interactions between brolgas and sarus cranes in australia · research institute for the...

185
Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy Donnelly NEVARD¹˒³ BSc Hons (Imperial College at Wye, University of London); MSc (Imperial College at Wye, University of London). Submission date 21 st January 2019 For the degree of Doctor of Philosophy Supervisors: Professor Stephen Garnett¹, Dr George Archibald CM², Professor Michael Lawes¹, succeeded by Dr Ian Leiper¹ July 2017. ¹Research Institute for the Environment and Livelihoods, Charles Darwin University, Darwin, NT 0909 ²International Crane Foundation, E11376 Shady Lane Road, Baraboo, WI 53913, United States ³Forever Wild (formerly Wildlife Conservancy of Tropical Queensland), C/- PO Box 505, Ravenshoe, QLD 4888

Upload: others

Post on 10-Aug-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

Research Institute for the Environment and Livelihoods

Interactions between

Brolgas and Sarus Cranes in Australia

Timothy Donnelly NEVARD¹˒³ BSc Hons (Imperial College at Wye, University of London); MSc (Imperial College at Wye,

University of London).

Submission date 21st January 2019

For the degree of Doctor of Philosophy

Supervisors: Professor Stephen Garnett¹, Dr George Archibald CM²,

Professor Michael Lawes¹, succeeded by Dr Ian Leiper¹ July 2017.

¹Research Institute for the Environment and Livelihoods,

Charles Darwin University, Darwin, NT 0909

²International Crane Foundation, E11376 Shady Lane Road, Baraboo, WI 53913, United

States

³Forever Wild (formerly Wildlife Conservancy of Tropical Queensland), C/- PO Box 505,

Ravenshoe, QLD 4888

Page 2: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

2

DEDICATION

This thesis is dedicated to my wife Gwyneth, who cautioned against starting and predicted I’d

never finish…so is as surprised as me.

