in place to limit

the exploitation of live whale sharks (Rowat & Brooks, 2012). Whale sharks

were recently included in Appendix II of the Convention on Migratory Species

(CMS, 2010), which could lead to the development of international conservation

measures (Rowat & Brooks, 2012). However, current management of whale

sharks occurs mostly within confined tourist locations through the imposition of

a maximum number of tourism operator licenses, and by implementing codes of

conduct to minimise disturbance in some countries (Cárdenas-Torres et al.,

2007; Pierce et al., 2010; Quiros, 2007). In the 30 years after the first initiative

for international management of marine resources (UNCLOS, 1982), and

despite whale sharks being classed as Vulnerable in the IUCN Red List

(www.iucn.org), a global effort to protect whale sharks is only indirectly

Page 321: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

323

supported through control of international trade of specimens (Appendix II of the

Convention of Trade in Endangered Species of Wild Fauna and Flora;

www.cites.org/eng/app/appendices.php). Also, national bans have been

imposed on targeted commercial fisheries of whale sharks, but they lack

enforcement (Riley et al., 2009), and the global-scale fisheries interaction with

whale sharks is currently not regulated in any way.”

p. 189 - 195 This is an excellent discussion of potential biases and limitations,

and should be worked into each of the relevant chapters (or at minimum the

relevant manuscripts for submission to peer-reviewed journals), as previously

discussed.

I have now included this in Chapters V and VI, as indicated in the answers to examiner

comments made on each section of the thesis.

Page 322: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

324

Examiner 2

Page 183: Although I agree that the PhD is a very interesting and well-conducted

one, I found the first paragraph quite over-stated. Cheung and colleagues have

developed far more advanced species distribution models based on Metabolic

Theory, physiology and trophic interactions.

The work by Cheung et al. (2011; 2009; 2010) has provided great advances in

projecting possible impacts of climate change on exploited marine fish and

invertebrates by using a model integrating population dynamics with a bioclimatic

envelope. Because the models I developed here were species-specific, and due to the

lack of knowledge on the biology of whale sharks, coupling the habitat suitability and

temporal models to population dynamics was not possible. I have now changed the

first paragraph to:

“My research provided the first ocean-scale and global distribution models of the

world’s largest fish, the whale shark (Rhincodon typus). Here, I applied a novel

and comprehensive modelling framework that took spatial variation and sub-

optimal sampling effort data into account while teasing out temporal patterns. I

showed how appropriate statistical approaches can be applied to oceanic

species for which data are incomplete, fragmented and uneven, such as for the

whale shark, highlighting the value of unique datasets collected

opportunistically. My approach also provides a mathematical platform for marine

climate change predictions that have typically lagged behind terrestrial systems,

and shows the possible impacts of a warming climate on a pan-oceanic shark

species.”

Page 323: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

325

REFERENCES

Abc News (2005) Shark fishing under scrutinity.

http://www.abc.net.au/news/newsitems/200511/s1519193.htm.

Aca EQ, Schmidt JV (2011) Revised size limit for viability in the wild: Neonatal and

young of the year whale sharks identified in the Philippines. Asia Life

Sciences, 20, 361-367.

Agardy T (2000) Effects of fisheries on marine ecosystems: a conservationist’s

perspective. ICES Journal of Marine Science, 57, 761-765.

Alava MNR, Dolumbaló ERZ, Yaptinchay AA, Trono RB (1997) Fishery and trade of

whale sharks and manta rays in the Bohol Sea, Philippines. In: Elasmobranch

Biodiversity, Conservation and Management (edited by Fowler, S. L., Reed, T.

M. and Dipper, F. A.). Species Survival Commission. IUCN - The World

Conservation Union.

Anderson, R. C. & Ahmed, H. (1993). The Shark Fisheries in the Maldives. Mal´e:

Ministryof Fisheries and Agriculture, Republic of the Maldives (Bay of Bengal

Programme).

Anderson C, Waheed A (2001) The economics of shark and ray watching in the

Maldives. Shark News - Newsletter of the IUCN Shark Specialist Group, 13, 1-

4. http://www.iucnssg.org/tl_files/Assets/pdf/SN/SN13%20LOW%20RES.pdf

Araújo MB, Guisan A (2006) Five (or so) challenges for species distribution modelling.

Journal of Biogeography, 33, 1677-1688.

Araújo MB, New M (2006) Ensemble forecasting of species distributions. Trends in

Ecology and Evolution, 22, 42-47.

Page 324: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

326

Araújo MB, Pearson R, G., Thuiller W, Erhard M (2005) Validation of species–climate

impact models under climate change. Global Change Biology, 11, 1504-1513.

Araújo MB, Williams PH (2000) Selecting areas for species persistence using

occurrence data. Biological Conservation, 96, 331-345.

Barbet-Massin M, Jiguet F, Albert CH, Thuiller W (2012) Selecting pseudo-absences

for species distribution models: how, where and how many? Methods in

Ecology and Evolution, 3, 327-338.

Barbet-Massin M, Thuiller W, Jiguet F (2010) How much do we overestimate future

local extinction rates when restricting the range of occurrence data in climate

suitability models? Ecography, 33, 878-886.

Barbosa AM, Real R, Vargas JM (2010) Use of coarse-resolution models of species’

distributions to guide local conservation inferences. Conservation Biology, 24,

1378-1387.

Baretta JW, Ebenhöh W, Ruardij P (1995) The European regional seas ecosystem

model, a complex marine ecosystem model. Netherlands Journal of Sea

Research, 33, 233-246.

Barker MJ, Schluessel V (2005) Managing global shark fisheries: suggestions for

prioritizing management strategies. Aquatic conservation: Marine and

Freshwater ecosystems, 15, 325-347.

Bassen V (2007) Conservation efforts of the East African whale shark trust in Kenya.

In: Irvine, T.R. and Keesing J.K. (Eds.) (2007). The First International Whale

Shark Conference: Promoting International Collaboration in Whale Shark

Conservation, Science and Management. Conference Overview, Abstracts

and Supplementary Proceedings, 20-22.

Page 325: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

327

Bates D, Maechler M, Bolker B (2011) lme4: Linear mixed-effects models using S4

classes. In: R package version 0.999375-42. http://CRAN.R-

project.org/package=lme4.

Baughman JL (1955) The oviparity of the whale shark, Rhineodon typus, with records

of this and other fishes in Texas waters. Copeia, 1955, 54-55.

Baum JK, Blanchard W (2010) Inferring shark population trends from generalized

linear mixed models of pelagic longline catch and effort data. Fisheries

Research, 102, 229-239.

Baum JK, Myers RA, Kehler DG, Worm B, Harley SJ, Doherty PA (2003) Collapse and

conservation of shark populations in the Northwest Atlantic. Science, 2999,

389-392.

Beaugrand G, Kirby RR (2010) Climate, plankton and cod. Global Change Biology, 16,

1268-1280.

Beaugrand G, Luczak C, Edwards M (2009) Rapid biogeographical plankton shifts in

the North Atlantic Ocean. Global Change Biology, 15, 1790-1803.

Beaumont LJ, Pitman AJ, Poulsen M, Hughes L (2007) Where will species go?

Incorporating new advances in climate modelling into projections of species

distributions. Global Change Biology, 13, 1368–1385.

Beckley LE, Cliff G, Smale MJ, Compagno LJV (1997) Recent strandings and

sightings of whale sharks in South Africa. Environmental Biology of Fishes,

50, 343-348.

Bedoya D, Manolakos ES, Novotny V (2011) Characterization of biological responses

under different environmental conditions: a hierarchical modeling approach.

Ecological Modelling, 222, 532–545.

Page 326: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

328

Beger M, Possingham HP (2008) Environmental factors that influence the distribution

of coral reef fishes: modeling occurrence data for broad-scale conservation

and management. Marine Ecology Progress Series, 361, 1-13.

Behera SK, Yamagata T (2003) Influence of the Indian Ocean Dipole on the Southern

Oscillation. Journal of the Meteorological Society of Japan, 81, 169-177.

Behrenfeld MJ, O'malley RT, Siegel DA et al. (2006) Climate-driven trends in

contemporary ocean productivity. Nature, 444, 752-755.

Berge J, Johnsen G, Nilsen F, Gulliksen B, Slagstad D (2005) Ocean temperature

oscillations enable reappearance of blue mussels Mytilus edulis in Svalbard

after a 1000 year absence. Marine Ecology Progress Series, 303, 165-175.

Block BA, Jonsen D, Jorgensen SJ et al. (2011) Tracking apex marine predator

movements in a dynamic ocean. Nature, 475, 86-90.

Borrell A, Aguilar A, Gazo M, Kumarran RP, Cardona L (2011) Stable isotope profiles

in whale shark (Rhincodon typus) suggest segregation and dissimilarities in

the diet depending on sex and size. Environmental Biology of Fishes, 92, 559-

567.

Boustany AM, Davis SF, Pyle P, Anderson SD, Le Boeuf BJ, Block BA (2002)

Expanded niche for white sharks. Nature, 415, 35-36.

Boyce MS, Vernier PR, Nielsen SE, Schmiegelow FKA (2002) Evaluating resource

selection functions. Ecological Modelling, 157, 281-300.

Bradshaw CJA (2007) Swimming in the deep end of the gene pool: global population

structure of an oceanic giant. Molecular Ecology, 16, 5111–5114.

Bradshaw CJA, Fitzpatrick BM, Steinberg CC, Brook BW, Meekan MG (2008) Decline

in whale shark size and abundance at Ningaloo Reef over the past decade:

Page 327: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

329

the world’s largest fish is getting smaller. Biological Conservation, 141, 1894-

1905.

Bradshaw CJA, Higgins J, Michael KJ, Wotherspoon SJ, Hindell MA (2004) At-sea

distribution of female southern elephant seals relative to variation in ocean

surface properties. ICES Journal of Marine Science, 61, 1014-1027.

Bradshaw CJA, Mollet HF, Meekan MG (2007) Inferring population trends for the

world’s largest fish from mark–recapture estimates of survival. Journal of

Animal Ecology, 76, 480–489.

Braicovich PE, Timi JT (2008) Parasites as biological tags for stock discrimination of

the Brazilian flathead Percophis brasiliensis in the south-west Atlantic. Journal

of Fish Biology, 73, 557-571.

Brooks K, Rowat D, Pierce SJ, Jouannet D, Vely M (2010) Seeing spots: photo-

identification as a regional tool for whale shark identification. Western Indian

Ocean Journal of Marine Sciences, 9, 185-194.

Brotons L, Thuiller W, Araújo MB, Hirzel AH (2004) Presence-absence versus

presence-only modelling methods for predicting bird habitat suitability.

Ecography, 27, 437-448.

Brunnschweiler JM (2010) The shark reef marine reserve: a marine tourism project in

Fiji involving local communities. Journal of Sustainable Tourism, 18, 29-42.

Brunnschweiler JM, Baensch H, Pierce SJ, Sims DW (2009) Deep-diving behaviour of

a whale shark Rhincodon typus during long-distance movement in the western

Indian Ocean. Journal of Fish Biology, 74, 706-714.

Brunnschweiler JM, Sims DW (2012) Diel oscillations in whale shark vertical

movements associated with meso- and bathypelagic diving. American

Fisheries Society Symposium, 76, 457-469.

Page 328: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

330

Buisson L, Thuiller W, Casajus N, Sovan L, Grenouillet G (2010) Uncertainty in

ensemble forecasting of species distribution. Global Change Biology, 16,

1145-1157.

Bullock TH (1955) Compensation for temperature in the metabolism and activity of

poikilotherms. Biological Reviews, 30, 311–342.

Burgers G, Stephenson DB (1999) The "Normality" of El Niño. Geophysical Research

Letters, 26, 1027-1030.

Burks CM, Driggers WB, Mullin KD (2006) Abundance and distribution of whale sharks

(Rhincodon typus) in the northern Gulf of Mexico. Fishery Bulletin, 104, 579-

585.

Burnham KP, Anderson DR (2004) Multimodel inference: understanding AIC and BIC

in model selection. Sociological Methods & Research, 33, 261-304.

Bush ABG, Philander SGH (1998) The role of ocean-atmosphere interactions in

tropical cooling during the last glacial maximum. Science, 279, 1341-1344.

Caldeira K, Wickett ME (2003) Anthropogenic carbon and ocean pH. Nature, 425, 365.

Cárdenas-Palomo N, Herrera-Silveira J, Reyes Ó (2010) Spatial and temporal

distribution of physicochemical features in the habitat of whale shark

Rhincodon typus (Orectolobiformes: Rhincodontidae) in the north of Mexican

Caribbean. Revista de biología tropical, 58, 399-412.

Cárdenas-Torres N, Enríquez-Andrade R, Rodríguez-Dowdell N (2007) Community-

based management through ecotourism in Bahia de Los Angeles, Mexico.

Fisheries Research, 84, 114 – 118.

Castro ALF, Stewart BS, Wilson SG et al. (2007) Population genetic structure of

Earth’s largest fish, the whale shark (Rhincodon typus). Molecular Ecology,

16, 5183–5192.

Page 329: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

331

Catlin J, Jones R (2009) Whale shark tourism at Ningaloo Marine Park: A longitudinal

study of wildlife tourism. Tourism Management, 31, 1-9.

Catlin J, Jones T, Norman B, Wood D (2010) Consolidation in a wildlife tourism

industry: the changing impact of whale shark tourist expenditure in the

Ningaloo coast region. International Journal of Tourism Research, 12, 134-

148.

Chahine MT (1992) The hydrological cycle and its influence on climate Nature, 359,

373-380.

Chang W-B, Ley M-Y, Fang L-S (1997) Embryos of the whale shark, Rhincodon typus:

early growth and size distribution. Copeia, 1997, 444-446.

Chassot E, Amandè MJ, Chavance P, Pianet R, Dédo Rg RG (2009) Some

preliminary results on tuna discards and bycatch in the frend purse seine

fishery of the easter Atlantic Ocean. Collective Volume of Scientific Papers

ICCAT, 64, 1054-1067.

Chen C-T, Liu K-M, Joung S-J (1997) Preliminary report on Taiwan's whale shark

fishery. Traffic Bulletin, 17, 53-57.

Chen VY, Phipps MJ (2002) Management and Trade of Whale Sharks in Taiwan,

Taipei, The David and Lucile Packard Foundation, ISBN 957986285-6.

Cheung WWL, Dunne J, Sarmiento JL, Pauly D (2011) Integrating ecophysiology and

plankton dynamics into projected maximum fisheries catch potential under

climate change in the Northeast Atlantic. ICES Journal of Marine Science, 68,

1008-1018.

Cheung WWL, Lam VWY, Sarmiento JL, Kearney K, Watson R, Pauly D (2009)

Projecting global marine biodiversity impacts under climate change scenarios.

Fish and Fisheries, 10, 235-251.

Page 330: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

332

Cheung WWL, Lam VWY, Sarmiento JL, Kearney K, Zeller D, Pauly D (2010) Large-

scale redistribution of maximum fisheries catch potential in the global ocean

under climate change. Global Change Biology, 16, 24-35.

Chin A, Kyne PM, Walker TI, Mcauley RB (2010) An integrated risk assessment for

climate change: analysing the vulnerability of sharks and rays on Australia’s

Great Barrier Reef. Global Change Biology, 16, 1936-1953.

Clark E, Nelson DR (1997) Young whale sharks, Rhincodon typus, feeding on a

copepod bloom near La Paz, Mexico. Environmental Biology of Fishes, 50,

63–73.

Clarke LE, Edmonds JA, Jacoby HD, Pitcher H, Reilly JM, Richels R (2007) Scenarios

of greenhouse gas emissions and atmospheric concentrations. A report by the

climate change science program and the subcommittee on global change

research. Washington, DC.

Cliff G, Anderson-Reade MD, Aitken AP, Charter GE, Peddemors VM (2007) Aerial

census of whale sharks (Rhincodon typus) on the northern KwaZulu-Natal

coast, South Africa. Fisheries Research, 84, 41 – 46.

CMS (2010) Meeting on international cooperation on migratory sharks under the

convention on migratory species. In: Convention on the Conservation of

Migratory Species of Wild Animals. Manila, Philippines,

UNEP/CMS/MS3/REPORT.

Cohen J (1960) A coefficient of agreement for nominal scales. Educational and

Psycological Measurement, 20, 37-46.

Colman JG (1997) A review of the biology and ecology of the whale shark. Journal of

Fish Biology, 51, 1219–1234.

Page 331: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

333

Colomer J (2007) Australian government conservation and management of whale

sharks under the Environment Protection and Biodiversity Conservation Act

1999. In: Irvine, T.R. and Keesing J.K. (Eds.) (2007). The First International

Whale Shark Conference: Promoting International Collaboration in Whale

Shark Conservation, Science and Management. Conference Overview,

Abstracts and Supplementary Proceedings, 26-30.

CSIRO (2005). Whale Sharks: six whale sharks tagged at Ningaloo Reef. Hobart,

TAS: Commonwealth Science and Industrial Research Organization–Marine

and Atmospheric Research. Available at

http://www.cmar.csiro.au/tagging/whale/ningaloo.html

Compagno LJV (2001) Sharks of the World: An Annotated and Illustrated Catalogue of

Shark Species Known to Date, vol. 2: Bullhead, Mackerel and Carpet Sharks

(Heterodontiformes, Lamniformes and Orectolobiformes), Rome, FAO Species

Catalogue for Fishery Purposes.

Council of Agriculture (2007) COA bans fishing for whale sharks. In: Taipei Times.

Taiwan News, 27/05/07, 4. Available at

http://www.taipeitimes.com/News/taiwan/archives/2007/05/27/2003362648.

Davies C (2007) Christmas Island – ecotourism and research. In: Irvine, T.R. and

Keesing J.K. (Eds.) (2007). The First International Whale Shark Conference:

Promoting International Collaboration in Whale Shark Conservation, Science

and Management. Conference Overview, Abstracts and Supplementary

Proceedings, p. 71.

Davis D, Banks S, Birtles A, Valentine P, Cuthill M (1997) Whale sharks in Ningaloo

Marine Park: managing tourism in an Australian marine protected area

Tourism Management, 18, 259-271.

Page 332: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

334

Davison AC, Hinkley DV (1997) Bootstrap methods and their application, Cambridge,

Cambridge University Press.

de La Parra Venegas R, Hueter R, Cano JG et al. (2011) An unprecedented

aggregation of whale sharks, Rhincodon typus, in Mexican coastal waters of

the Caribbean Sea. PLoS ONE, 6 (4), e18994.

doi:18910.11371/journal.pone.0018994.

De’ath G (2007) Boosted trees for ecological modelling and prediction. Ecology, 88,

243-251.

Dennis RLH, Sparks TH, Hardy PB (1999) Bias in butterfly distribution maps: the

effects of sampling effort. Journal of Insect Conservation, 3, 33-42.

Deyoung B, Heath M, Werner F, Chai F, Megrey B, Monfray P (2004) Challenges of

modeling ocean basin ecosystems. Science, 304, 1463-1466.

Diaz LE (2007) The whale shark (Rhincodon typus): proposal of fishing regulation in

Cuban waters. In: Irvine, T.R. and Keesing J.K. (Eds.) (2007). The First

International Whale Shark Conference: Promoting International Collaboration

in Whale Shark Conservation, Science and Management. Conference

Overview, Abstracts and Supplementary Proceedings, p. 72.

Dickey-Collas M, Clarke M, Slotte A (2009) “Linking herring”: do we really understand

plasticity? ICES Journal of Marine Science, 66, 1649-1651.

Diggle P, Ribeiro PJ (2007) Model-based geostatistics New York, Springer.

Dimarco SF, Chapman P, Nowlin Jr. WD et al. (2002) Volume transport and property

distributions of the Mozambique Channel. Deep-Sea Research II, 49, 1481-

1511.

Dingle H (1996) Migration: the biology of life on the move, New York, Oxford university

Press.

Page 333: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

335

Dormann CF, Mcpherson JM, Araújo MB et al. (2007) Methods to account for spatial

autocorrelation in the analysis of species distributional data: a review.

Ecography, 30, 609628.

Dove ADM, Arnold J, Clauss TM (2010) Blood cells and serum chemistry in the world’s

largest fish: the whale shark Rhincodon typus. Aquatic Biology, 9, 177-183.