Page 3: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

3

TABLE OF CONTENTS

Page

DEDICATION 2

TABLE OF CONTENTS 3

ABSTRACT 7

CANDIDATE’S DECLARATION 9

AUTHOR CONTRIBUTIONS TO PUBLISHED PAPERS 10

ACKNOWLEDGEMENTS 13

LIST OF TABLES 15

LIST OF FIGURES 17

LIST OF ABBREVIATIONS 20

CHAPTER 1 – INTRODUCTION, LITERATURE REVIEW AND THESIS

STRUCTURE 21

1.1 Preamble 21

1.2 A Paleoenvironmental Context 22

1.3 Agriculture and Biodiversity 23

1.4 Agricultural Sustainability 26

1.5 Agriculture and Cranes 28

1.6 Australian Cranes 31

1.6.1 The Brolga 32

1.6.2 The Australian Sarus Crane 34

1.6.3 The ‘Sarolga’ 35

1.6.4 The Conservation Status of Australian Cranes 36

1.7 Research Questions and Thesis Structure 36

Page 4: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

4

1.7.1 Research Questions 37

1.7.2 Thesis Structure 38

CHAPTER 2 – STUDY AREA AND METHODS 41

2.1 Study Area and Geographical Scope 41

2.2 Methodology 43

CHAPTER 3 - THE AUSTRALIAN SARUS CRANE 48

3.1 Abstract 48

3.2 Introduction 48

3.3 Methods 53

3.3.1 Samples 53

3.3.2 Extraction 56

3.3.2.1 Nuclear DNA 56

3.3.2.2 Mitochondrial DNA 57

3.3.3 Statistical Analysis 58

3.3.3.1 Nuclear DNA 58

3.3.3.2 Mitochondrial DNA 59

3.4 Results 59

3.5 Discussion 66

3.6 Conclusions and Recommendations 68

CHAPTER 4 – EXTENT AND TRAJECTORY OF INTROGRESSION 70

4.1 Abstract 70

4.2 Introduction 71

4.2.1 Background 71

4.2.2 The Brolga 75

4.2.3 The Australian Sarus Crane 76

Page 5: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

5

4.2.4 Aims 77

4.3 Methods 78

4.3.1 Field Observations and Sample Collection 78

4.3.2 Molecular Analysis 80

4.3.3 Lincoln Petersen Index 84

4.4 Results 84

4.4.1 Field Observations 84

4.4.2 Population Genetics 86

4.4.3 Lincoln Petersen Index 88

4.5 Discussion and Conclusions 89

CHAPTER 5 - RESOURCE PARTITIONING AND COMPETITION 94

5.1 Abstract 94

5.2 Introduction 95

5.3 Methods 97

5.3.1 Atherton Tablelands Study Area 97

5.3.2 Field Surveys 100

5.3.3 Data Analysis 100

5.4 Results 103

5.4.1 Temporal Patterns 103

5.4.2 Spatial Patterns 104

5.4.3 Habitat Use 107

5.4.4 Crop Seasonality 109

5.5 Discussion 110

5.6 Conservation Management 112

5.7 Conclusions and Recommendations 113

Page 6: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

6

CHAPTER 6 - FARMER ATTITUDES 115

6.1 Abstract 115

6.2 Introduction 115

6.3 Study Area 119

6.4 Methods 121

6.5 Results 122

6.5.1 Primary Production on the Atherton Tablelands 122

6.5.2 Farmer Intentions and Attitudes 123

6.6 Discussion and Recommendations 127

6.6.1 Crane Management Options 130

6.6.2 Consequences of Failing to Conserve Cranes on the Atherton Tablelands 133

CHAPTER 7 –DISCUSSION 135

7.1 Introduction 135

7.2 Key Findings 137

7.3 Future Scenarios 142

CHAPTER 8 - CONCLUSIONS AND RECOMMENDATIONS 146

8.1 Conclusions 146

8.2 Recommendations 147

8.3 A Final Word 149

ACKNOWLEDGEMENT OF ASSISTANCE 150

APPENDIX – PERMITS, APPROVALS AND AUTHORITIES 152

BIBLIOGRAPHY 153

Page 7: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

7

ABSTRACT

Brolgas and Sarus Cranes use agricultural lands on the Atherton Tablelands in Australia,

where they come into conflict with farmers. To help frame their conservation management,

this research investigates genetic, spatial, temporal and socioeconomic patterns.

Work on Brolga genetics (incorporating data from this research) which identifies only small

variation between northern and southern populations (exluding New Guinea) has recently

been published, so the genetic component of this thesis has focussed on the Australian Sarus

Crane. Analyses of 10 microsatellite loci and 49 mitochondrial control sequences have

established that it differs significantly from both South and South-east Asian subspecies. A

sample from the extinct Philippine subspecies also clustered with the Australian Sarus Crane,

hinting at its potential importance in relation to reintroduction to Luzon.

Early observations of suspected Brolga and Sarus Crane introgression coincided with

significantly increased opportunities for interaction when foraging on agricultural crops.

Bayesian clustering based on ten microsatellite loci has confirmed that Australian cranes

hybridise and that their hybrids are fertile. Genetic analyses have also provided definitive

evidence that Brolgas and Sarus cranes migrate between the Gulf Plains and Atherton

Tablelands.

Abundances of both species on the Atherton Tablelands were positively correlated, with

Sarus Cranes more abundant on fertile volcanic soils and Brolgas in outlying cropping areas,

closer to roost sites. Both occurred at their highest densities on cultivated land, as well as

harvested maize and peanut stubbles.

Farmers were interviewed to explore their current attitudes and future intentions towards

cranes. Most are beginning to reduce post-harvest crop residues and move to perennial crops;

Page 8: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

8

diminishing traditional crane foraging opportunities and increasing the potential for conflict

resulting from crane damage to new plantings. By creating specific foraging areas, crane

tourism could provide new feeding opportunities for cranes but as they are mobile and able to

take advantage of newly created cropping areas, abandonment of the Atherton Tablelands

Key Biodiversity Area is therefore also possible.

Page 9: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

9

CANDIDATE’S DECLARATION

I hereby warrant that 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.

Whilst every reasonable effort has been made to gain permission and acknowledge the

owners of copyright material, I would be pleased to hear from any copyright owner who has

been omitted or incorrectly acknowledged.

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

made available for loan and photocopying online via the University’s Open Access repository

eSpace.

Signed by: Timothy Donnelly NEVARD Dated: 10th

August 2019

Page 10: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

10

AUTHOR CONTRIBUTIONS TO PAPERS PUBLISHED AND

SUBMITTED FOR PUBLICATION

During the course of my research for this thesis, I led the preparation and submission of the

four papers for peer reviewed publication listed below. They respectively form the basis of

Chapters 3, 4, 5 and 6.

Publications:

1. Nevard, T.D., Haase, M., Archibald, G., M., Leiper, I., Van Zalinge, R., Purchkoon,

N., Siriaroonrat, B., Latt, T.N., Wink, M. and Garnett, S.T. (in press). Subspecies in

the Sarus Crane Antigone antigone revisited; with particular reference to the

Australian population. Submitted to PLOS ONE (26/17/2019).

2. Nevard, T., Haase, M., Archibald, G., Leiper, I. and Garnett, S.T. (2019). The

Sarolga: conservation implications of genetic and visual evidence for hybridization

between the Brolga Antigone rubicunda and the Australian Sarus Crane Antigone

antigone gillae. Oryx. doi:10.1017/S003060531800073X.

3. Nevard, T.D., Franklin D.C., Leiper, I., Archibald, G. and Stephen T. Garnett, S.T

(2019). Agriculture, Brolgas and Australian Sarus Cranes on the Atherton Tablelands,

Australia. Pacific Conservation Biology, doi:10.1071/PC18055.

4. Nevard, T., Leiper, I., Archibald, G and Garnett S (2018). Farming and cranes on the

Atherton Tablelands, Australia. Pacific Conservation Biology,

doi.org/10.1071/PC18055.