Duffy CaJ (2002) Distribution, seasonality, lengths, and feeding behaviour of whale

sharks (Rhincodon typus) observed in New Zealand waters. New Zealand

Journal of Marine and Freshwater Research, 36, 565–570.

Dulvy NK, Rogers SI, Jennings S, Stelzenmüller V, Dye SR, Skjoldal HR (2008)

Climate change and deepening of the North Sea fish assemblage: a biotic

indicator of warming seas. Journal of Applied Ecology, 45, 1029–1039.

Eckert SA, Dolar LL, Kooyman GL, Perrin W, A. RR (2002) Movements of whale

sharks (Rhincodon typus) in South-east Asian waters as determined by

satellite telemetry. The Zoological Society of London, 257, 111-115.

Eckert SA, Stewart BS (2001) Telemetry and satellite tracking of whale sharks,

Rhincodon typus, in the Sea of Cortez (Mexico) and the north Pacific Ocean

Environmental Biology of Fishes, 60, 299–308.

Edwards Jr. TC, Cutler DR, Zimmermann NE, Geiser L, Alegria J (2005) Model-based

stratifications for enhancing the detection of rare ecological events. Ecology,

86, 1081-1090.

Edwards M (2004) Preface Marine Ecology Progress Series, Supplement: 1-2, 1.

Edwards M, Richardson AJ (2004) Impact of climate change on marine pelagic

phenology and trophic mismatch. Nature, 430, 881-884.

Elith J, Graham CH, Anderson RP et al. (2006) Novel methods improve prediction of

species’ distributions from occurrence data. Ecography, 29, 129-151.

Page 334: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

336

Elith J, Leathwick JR (2009) Species distribution models: ecological explanation and

prediction across space and time. Annual Review of Ecology, Evolution and

Systematics, 40, 677-697.

Elith J, Leathwick JR, T. H (2008) A working guide to boosted regression trees.

Journal of Animal Ecology, 77, 802 - 813.

Elith J, Phillips SJ, Hastie T, Dudík M, Chee YE, Yates CJ (2011) A statistical

explanation of MaxEnt for ecologists. Diversity and Distributions, 17, 43–57.

Engler R, Guisan A, Rechsteiner L (2004) An improved approach for predicting the

distribution of rare and endangered species from occurrence and pseudo-

absence data. Journal of Applied Ecology, 41, 263-274.

Essington TE, Schindler DE, Olson RJ, Kitchell JF, Boggs C, Hilborn R (2002)

Alternative fisheries and the predation rate of yellowfin tuna in the eastern

Pacific Ocean. Ecological Applications, 12, 724-734.

Evans K, Thresher R, Warneke RM, Bradshaw CJA, Pook M, Thiele D, Hindell MA

(2005) Periodic variability in cetacean strandings: links to large-scale climate

events. Biology Letters, 1, 147-150.

Ferreira JG, Hawkins AJS, Monteiro P et al. (2008) Integrated assessment of

ecosystem-scale carrying capacity in shellfish growing areas. Aquaculture,

275, 138-151.

Field IC, Meekan MG, Buckworth RC, Bradshaw CJA (2009) Susceptibility of Sharks,

Rays and Chimaeras to Global Extinction (Chapter 4). In: Advances in Marine

Biology.

Fitzpatrick B, Meekan M, Richards A (2006) Shark attacks on a whale shark

(Rhincodon typus) at Ningaloo reef, Western Australia. Bulletin of Marine

Science, 78, 397-402.

Page 335: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

337

Folke C, Carpenter S, Walker B, Scheffer M, Elmqvist T, Gunderson L, Holling CS

(2004) Regime shifts, resilience, and biodiversity in ecosystem management.

Annual Review of Ecology, Evolution and Systematics, 35, 557-581.

Fordham DA, Wigley TML, Brook BW (2011) Multi-model climate projections for

biodiversity risk assessments. Ecological Applications, doi:10.1890/11-0314.1.

Fordham DA, Wigley TML, Watts MJ, Brook BW (2012) Strengthening forecasts of

climate change impacts with multi-model ensemble averaged projections

using MAGICC/SCENGEN 5.3. Ecography, 35, 4-8.

Fordham SV (2007) International conservation of whale sharks. In: Irvine, T.R. and

Keesing J.K. (Eds.) (2007). The First International Whale Shark Conference:

Promoting International Collaboration in Whale Shark Conservation, Science

and Management. Conference Overview, Abstracts and Supplementary

Proceedings, p. 73.

Franklin J (2010) Mapping species distributions - spatial inference and prediction, New

York, United States of America, Cambridge University Press.

Freeman E (2012) Presence-Absence Model Evaluation (package

‘PresenceAbsence’). Vienna, Austria, CRAN repository; R Foundation for

Statistical Computing. ISBN 3-900051-07-0, URL http://www.R-project.org/.

Freeman EA, Moisen GG (2008) A comparison of the performance of threshold criteria

for binary classification in terms of predicted prevalence and kappa. Ecological

Modelling, 217, 48-58.

Frid A, Baker GG, Dill LM (2008) Do shark declines create fear-released systems?

Oikos, 117, 191-201.

Gadig O (2007) Occurrence, distribution and conservation of the whale shark in the

western-south Atlantic. In: Irvine, T.R. and Keesing J.K. (Eds.) (2007). The

Page 336: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

338

First International Whale Shark Conference: Promoting International

Collaboration in Whale Shark Conservation, Science and Management.

Conference Overview, Abstracts and Supplementary Proceedings - Poster

Presentation, p. 73.

García VB, Lucifora LO, Myers RA (2007) The importance of habitat and life history to

extinction risk in sharks, skates, rays and chimaeras. Proceedings of the

Royal Society B, 275, 83-89.

Garza-Pérez JR, Lehmann A, Arias-González JE (2004) Spatial prediction of coral reef

habitats: integrating ecology with spatial modeling and remote sensing. Marine

Ecology Progress Series, 269, 141-152.

Gebco (2003) General Bathymetric Chart of the Oceans - One Minute Grid, version

2.0. In: http://www.gebco.net.

Gelman A, Hill J (2007) Data analysis using regression and multilevel/hierarchical

models, New York, Cambridge University Press.

Gifford A, Compagno LJV, Levine M (2007a) Aerial surveys of whale sharks

(Rhincodon typus) off the East Coast of Southern Africa. In: Irvine, T.R. and

Keesing J.K. (Eds.) (2007). The First International Whale Shark Conference:

Promoting International Collaboration in Whale Shark Conservation, Science

and Management. Conference Overview, Abstracts and Supplementary

Proceedings, 53-57.

Gifford A, Compagno LJV, Levine M, Antoniou A (2007b) Satellite tracking of whale

sharks using tethered tags. Fisheries Research, 84, 17-24.

Gotelli NJ, Dorazio RM, Ellison AM, Grossman GD (2010) Detecting temporal trends in

species assemblages with bootstrapping procedures and hierarchical models.

Philosophical transactions of the royal society, 365, 3621-3631.

Page 337: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

339

Graham R (2007) Estimating a global population of whale sharks: pitfalls and

opportunities. In: Irvine, T.R. and Keesing J.K. (Eds.) (2007). The First

International Whale Shark Conference: Promoting International Collaboration

in Whale Shark Conservation, Science and Management. Conference

Overview, Abstracts and Supplementary Proceedings, p. 74.

Graham, R. T., Jacquin, S., Bakary, G., Witt, M. & Randriamahazo, H. (2008).

Characteristics of whale shark seasonality and abundance in NW Madagascar

and dispersal in the Western Indian Ocean. Second International Whale Shark

Symposium, Holbox, Quinta Roo, Mexico (PowerPoint display only). Available

at http://www.domino.conanp.gob.mx/presentaciones.html/ (accessed 26

October 2012).

Graham RT, Roberts CM (2007) Assessing the size, growth rate and structure of a

seasonal population of whale sharks (Rhincodon typus Smith 1828) using

conventional tagging and photo identification. Fisheries Research, 84, 71-80.

Graham RT, Roberts CM, Smart JCR (2006) Diving behaviour of whale sharks in

relation to a predictable food pulse. Journal of the Royal Society, 3, 109-116.

Green JR (2011) Tracks of all 14 tagged animals. In: Galapagos Whale Shark Project -

researching for knowledge. Available at

http://www.galapagoswhaleshark.com/index.php?option=com_content&view=

article&id=2:track-of-all-tagged-animals&catid=1:lastest&Itemid=22.

Grémillet D, Lewis S, Drapeau L et al. (2008) Spatial match–mismatch in the Benguela

upwelling zone: should we expect chlorophyll and sea-surface temperature to

predict marine predator distributions? Journal of Applied Ecology, 45, 610-

621.

Page 338: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

340

Gudger EW (1932) The whale shark, Rhineodon typus, among the Seychelles islands.

Nature, 130, 169.

Gudger EW (1934) The geographical distribution of the whale shark (Rhineodon

typus). American Museum of Natural History, 863-893.

Guisan A, Broennimann O, Engler R, Vust M, Yoccoz NG, Lehmann A, Zimmermann

NE (2006) Using niche-based models to improve the sampling of rare species.

Biological Conservation, 20, 501-511.

Guisan A, Thuiller W (2005) Predicting species distribution: offering more than simple

habitat models. Ecology Letters, 8, 993-1009.

Guisan A, Zimmermann NE (2000) Predictive habitat distribution models in ecology.

Ecological Modelling, 135, 147-186.

Gunn JS, Stevens JD, Davis TLO, Norman BM (1999) Observations on the short-term

movements and behaviour of whale sharks (Rhincodon typus) at Ningaloo

Reef (Western Australia). Marine Biology, 135, 553-559.

Gunn, J. S., Whitelaw, A. W., Davis, T. L. O., Bailey, K. & Itano, D. G. (1992). Tuna

tagging in the Coral Sea. Australian Fisheries 51, 22–24.

Gutierrez AP, Ponti L, D’oultremont T, Ellis CK (2008) Climate change effects on

poikilotherm tritrophic interactions. Climatic change, 97, (Suppl 1) S167-S192.

Haetrakul T, Munanansup S, Assawawongkasem N, Chansue N (2007) A case report:

stomach foreign object in whale shark (Rhincodon typus) stranded in

Thailand. In 4th International Symposium on SEASTAR2000 and Asian

Biologging Science, pp. 83–85. Kyoto: Kyoto University Research Information

Repository 2009-03-12T05:56:49Z. Available at http://repository.kulib.kyoto-

u.ac. jp/dspace/bitstream/2433/71015/1/8thSEASTAR_83.pdf (accessed 26

October 2012).

Page 339: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

341

Hall MA, Alverson DL, Metuzals KI (2000) By-catch: problems and solutions. Marine

Pollution Bulletin, 41, 2004-2219.

Hammerschlag N, Gallagher AJ, Lazarre DM (2011) A review of shark satellite tagging

studies. Journal of Experimental Marine Biology and Ecology, 398, 1-8.

Harborne, A. R., Afzal, D. C. & Andrews, M. J. (2001). Honduras: Caribbean coast.

Marine Pollution Bulletin 42, 1221–1235.

Hanfee F (1997) Trade in sharks and shark products in India: A preliminary survey. In:

TRAFFIC, WWF India.

Hanfee F (2001) Gentle giants of the sea - India's whale shark fishery: A report on

trade in whale shark off Gujarat coast. In: TRAFFIC, WWF India.

Hannah L, Midgley GF, Millar D (2002) Climate change-integrated conservation

strategies. Global Ecology and Biogeography, 11, 485-495.

Hanski I (1994) A practical model of metapopulation dynamics. Journal of Animal

Ecology, 63, 151-162.

Hanski I (1998) Metapopulation dynamics. Nature, 396, 41-49.

Hanski I, Gilpin M (1991) Metapopulation dynamics: brief history and conceptual

domain. Biological Journal of the Linnean Society, 42, 3-16.

Hartl DL, Clark AG (1989) Principles of Population Genetics, 2nd edition, Sinauer

Associates Inc.; Sunderland, Massachusetts.

Hastie T, Tibshirani R (1986) Generalized Additive Models. Statistical science, 1, 297-

318.

Hays GC, Bradshaw CJA, James MC, Lovell P, Sims DW (2007) Why do Argos

satellite tags deployed on marine animals stop transmitting? Journal of

Experimental Marine Biology and Ecology, 349, 52–60.

Page 340: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

342

Hays GC, Richardson AJ, Robinson C (2005) Climate change and marine plankton.

Trends in Ecology and Evolution, 20, 337-344.

Hazin FHV, Vaske Júnior T, Oliveira PG, Macena BCL, Carvalho F (2008)

Occurrences of whale shark (Rhincodon typus Smith, 1828) in the Saint Peter

and Saint Paul archipelago, Brazil. Brazilian Journal of Biology, 68, 385-389.

Hegerl, G. C. & Bindoff, N. L. (2005). Warming the world’s oceans. Science 309, 254–

255.

Hellgren EC, Onorato DP, Skiles JR (2005) Dynamics of a black bear population within

a desert metapopulation. Biological Conservation, 122, 131-140.

Hendriks IE, Duarte CM, Heip CHR (2006) Biodiversity research still grounded.

Science, 312, 1715.

Heyman WD, Carr LM, Lobel PS (2010) Diver ecotourism and disturbance to reef fish

spawning aggregations: it is better to be disturbed than to be dead. Marine

Ecology Progress Series, 419, 201-210.

Heyman WD, Graham RT, Kjerfve B, Johannes RE (2001) Whale sharks Rhincodon

typus aggregate to feed on fish spawn in Belize. Marine Ecology Progress

Series, 215, 275-282.

Hiddink JG, Ter Hofstede R (2008) Climate induced increases in species richness of

marine fishes. Global Change Biology, 14, 453-460.

Hijmans RJ, Graham CH (2006) The ability of climate envelope models to predict the

effect of climate change on species distributions. Global Change Biology, 12,

2272-2281.

Hinder SL, Hays GC, Edwards M, Roberts EC, Walne AW, Gravenor MB (2012)

Changes in marine dinoflagellate and diatom abundance under climate

change. Nature Climate Change, 2, 271-275.

Page 341: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

343

Hirzel AH, Hausser J, Chessel D, Perrin N (2002) Ecological-Niche Factor Analysis:

how to compute habitat-suitability maps without absence data? Ecology, 83,

2027–2036.

Hobbs J-PA, Hamanaka T, Mcdonald CA, Gilligan JJ, Neilson J (2009) Seasonal

aggregation of juvenile whale sharks (Rhincodon typus) at Christmas Island,

Indian Ocean. Coral Reefs, 28, 577.

Hoegh-Guldberg O, Bruno JF (2010) The impact of climate change on the world’s

marine ecosystems. Science, 328, 1523.

Hoffmayer ER, Franks JS, Driggers WB, Oswald KJ, Quattro J, M. (2007)

Observations of a feeding aggregation of whale sharks, Rhincodon typus, in

the north central Gulf of Mexico. Gulf and Caribbean Research, 19, 69-73.

Hoffmayer ER, Franks JS, Shelley JP (2006) Whale sharks (Rhincodon typus) in the

Northcentral Gulf of Mexico: A rationale for research. Gulf and Caribbean

Fisheries Institute, 57, 255-262.

Holmberg J, Norman B, Arzoumanian Z (2008) Robust, comparable population metrics

through collaborative photo-monitoring of whale sharks Rhincodon typus.

Ecological Applications, 18, 222-233.

Holmberg J, Norman B, Arzoumanian Z (2009) Estimating population size, structure,

and residency time for whale sharks Rhincodon typus through collaborative

photo-identification. Endangered Species Research, 7, 39-53.

Hsu H-H, Joung S-J, Liao Y-Y, Liu K-M (2007) Satellite tracking of juvenile whale

sharks, Rhincodon typus, in the Northwestern Pacific. Fisheries Research, 84,

25-31.

Page 342: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

344

Hsu HH, Joung SJ, Liu KM (2012) Fisheries, management and conservation of the

whale shark Rhincodon typus in Taiwan. Journal of Fish Biology, 80, 1595-

1607, doi: 1510.1111/j.1095-8649.2012.03234.x.

Hueter R (2007) Biological studies of large feeding aggregations of whale sharks

(Rhincodon typus) in the southern Gulf of Mexico. In: Irvine, T.R. and Keesing

J.K. (Eds.) (2007). The First International Whale Shark Conference: Promoting

International Collaboration in Whale Shark Conservation, Science and

Management. Conference Overview, Abstracts and Supplementary

Proceedings, p. 76.

Hueter R, Tyminski J, De La Parra R (2008) The geographical movements of whale

sharks tagged with pop-up archival satellite tags off Quintana Roo, Mexico. In:

Second International Whale Shark Symposium. Holbox, Gulf of Mexico,

(PowerPoint display only) available at:

http://www.domino.conanp.gob.mx/presentaciones.html.

Hutchings JA, Myers RA, Garcia VB, Lucifora LO, Kuparinen A (2012) Life-history

correlates of extinction risk and recovery potential. Ecological Applications, 22,

1061-1067.

Hutchison GE (1957) Concluding remarks. Cold Spring Harbor Symposium on

Quantitative Biology, 22, 415-457.

Irvine, T. R. & Keesing, J. K. E. (Eds) (2007). Whale sharks: science, conservation and

management. In: Proceedings of the First International Whale Shark

Conference. Fisheries Research 84.

IUCN (2010) The International Union for Conservation of Nature red list of threatened

species. Available at: http://www.iucnredlist.org/apps/redlist/details/19488/0.

Page 343: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

345

Iwasaki Y (1970) On the distribution and environment of the whale shark, Rhincodon

typus, in Skipjack fishing grounds in the Western Pacific Ocean. Journal of the

College of Marine Science and Technology, Tokai University, 4, 37-51.

Jackson JBC, Kirby MX, Berger WH et al. (2001) Historical overfishing and the recent

collapse of coastal ecosystems. Science, 293, 629-638.

Jarman SN, Wilson SG (2004) DNA-based species identification of krill consumed by

whale sharks. Journal of Fish Biology, 65, 586 – 591.

Jessup D (2003) Opportunistic research and sampling combined with fish and wildlife

management actions or crisis response. Institute for Laboratory Animal

Research, 44, 277-285.

Jiménez-Valverde A (2011) Insights into the area under the receiver operating

characteristic curve (AUC) as a discrimination measure in species distribution

modelling. Global Ecology and Biogeography, 21, 498-507.

Jiménez-Valverde A, Lobo JM (2007) Threshold criteria for conversion of probability of

species presence to either–or presence–absence. Acta oecologica, 31, 361-

369.

Jonahson M, Harding S (2007) Occurrence of whale sharks (Rhincodon typus) in

Madagascar. Fisheries Research, 84, 132-135.

Jones GP, Almany GR, Russ GR, Sale PF, Steneck RS, Van Oppen MJH, Willis BL

(2009) Larval retention and connectivity among populations of corals and reef

fishes: history, advances and challenges. Coral Reefs, 28, 307-325.

Joshi D, Talwar V, Gandhi R, Mookerjee A (2007) Campaign for whale shark

conservation: experiences from coastal Gujarat, India. In: Irvine, T.R. and

Keesing J.K. (Eds.) (2007). The First International Whale Shark Conference:

Promoting International Collaboration in Whale Shark Conservation, Science

Page 344: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

346

and Management. Conference Overview, Abstracts and Supplementary

Proceedings - CSIRO Marine and Atmospheric Research, Australia, 14-19.

Joung S-J, Chen C-T, Clark E, Uchida S, Huang WYP (1996) The whale shark

(Rhincodon typus) is a livebearer: 300 embryos found in one ‘megamamma’

supreme. Environmental Biology of Fishes, 46, 219 – 223.

Jury M, Mcclanahan T, Maina J (2010) West Indian Ocean variability and East African

fish catch. MArine Environmental Research, 70, 162-170.

Kaschner K, Watson R, Trites AW, Pauly D (2006) Mapping world-wide distributions of

marine mammal species using a relative environmental suitability (RES)

model. Marine Ecology Progress Series, 316, 285-310.

Kearney M, Porter W (2009) Mechanistic niche modelling: combining physiological and

spatial data to predict species ranges. Ecology Letters, 12, 334-350.

Kearney MR, Wintle BA, Porter WP (2010) Correlative and mechanistic models of

species distribution provide congruent forecasts under climate change.

Conservation Letters, 3, 203-213.

Kelly NM, Fonseca M, Whitfield P (2001) Predictive mapping for management and

conservation of seagrass beds in North Carolina. Aquatic conservation:

Marine and Freshwater ecosystems, 11, 437-451.

Kirby RR, Beaugrand G (2009) Trophic amplification of climate warming. Proceedings

of the Royal Society B, 276, 4095-4103.

Klein B, Siedler G (1989) On the origin of the Azores current. Journal of Geophysical

Research, 94, 6159-6168.

Knutti R, Furrer R, Tebaldi C, Cermak J, Meehl GA (2010) Challenges in combining

projections from multiple climate models. Journal of Climate, 23, 2739-2758.

Page 345: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

347

Kreakie BJ, Fan Y, Keitt TH (2012) Enhanced migratory waterfowl distribution

modeling by inclusion of depth to water table data. PLoS ONE, 7, e30142;

doi:30110.31371/journal.pone.0030142.

Kritzer JP, Sale PF (2006) Marine metapopulations. Burlington, MA, USA, Elsevier

Academic Press

Kukuyev, E. I. (1996). The new finds in recently born individuals of the whaleshark

Rhincodon typus (Rhiniodontidae) in the Atlantic Ocean. Journal of

Ichthyology 36, 203–205.

Kumari B, Raman M (2007) Satellite remote sensing as a tool to determine whale

shark distribution. In: Irvine, T.R. and Keesing J.K. (Eds.) (2007). The First

International Whale Shark Conference: Promoting International Collaboration

in Whale Shark Conservation, Science and Management. Conference

Overview, Abstracts and Supplementary Proceedings, 58-62.

Kumari B, Raman M (2010) Whale shark habitat assessments in the northeastern

Arabian Sea using satellite remote sensing. International Journal of Remote

Sensing, 31, 379-389.

Landis JR, Koch GG (1977) The measurement of observer agreement for categorical

data. Biometrics, 33, 159-174.

Last PR, Stevens JD (1994) Sharks and Rays of Australia, CSIRO, Australia.

Last PR, Stevens JD (2009) Sharks and Rays of Australia, 2nd edition; Collingwood,

Vic., CSIRO Pub.

Last PR, White WT, Gledhill DC, Hobday AJ, Brown R, Edgar GJ, Pecl G (2011) Long-

term shifts in abundance and distribution of a temperate fish fauna: a

response to climate change and fishing practicesgeb. Global Ecology and

Biogeography, 20, 58-72.

Page 346: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

348

Lawler JJ, White D, Neilson RP, Blaustein AR (2006) Predicting climate-induced range

shifts: model differences and model reliability. Global Change Biology, 12,

1568-1584.

Lazure P, Dumas F (2008) An external–internal mode coupling for a 3D

hydrodynamical model for applications at regional scale (MARS). Advances in

Water Resources, 31, 233-250.

Leathwick JR, Rowe D, Richardson J, Elith J, Hastie T (2005) Using multivariate

adaptive regression splines to predict the distributions of New Zealand’s

freshwater diadromous fish. Freshwater Biology, 50, 2034–2052.

Legendre P, Legendre L (1998) Numerical ecology, Amsterdam, Elsevier.

Lehmann A, Overton JM, Austin MP (2002) Regression models for spatial prediction:

their role for biodiversity and conservation. Biodiversity and conservation, 11,

2085-2092.

Lester RJG (1990) Reappraisal of the use of parasites for fish stock identification.

Australian Journal of Marine and Freshwater Research, 41, 855-864.

Li W, Wang Y, Norman B (2012) A preliminary survey of whale shark Rhincodon typus

catch and trade in China: an emerging crisis. Journal of Fish Biology, 80,

1608-1618, doi:1610.1111/j.1095-8649.2012.03250.x.

Lichstein JW, Simons TR, Shriner SA, Franzreb KE (2002) Spatial autocorrelation and

autoregressive models in ecology. Ecological Monographs, 72, 445-463.

Link WA, Barker RJ (2006) Model weights and the foundations of multimodel

inference. Ecology, 87, 2626-2635.

Liu C, Berry PM, Dawson TP, Pearson RG (2005) Selecting thresholds of occurrence

in the prediction of species distributions. Ecography, 28, 385-393.

Page 347: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

349

Lobo JM, Jiménez-Valverde A, Hortal J (2010) The uncertain nature of absences and

their importance in species distribution modelling. Ecography, 33, 103-114.

Lobo JM, Jiménez-Valverde A, Real R (2007) AUC: a misleading measure of the

performance of predictive distribution models. Global Ecology and

Biogeography, 17, 145-151.

Machete M, Afonso P (pers. comm.) Azores Fisheries Observer Programme

(Programa de Observação para as Pescas dos Açores - POPA). Dept. of

Oceanography and Fisheries; IMAR - University of the Azores.

Macneil MA, Graham NaJ, Polunin NVC, Kulbicki M, Galzin R, Harmelon-Vivien M,

Rushton SP (2009) Hierarchical drivers of reef-fish metacommunity structure.

Ecology, 90, 252-264.

Maity R, Kumar DN, Nanjundiah RS (2007) Review of hydroclimatic teleconnection

between hydrologic variables and large-scale atmospheric circulation patterns

with Indian perspective. ISH Journal of Hydraulic Engineering, 13, 77-92.

Manel S, Williams HC, Ormerod SJ (2001) Presence-Absence Models in Ecology: The

Need to Account for Prevalence. Journal of Applied Ecology, 38, 921-931.

MCSS (2009) China splash tag starts to transmit. (Marine Conservation of the

Seychelles) In: Seychelles, whale sharks. Available at: http://seychelles-

whale-sharks.blogspot.com.au/2009/07/china-splash-tag-starts-to-

transmit.html.

Martin RA (2007) A review of behavioural ecology of whale sharks (Rhincodon typus).

Fisheries Research, 84, 10-16.

Matsunaga H, Nakano H, Okamoto H, Suzuki Z (2003) Whale shark migration

observed by pelagic tuna fishery near Japan. In: SCTB 16 Working Paper

from 16th Meeting of the standing committee on tuna and billfish at

Page 348: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

350

Mooloolaba, Australia. BBRG-12; National Research Institute of Far Seas

Fisheries, Fisheries Research Agency, Japan.

Mau R, Wilson E (2007) Industry trends and whale shark ecology based on tourism

operator logbooks at Ningaloo Marine Park. In: Irvine, T.R. and Keesing J.K.

(Eds.) (2007). The First International Whale Shark Conference: Promoting

International Collaboration in Whale Shark Conservation, Science and

Management. Conference Overview, Abstracts and Supplementary

Proceedings, 45-52.

Mckinney JA, Hoffmayer ER, Wu W, Fulford R, Hendon JM (2012) Feeding habitat of

the whale shark Rhincodon typus in the northern Gulf of Mexico determined

using species distribution modelling. Marine Ecology Progress Series, 458,

199-211.

Mcmahon C, Hays GC (2006) Thermal niche, large-scale movements and implications

of climate change for a critically endangered marine vertebrate. Global

Change Biology, 12, 1330–1338.

Meekan MG, Bradshaw CJA, Press M, Mclean C, Richards A, Quasnichka S, Taylor

JG (2006) Population size and structure of whale sharks Rhincodon typus at

Ningaloo Reef, Western Australia. Marine Ecology Progress Series, 319, 275–

285.

Meekan MG, Jarman SN, Mclean C, Schultz MB (2009) DNA evidence of whale

sharks (Rhincodon typus) feeding on red crab (Gecarcoidea natalis) larvae at

Christmas Island, Australia. Marine and Freshwater Research, 60, 607–609.

Mellin C, Bradshaw CJA, Meekan MG, Caley MJ (2010a) Environmental and spatial

predictors of species richness and abundance in coral reef fishes. Global

Ecology and Biogeography, 19, 212–222.

Page 349: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

351

Mellin C, Bradshaw CJA, Meekan MG, Caley MJ (2010b) Environmental and spatial

predictors of species richness and abundance in coral reef fishes. Global

Ecology and Biogeography, 19, 212-222.

Mellin C, Russell BD, Connell SD, Brook BW, Fordham DA (2011) Geographic range

determinants of two commercially important marine molluscs. Diversity and

Distributions, 18, 133-146.

Moler C (2004) Numerical Computing with MATLAB. In:

http://www.mathworks.com/moler. Electronic edition: The MathWorks, Inc.,

Natick, MA, 2004.

Morjan CL, Rieseberg LH (2004) How species evolve collectively: implications of gene

flow and selection for the spread of advantageous alleles. Molecular Ecology,

13, 1341-1356.

MML (2010) Dubai whale shark’s movements mapped: whale shark released from

Atlantis, The Palm is the first tracked in Arabian Gulf . Sarasota, FL: Mote

Marine Laboratory. Available at

http://www.mote.org/index.php?src=news&refno=379&category=News room

Motta PJ, Maslanka M, Hueter RE et al. (2010) Feeding anatomy, filter-feeding rate,

and diet of whale sharks Rhincodon typus during surface ram filter feeding off

the Yucatan Peninsula, Mexico. Zoology, 113, 190-212.

Musick JA, Burgess G, Cailliet G, Camhi M, Fordham S (2000) Management of sharks

and their relatives (Elasmobranchii). Fisheries, 25, 9-13.

Myers RA, Baum JK, Shepherd TD, Powers SP, Peterson CH (2007) Cascading

effects of the loss of apex predatory sharks from a coastal ocean. Science,

315, 1846-1850.

Page 350: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

352

Nelder JA, Wedderburn RWM (1972) Generalized Linear Models. Journal of Royal

Statistics Society, 135.

Nelson JD, Eckert SA (2007) Foraging ecology of whale sharks (Rhincodon typus)

within Bahía de Los Angeles, Baja California Norte, México. Fisheries

Research, 84, 47-64.

Nielsen CLR, Wakamiya SM, Nielsen CK (2008) Viability and patch occupancy of a

swamp rabbit metapopulation at the northern edge of its distribution. Biological

Conservation, 141, 1043-1054.

Njonjo N (2007) Community initiative projects of the East African whale shark trust. In:

Irvine, T.R. and Keesing J.K. (Eds.) (2007). The First International Whale

Shark Conference: Promoting International Collaboration in Whale Shark

Conservation, Science and Management. Conference Overview, Abstracts

and Supplementary Proceedings, 23-25.

Norman B (2005) Rhincodon typus. In: IUCN 2012. IUCN Red List of Threatened

Species. Version 2012.1. Downloaded on 27 September 2012.

www.iucnredlist.org.

Norman B (2007) The ECOCEAN whale shark photoidentification library: A centralized

and scalable approach to whale shark data collection, management, and

analysis. In: Irvine, T.R. and Keesing J.K. (Eds.) (2007). The First International

Whale Shark Conference: Promoting International Collaboration in Whale

Shark Conservation, Science and Management. Conference Overview,

Abstracts and Supplementary Proceedings, 84.

Norman BM, Stevens JD (2007) Size and maturity status of the whale shark

(Rhincodon typus) at Ningaloo Reef in Western Australia. Fisheries Research,

84, 81-86.

Page 351: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

353

O’brien CM, S. CM, Dickman CR, Keating J (2008) Metapopulation dynamics and

threatened species management: why does the broad-toothed rat

(Mastacomys fuscus) persist? Biological Conservation, 141, 1962-1971.

Ogle K (2009) Hierarchical Bayesian statistics: merging experimental and modeling

approaches in ecology. Ecological Applications, 19, 577–581.

Oviedo L, Solís M (2008) Underwater topography determines critical breeding habitat

for humpback whales near Osa Peninsula, Costa Rica: implications for marine

protected areas. Revista de Biologia Tropical (International Journal of Tropical

Biology and Conservation), 56, 591-602.

Parmesan C (2006) Ecological and evolutionary responses to recent climate change.

Annual Review of Ecology, Evolution and Systematics, 37, 637-669.

Pauly D (1997) Growth and mortality of the basking shark Cetorhinus maximus and

their implications for management of whale sharks Rhincodon typus. In

Elasmobranch Biodiversity, Conservation and Management, International

Seminar and Workshop, Sabah, Malaysia (Fowler, S. L., Reed, T. M. &

Dipper, F. A., eds), 199–208. Gland: IUCN SSC Shark Specialist Group.

http://data.iucn.org/dbtw-wpd/edocs/ssc-op-025.pdf

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

marine food webs. Science, 279, 860.

Pauly D, Watson R, Alder J (2005) Global trends in world fisheries: impacts on marine

ecosystems and food security. Philosophical transactions of the royal society,

360, 5-12.

Penven P, Roy C, Brundrit GB et al. (2001) A regional hydrodynamic model of

upwelling in the Southern Benguela. South African Journal of Science, 97,

472-475.

Page 352: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

354

Perry AL, Low PJ, Ellis JR, Reynolds JD (2005) Climate change and distribution shifts

in marine fishes. Science, 308, 1912-1915.

Phillips SJ, Anderson RP, Schapire RE (2006) Maximum entropy modeling of species

geographic distributions. Ecological Modelling, 190, 231–259.

Phillips SJ, Dudík M (2008) Modeling of species distributions with Maxent: new

extensions and a comprehensive evaluation. Ecography, 31, 161-175.

Phillips SJ, Dudík M, Elith J, Graham CH, Lehmann A, Leathwick J, Ferrier S (2009)

Sample selection bias and presence-only distribution models: implications for

background and pseudo-absence data. Ecological Applications, 19, 181–197.

Phillips SJ, Dudík M, Schapire RE (2004) A Maximum Entropy approach to Species

Distribution Modeling. In: Proceedings of the 21st International Conference on

Machine Learning. Banff, Canada.

Pianet R, Molina De AD, Doriso J, Dewals P, Norström V, Ariz J (2009) Statistics of

the main purse seine fleets fishing in the Indian Ocean (1981-2008). Indian

Ocean Tuna Commission, Mombasa, Kenya, IOTC-2009-WPTT-22.

Pierce SJ, Méndez-Jiménez A, Collins K, Rosero-Caicedo R, Monadjem A (2010)

Developing a code of conduct for whale shark interactions in Mozambique.

Aquatic conservation: Marine and Freshwater ecosystems, 20, 782-788.

Pillai SK (1998) On a whale shark Rhincodon typus found accompanied by its young

ones. Marine Fisheries Information Service, Technical and Extension Series,

152, 15.

Pimm SL, Raven P (2000) Extinction by numbers. Nature, 403, 843-845

Pimm SL, Russell GJ, Gittleman JL, Brooks TM (1995) The future of biodiversity.

Science, 269, 347.

Page 353: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

355

Pine R, Alava MNR, Yaptinchay AA (2007) Challenges and lessons learned in setting-

up a community-based whale shark ecotourism program: The case in Donsol,

Philippines. In: Irvine, T.R. and Keesing J.K. (Eds.) (2007). The First

International Whale Shark Conference: Promoting International Collaboration

in Whale Shark Conservation, Science and Management. Conference

Overview, Abstracts and Supplementary Proceedings, 36-44.

Platt T, Denman KL (1975) Spectral analysis in ecology. Annual Review of Ecology,

Evolution and Systematics, 6, 189-210.

Pollard KS, Dudoity S, Van Der Laan MJ (2005) Multiple testing procedures: R

multtest package and applications to genomics. In: Bioinformatics and

Computational Biology Solutions Using R and Bioconductor. (ed Springer).

Pörtner HO (2001) Climate change and temperature-dependent biogeography: oxygen

limitation of thermal tolerance in animals. Naturwissenschaften, 88, 137-146.

Praca E, Gannier A (2007) Ecological niche of three teuthophageous odontocetes in

the northwestern Mediterranean Sea. Ocean Science Discussions, 4, 785-

815.

Pravin P (2000) Whale shark in the Indian coast – need for conservation. Current

science, 79, 310-315.

Prince ED, Goodyear CP (2006) Hypoxia-based habitat compression of tropical

pelagic fishes. Fisheries Oceanography, 15, 451-464.

Qian SS, Cuffney TF, Alameddine I, Mcmahon G, Reckow KH (2010) On the

application of multilevel modeling in environmental and ecological studies.

Ecology, 91, 355–361.

Quiros A (2005) Whale shark “ecotourism” in the Philippines and Belize: evaluating

conservation and community benefits. Tropical Resources Bulletin, 24, 7.

Page 354: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

356

Quiros AL (2007) Tourist compliance to a code of conduct and the resulting effects on

whale shark (Rhincodon typus) behavior in Donsol, Philippines. Fisheries

Research, 84, 102-108.

R Development Core Team (2011) R: A language and environment for statistical

computing. In: R Foundation for Statistical Computing. ISBN 3-900051-07-0,

URL http://www.R-project.org/. Vienna, Austria.

R Development Core Team (2012) R: A language and environment for statistical

computing. In: R Foundation for Statistical Computing. ISBN 3-900051-07-0,

URL http://www.R-project.org/. Vienna, Austria.

Ramírez-Macías D, Vázquez-Juárez R, Galván-Magaña F, Munguía-Vega A (2007)

Variations of the mitochondrial control region sequence in whale sharks

(Rhincodon typus) from the Gulf of California, Mexico. Fisheries Research, 84,

87-95.

Reygondeau G, Beaugrand G (2011) Future climate-driven shifts in distribution of

Calanus finmarchicus. Global Change Biology, 17, 756-766.

Richardson AJ, Poloczanska ES (2008) Under-resourced, under threat. Science, 320,

1294-1295.

Ridgeway G (2013) Generalized boosted regression models - package 'gbm'. Vienna,

Austria, CRAN repository; R Foundation for Statistical Computing. ISBN 3-

900051-07-0, URL http://www.R-project.org/.

Riley MJ, Hale MS, Harman A, Rees RG (2010) Analysis of whale shark Rhincodon

typus aggregations near South Ari atoll, Maldives Archipelago. Aquatic

Biology, 8, 145-150.

Page 355: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

357

Riley MJ, Harman A, Rees RG (2009) Evidence of continued hunting of whale sharks

Rhincodon typus in the Maldives. Environmental Biology of Fishes, 86, 371–

374.

Roberts CM (2002) Deep impact: the rising toll of fishing in the deep sea. Trends in

Ecology and Evolution, 17, 242-245.

Robertson MP, Peter CI, Villet MH, Ripley BS (2003) Comparing models for predicting

species’ potential distributions: a case study using correlative and mechanistic

predictive modelling techniques. Ecological Modelling, 164, 153-167.

Robinson LM, Elith J, Hobday AJ, Pearson RG, Kendall BE, Possingham HP,

Richardson AJ (2011) Pushing the limits in marine species distribution

modelling: lessons from the land present challenges and opportunities. Global

Ecology and Biogeography, 20, 7889-7802.

Rodrigues NV, Correia JPS, Graça JTC, Rodrigues F, Pinho R, Hirofumi M (2012)

First record of a whale shark Rhincodon typus in continental Europe. Journal

of Fish Biology, doi:10.1111/j.1095-8649.2012.03392.x.

Rodríguez-Dowdell N, Enríquez-Andrade R, Cárdenas-Torres N (2007) Property

rights-based management: whale shark ecotourism in Bahia de los Angeles,

Mexico. Fisheries Research, 84, 119-227.

Romanov EV (2002) Bycatch in the tuna purse-seine fisheries of the western Indian

Ocean. Fishery Bulletin, 100, 90-105.

Rose GA (2005) Capelin (Mallotus villosus) distribution and climate: a sea ‘‘canary’’ for

marine ecosystem change. ICES Journal of Marine Science, 62, 1524-1530.

Roughgarden J, Iwasa Y, Baxter C (1985) Demographic theory for an open marine

population with space-limited recruitment. Ecology, 66, 54-67.

Page 356: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

358

Rowat D (2007) Occurrence of whale shark (Rhincodon typus) in the Indian Ocean: a

case for regional conservation. Fisheries Research, 84, 96-101.

Rowat D, Brooks K, March A et al. (2011) Long-term membership of whale sharks

(Rhincodon typus) in coastal aggregations in Seychelles and Djibouti. Marine

and Freshwater Research, 62, 621-627.

Rowat D, Brooks KS (2012) A review of the biology, fisheries and conservation of the

whale shark Rhincodon typus. Journal of Fish Biology, 80, 1019–1056.

Rowat D, Engelhardt U (2007) Seychelles: a case study of community involvement in

the development of whale shark ecotourism and its socio-economic impact.

Fisheries Research, 84, 109-113.

Rowat D, Gore M (2007) Regional scale horizontal and local scale vertical movements

of whale sharks in the Indian Ocean off Seychelles. Fisheries Research, 84,

32–40.

Rowat D, Gore M, Meekan MG, Lawler IR, Bradshaw CJA (2009a) Aerial survey as a

tool to estimate whale shark abundance trends. Journal of Experimental

Marine Biology and Ecology, 368, 1-8.

Rowat D, Gore MA, Baloch BB et al. (2008) New records of neonatal and juvenile

whale sharks (Rhincodon typus) from the Indian Ocean. Environmental

Biology of Fishes, 82, 215-219.

Rowat D, Meekan MG, Engelhardt U, Pardigon B, Vely M (2007) Aggregations of

juvenile whale sharks (Rhincodon typus) in the Gulf of Tadjoura, Djibouti.

Environmental Biology of Fishes, 80, 465-472.

Rowat D, Speed CW, Meekan MG, Gore MA, Bradshaw CJA (2009b) Population

abundance and apparent survival of the vulnerable whale shark Rhincodon

typus in the Seychelles aggregation. Orix, 43, 591-598.

Page 357: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

359

Ruiz P, Laplanche C (2010) A hierarchical model to estimate the abundance and

biomass of salmonids by using removal sampling and biometric data from

multiple locations. Can. J. Fish. Aquat. Sci., 67, 2032–2044.

Runge JA (1998) Should we expect a relationship between primary production and

fisheries? The role of copepod dynamics as a filter of trophic variability.

Hydrobiologia, 167/168, 61-71.

Sa N (2008) Gigantes na ilha de Santa Maria. In: National Geographic. Portugal,

November 2008, Grupo RBA. Available at

http://www.nationalgeographic.pt/sumario.jsp?id=1744887/ (accessed 26

October 2012)

Sabatés A, Martín P, Lloret J, Raya V (2006) Sea warming and fish distribution: the

case of the small pelagic fish, Sardinella aurita, in the western Mediterranean.

Global Change Biology, 12, 2209-2219.

Saji NH, Goswami BN, Vinayachandran PN, Yamagata T (1999) A dipole mode in the

tropical Indian Ocean. Nature, 401.

Sale PF, Kritzer JP (2003) Determining the extent and spatial scale of population

connectivity: decapods and coral reef fishes compared. Fisheries Research,

65, 153-172.

Sambrotto R (1999) Oceanography: colour renewal from space. Nature, 397, 301-302.

Scheffer M, Carpenter S, Young BD (2005) Cascading effects of overfishing marine

systems. Trends in Ecology and Evolution, 20, 579-581.

Schindler DE, Essington TE, Kitchell JF, Boggs C, Hilborn R (2002) Sharks and tunas:

fisheries impacts on predators with constrasting life histories. Ecological

Applications, 12, 735-748.

Page 358: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

360

Schmidt JV, Chen C-C, Sheikh SI, Meekan MG, Norman BM, Joung S-J (2010)

Paternity analysis in a litter of whale shark embryos. Endangered Species

Research, 12, 117-124.

Schmidt JV, Schmidt CL, Ozer F, Ernst RE, Feldheim KA, Ashley MV, Levine M (2009)

Low genetic differentiation across three major ocean populations of the whale

shark, Rhincodon typus. PLoS ONE, 4 (4), e4988.

doi:4910.1371/journal.pone.0004988.

Schmittner A, Yoshimori M, Weaver AJ (2002) Instability of glacial climate in a model

of the ocean-atmosphere-cryosphere system. Science, 295, 1489-1493.

Seo C, Thorne JH, Hannah L, Thuiller W (2009) Scale effects in species distribution

models: implications for conservation planning under climate change. Biology

Letters, 5, 39-43.

Sequeira A, Ferreira JG, Hawkins AJS et al. (2008) Trade-offs between shellfish

aquaculture and benthic biodiversity: A modelling approach for sustainable

management. Aquaculture, 274, 313-328.

Sequeira A, Mellin C, Delean S, Meekan MG, Bradshaw CJA (in press) Spatial and

temporal predictions of decadal trends in Indian Ocean whale sharks. Marine

Ecology Progress Series.

Sequeira A, Mellin C, Rowat D, Meekan MG, Bradshaw CJA (2012) Ocean-scale

prediction of whale shark distribution. Diversity and Distributions, 18, 504-518.

Sequeira AMM, Mellin C, Meekan MG, Sims DW, Bradshaw CJA (2013) Inferred

global connectivity of whale shark Rhincodon typus populations. Journal of

Fish Biology, 82, 367-389.

Page 359: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

361

Sérgio C, Figueira R, Draper D, Menezes R, Sousa AJ (2007) Modelling bryophyte

distribution based on ecological information for extent of occurrence

assessment. Biological Conservation, 135, 341-351.

Sims DW (2003) Tractable models for testing theories about natural strategies:

foraging behaviour and habitat selection of free-ranging sharks. Journal of

Fish Biology, 63, (Supplement A) 53-73.

Sims DW, Reid PC (2002) Congruent trends in long-term zooplankton decline in the

north-east Atlantic and basking shark (Cetorhinus maximus) fishery catches

off west Ireland. Fisheries Oceanography, 11, 59-63.

Sims DW, Southall EJ, Richardson AJ, Reid PC, Metcalfe JD (2003) Seasonal

movements and behaviour of basking sharks from archival tagging: no

evidence of winter hibernation. Marine Ecology Progress Series, 248, 187-

196.

Sims DW, Southall EJ, Tarling GA, Metcalfe JD (2005) Habitat-specific normal and

reverse diel vertical migration in the plankton-feeding basking shark. Journal

of Animal Ecology, 74, 755-731.

Sleeman JC, Meekan MG, Fitzpatrick BJ, Steinberg CR, Ancel R, Bradshaw CJA

(2010a) Oceanographic and atmospheric phenomena influence the

abundance of whale sharks at Ningaloo Reef, Western Australia. Journal of

Experimental Marine Biology and Ecology, 382, 77-81.

Sleeman JC, Meekan MG, Wilson SG et al. (2007) Biophysical correlates of relative

abundances of marine megafauna at Ningaloo Reef, Western Australia.

Marine and Freshwater Research, 58, 608-623.

Sleeman JC, Meekan MG, Wilson SG, Polovina JJ, Stevens JD, Boggs GS, Bradshaw

CJA (2010b) To go or not to go with the flow: environmental influences on

Page 360: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

362

whale shark movement patterns. Journal of Experimental Marine Biology and

Ecology, 390, 84-98.

Smith A (1828) Descriptions of new, or imperfectly known objects of the animal

kingdom, found in South Africa. South African Commercial Advertiser, 145, p.

2.

Southall EJ, Sims DW, Witt MJ, Metcalfe JD (2006) Seasonal space-use estimates of

basking sharks in relation to protection and political-economic zones in the

North-east Atlantic. Biological Conservation, 132, 33-39.

Southward AJ, Hawkins SJ, Burrows MT (1995) Seventy years observations of

changes in distribution and abundance of zooplancton and intertida organisms

in the western English Channel in relation to rising sea temperature. Journal of

Thermal Biology, 20, 127-155.

Speed CW, Meekan MG, Bradshaw CJA (2007) Spot the match – wildlife photo-

identification using information theory. Frontiers in Zoology, 4, 2. doi:

10.1186/1742-9994-1184-1182

Speed CW, Meekan MG, Rowat D, Pierce SJ, Marshall AD, Bradshaw CJA (2008)

Scarring patterns and relative mortality rates of Indian Ocean whale sharks.

Journal of Fish Biology, 72, 1488–1503.

Speed CW, Meekan MG, Russel BC, Bradshaw CJA (2009) Recent whale shark

(Rhincodon typus) beach strandings in Australia. Marine Biodiversity Records,

2, 1-3.

Stacey N, Karam J, Dwyer D, Speed C, Meekan M (2008) Assessing traditional

ecological knowledge of whale sharks (Rhincodon typus) in eastern Indonesia:

a pilot study with fishing communities in Nusa Tenggara Timur. In: Report

prepared for the Department of the Environment, Water, Heritage and the

Page 361: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

363

Arts, Canberra. Darwin, School for Environmental Research, Charles Darwin

University. Available at

http://www.environment.gov.au/coasts/publications/assessing-whale-

sharks.html/ (accessed 26 October 2012).

Stearns C (1976) Life-history tactics: a review of ideas. The Quartely Review of

Biology, 51, 33-47.

Stevens JD (2007) Whale shark (Rhincodon typus) biology and ecology: a review of

the primary literature. Fisheries Research, 84, 4-9.

Stewart BS, Wilson SG (2005) Threatened fishes of the world: Rhincodon typus (Smith

1828) (Rhincodontidae). Environmental Biology of Fishes, 74, 184–185.

Stocker TF, Raible CC (2005) Climate change: water cycle shifts gear. Nature, 434,

830-833.

Suárez JFR, Ramírez JJP, Cano JMG et al. (2007) Whale shark management

strategies, with the participation of local stakeholders, in Yum Balam, Mexico.

In: Irvine, T.R. and Keesing J.K. (Eds.) (2007). The First International Whale

Shark Conference: Promoting International Collaboration in Whale Shark

Conservation, Science and Management. Conference Overview, Abstracts

and Supplementary Proceedings, 31-35.

Sumaila UR, Cheung WWL, Lam VWY, Pauly D, Herrick S (2011) Climate change

impacts on the biophysics and economics of world fisheries. Nature Climate

Change, 1, 449-456.

Taylor JG (1996) Seasonal occurrence, distribution and movements of the whale

shark, Rhincodon typus, at Ningaloo Reef, Western Australia. Marine and

Freshwater Research, 47, 637-642.

Page 362: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

364

Taylor JG (2007) Ram filter-feeding and nocturnal feeding of whale sharks (Rhincodon

typus) at Ningaloo Reef, Western Australia. Fisheries Research, 84, 65-70.

Taylor JG, Pearce AF (1999) Ningaloo Reef currents: implications for coral spawn

dispersal, zooplankton and whale shark abundance. Journal of the Royal

Society of Western Australia, 82, 57-65.

Theberge MM, Dearden P (2006) Detecting a decline in whale shark Rhincodon typus

sightings in the Andaman Sea, Thailand, using ecotourist operator-collected

data. Oryx, 40, 337–342.

Thomas CD, Cameron A, Green RE et al. (2004) Extinction risk from climate change.

Nature, 427.

Thuiller W (2003) BIOMOD – optimizing predictions of species distributions and

projecting potential future shifts under global change. Global Change Biology,

9, 1353-1362.

Thuiller W (2004) Patterns and uncertainties of species’ range shifts under climate

change. Global Change Biology, 10, 2020-2027.

Thuiller W (2007) Climate change and the ecologist. Nature, 448, 550-552.

Thuiller W, Araújo MB, Pearson RG, Whittaker RJ, Brotons L, Lavorel S (2004a)

Uncertainty in predictions of extinction risk. Nature, 427, 145-148.

Thuiller W, Brotons L, Araújo MB, Lavorel S (2004b) Effects of restricting

environmental range of data to project current and future species distributions.

Ecography, 27, 165-172.

Thuiller W, Lafourcade B, Engler R, Araújo MB (2009) BIOMOD: a platform for

ensemble forecasting of species distributions. Ecography, 32, 1-5.

Page 363: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

365

Thums M, Meekan M, Stevens J, Wilson S, Polovina J (2012) Evidence for

behavioural thermoregulation by the world’s largest fish. Journal of the Royal

Society Interface, published online 17 October 2012.

Tilman D, Kareiva P (1997) Spatial ecology - the role of space in population dynamics

and interspecific interactions, Princeton, New Jersey, Princeton University

Press.

Tittensor DP, Baco AR, Brewin PE et al. (2009) Predicting global habitat suitability for

stony corals on seamounts. Journal of Biogeography, 36, 1111–1128.

Topelko KN, Dearden P (2009) The shark watching industry and its potential

contribution to shark conservation. Journal of Ecotourism, 4, 108-128.

Traill LW, Lim MLM, Sodhi NS, Bradshaw CJA (2010) Mechanisms driving change:

altered species interactions and ecosystem function through global warming.

Journal of Animal Ecology, 79, 937–947.

Tsoar A, Allouche O, Steinitz O, Rotem D, Kadmon R (2007) A comparative evaluation

of presenceonly methods for modelling species distribution. Diversity and

Distributions, 13, 397-405.

Turnbull SD, Randell JE (2006) Rare occurrence of a Rhincodon typus (Whale shark)

in the Bay of Fundy, Canada. Northeastern Naturalist, 13, 57-58.

Uchida S (1984) Investigations of whale shark in Taiwan and Philippines. Ibid., 18, 4-

10. Report of Japanese Group for Elasmobranch Studies [Itaera rui kenkyu

renrakukai ho]. Tokyo: Ocean Research Institute, University of Tokyo (in

Japanese with English abstract).

UNCLOS (1982) United Nations Convention on the Law of the Sea. Division for Ocean

Affairs and the Law of the Sea. New York, NY: United Nations. Available at

Page 364: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

366

http://www.un.org/Depts/los/convention_agreements/convention_overview_co

nvention.htm/.

Urbina-Cardona JN, Flores-Villela O (2009) Ecological-Niche modeling and

prioritization of conservation-area networks for mexican herpetofauna.

Conservation Biology, 24, 1031-1041.

Van Tienhoven, A. M., Den Hartog, J. E., Reijns, R. A. & Peddemors, V. M. (2007). A

computer-aided program for pattern-matching of natural marks on the spotted

ragged tooth shark Carcharias taurus. Journal of Applied Ecology 44, 273–

280.

Vanderwal J, Shoo LP, Graham C, Williams SE (2009) Selecting pseudo-absence

data for presence-only distribution modeling: How far should you stray from

what you know? Ecological Modelling, 220, 589-594.

Venables WN, Ripley BD (eds) (2002a) Modern Applied Statistics with S, New york,

Springer.

Venables WN, Ripley BD (2002b) Modern Applied Statistics with S; 4th edition, New

York, Springer.

Vianna GMS, Meekan MG, Pannell DJ, Marsh SP, Meeuwig JJ (2012) Socio-

economic value and community benefits from shark-diving tourism in Palau: a

sustainable use of reef shark populations. Biological Conservation, 145, 267-

277.

Wafar M, Le Corre P, Birrien J-L (1984) Seasonal changes of dissolved organic matter

(C, N, P) in permanently well mixed temperate waters. Limnology and

Oceanograohy, 29, 1127-1132.

Walker G (1925) Correlation in seasonal variations of weather - a further study of

world weather. Monthly Weather Review, 252-254.

Page 365: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

367

Walther G-R, Post E, Convey P et al. (2002) Ecological responses to recent climate

change. Nature, 416, 389-395.

Weir CR (2010) Sightings of whale sharks (Rhincodon typus) off Angola and Nigeria.

Marine Biodiversity Records, 3, 1-5.

Weng KC, Castilho PC, Morrissette JM et al. (2005) Satellite tagging and cardiac

physiology reveal niche expansion in Salmon sharks. Science, 310, 104.

Wernberg T, Russell BD, Thomsen MS, Gurgel CFD, Bradshaw CJA, Poloczanska

ES, Connell SD (2011) Seaweed communities in retreat from ocean warming.

Current Biology, 21, 1828-1832.

White WT, Cavanagh RD (2007) Whale shark landings in Indonesian artisanal shark

and ray fisheries. Fisheries Research, 84, 128-131.

Williams GJ, Aeby GS, Cowie ROM, Davy SK (2010) Predictive modeling of coral

disease distribution within a reef system. PLoS ONE, 5, e9264.

Wilson SG (2002) A whale shark feeding in association with a school of giant herring

at Ningaloo Reef, Western Australia. Journal of the Royal Society of Western

Australia, 85, 43-44.

Wilson SG, Martin RA (2001) Body markings of the whale shark: vestigial or

functional? Western Australian naturalist, 24, 115-134.

Wilson SG, Newbound DR (2001) Two whale shark faecal samples from Ningaloo

Reef, Western Australia. Bulletin of Marine Science, 68, 361-362.

Wilson SG, Polovina JJ, Stewart BS, Meekan MG (2006) Movements of whale sharks

(Rhincodon typus) tagged at Ningaloo Reef, Western Australia. Marine

Biology, 148, 1157-1166.

Wilson SG, Stewart BS, Polovina JJ, Meekan MG, Stevens JD, Galuardi B (2007)

Accuracy and precision of archival tag data: a multiple-tagging study

Page 366: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

368

conducted on a whale shark (Rhincodon typus) in the Indian Ocean. Fisheries

Oceanography, 16, 547-554.

Wilson SG, Taylor JG, Pearce AF (2001) The seasonal aggregation of whale sharks at

Ningaloo Reef, Western Australia: currents, migrations and the El Niño /

Southern Oscillation. Environmental Biology of Fishes, 61, 1-11.

Wintle BA, Bardos DC (2006) Modeling species-habitat relationships with spatially

autocorrelated observation data. Ecological Applications, 16, 1945-1958.

Wintner SP (2000) Preliminary study of vertebral growth rings in the whale shark,

Rhincodon typus, from the east coast of South Africa. Environmental Biology

of Fishes, 59, 441-451.

Wirtz P, Fricke R, Biscoito MJ (2008) The coastal fishes of Madeira Island - new

records and an annotated check-list. Zootaxa, 1715, 1-26.

Wisz MS, Guisan A (2009) Do pseudo-absence selection strategies influence species

distribution models and their predictions? An information-theoretic approach

based on simulated data. BioMed Central Ecology, 9.

Wolfson FH (1983) Records of seven juveniles of the whale shark, Rhiniodon typus.

Journal of Fish Biology, 22, 647-655.

Wolfson FH (1986) Occurrences of whale shark, Rhincodon typus Smith. In: Indo-

Pacific Fish Biology: Proceedings of the Second International Conference on

Indo-Pacific Fishes (Ureno, T., Arai, R., Taniuchi, T. & Matsuura, K., eds), pp.

208–226. Tokyo: Ichthyological Society of Japan

Wolfson FH, Notarbartolo-Di-Sciara G (1981) The whale shark bibliography. Atti della

Societa Italiana di Scienze Naturali e del Museo Civico di Storia Naturale di

Milano, 122, 171-203.

Page 367: Global Distribution Models for Whale Sharks · Global Distribution Models for Whale Sharks (Assessing Occurrence Trends of Highly Migratory Marine Species) Ana Micaela Martins Sequeira

369

Woodby D, Carlile D, Hulbert L (2009) Predictive modeling of coral distribution in the

Central Aleutian Islands, USA. Marine Ecology Progress Series, 397, 227-

240.

Woods JD (1985) The world Ocean Circulation Experiment. Nature, 314, 501-511.

World Widlife Foundation (2012) Dead whale shark found off Manila Bay. In: WWF -

Philippines, Newsroom. Available at:

http://www.wwf.org.ph/newsfacts.php?pg=det&id=169.

Worm B, Barbier EB, Beaumont N et al. (2006) Impacts of biodiversity loss on ocean

ecosystem services. Science, 314, 787-790.

Yopak KE, Frank LR (2009) Brain size and brain organization of the whale shark,

Rhincodon typus, using magnetic resonance imaging. Brain, Behaviour and

Evolution, 74, 121-142.

Zaniewski AE, Lehmann A, Overton JM (2002) Predicting species spatial distributions

using presence-only data: a case study of native New Zealand ferns.

Ecological Modelling, 157, 261-280.

Ziegler J, Dearden P, Rollins R (2011) But are tourists satisfied? Importance-

performance analysis of the whale shark tourism industry on Isla Holbox,

Mexico. Tourism Management, 33, 692-701.

Zuur AF, Ieno EN, Walker N, Saveliev AA, Smith GM (2009) Mixed Effects Models and

Extensions in Ecology with R, first edition. ISBN 978-0-387-87457-9.