Page 11: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

11

Nature and extent of the author’s intellectual contributions:

1. Nevard led the development of the research questions and collected/sourced samples

and data with the assistance of Haase, Archibald, Van Zalinge, Purchkoon,

Siriaroonrat, Latt and Wink. The University of Greifswald laboratory performed the

genetic analyses under the supervision of Haase. Nevard wrote the first draft of the

paper with input from Haase; revised with editorial input from Garnett, Archibald and

Leiper. Van Zalinge, Purchkoon, Siariaroonrat, Latt and Wink all provided minor

amendments. Nevard and Haase developed the tables. Nevard and Leiper developed

the maps.

2. Nevard led the development of the research questions and collected the data. Nevard

worked with Leiper on spatial data analyses. The University of Greifswald laboratory

performed the genetic analyses under the supervision of Haase. Nevard wrote the first

draft of the paper with input from Haase; revised with editorial input from Garnett,

Archibald and Leiper. Nevard and Haase developed the tables. Nevard and Leiper

developed the maps.

3. Nevard led the development of the research questions, collected the data and worked

with Franklin and Leiper on statistical analyses. Nevard wrote the first draft of the

paper with input from Franklin; revised with editorial input from Garnett, Archibald,

and Leiper. Nevard and Franklin developed the tables. Nevard and Leiper developed

the maps.

Page 12: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

12

4. Nevard led the development of the research questions, collected the data and worked

with Garnett and Leiper on data analyses. Nevard wrote the first draft of the paper

with input from Garnett; revised with editorial input from Garnett, Archibald and

Leiper. Nevard developed the tables. Nevard and Leiper developed the maps.

We confirm the candidate’s contribution to the abovementioned papers and consent to the

inclusion of these papers in this thesis.

Martin Haase Robert Van Zalinge

Stephen Garnett Nuchjaree Purchkoon

George Archibald Boripat Siriaroonrat

Ian Leiper Tin Nwe Latt

Don Franklin Michael Wink

Page 13: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

13

ACKNOWLEDGEMENTS

First and foremost, I would like to thank my supervisory team: Professor Stephen Garnett

(principal advisor), Dr George Archibald CM, Dr Ian Leiper and Professor Michael Lawes

(retired 2017). Without their dedicated professional and personal support, I would not have

been able to either envisage or complete this project. In this context, I would particularly like

to acknowledge Professor Garnett, who provided outstanding personal, academic and logistic

direction at every unforeseen turn in the road; and Dr Archibald, who has so generously

shared his unparalleled knowledge of cranes.

I would also like to especially thank Dr Martin Haase of the University of Greifswald, with

whom I have worked on cranes for many years and who, notwithstanding the ‘tyranny of

distance’, has provided unfailing insight and encouragement throughout. I would also like to

acknowledge Dr Don Franklin, who provided great support with statistical analyses.

There are so many other people with whom I have also shared my PhD journey. These

include: Elinor Scambler, Dr Graham Harrington, Ceinwen Edwards and Dr John Grant, who

have all generously shared their insights on cranes of the Atherton Tablelands and Gulf

Plains; Trevor Adil, who helped me break the ice with farmers on the Atherton Tablelands

Dr Barry Hartup, Yulia Momose, Dr Baz Hughes, Dr Geoff Hilton and Dr Inka Veltheim

who have all provided invaluable advice in relation to crane catching and banding techniques;

Yossi Leshem, who shared insights into crane-related tourism; Dr Adam Miller, who was a

great collaborator on Brolga genetics; Dr David Westcott and Adam McKeown of CSIRO,

who provided advice in relation to potential tracking; Silke Fregin, who has marvellously

managed a huge number of samples and complex lab work at the University of Greifswald;

Dr Annabelle Olsson, who facilitated veterinary health clearances; Robert Van Zalinge,

Page 14: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

14

Nuchjaree Purchkoon, Dr Boripat Sinaroonrat, Professor Michael Wink and Tin Nwe Latt

who gathered and contributed genetic samples from Sarus Cranes in S.E. Asia; Emily Imhoff

of Cincinnati Museum and Dr Chris Milensky, the Collections Manager of the Division of

Birds at the Smithsonian Institution, who single-mindedly tracked-down and sampled their

museum specimens; and U Win Naing Thaw (Director of the Myanmar Nature and Wildlife

Conservation Division) who facilitated the collection of Myanmar samples.

A big thank you also to Betsy Didrickson, who unfailingly sourced references and literature

from the International Crane Foundation library; Dr Kerryn Morrison, Dr Gopi Sundar,

Tanya Smith, Dr Claire Mirande, and other colleagues at the International Crane Foundation,

who provided fantastic advice on crane catching, sampling and counting techniques.

In relation to volunteer fieldworkers and landholders, I would like to thank Dr Ruairidh,

Harry and Bronwen Nevard, Stella and Jürgen Freund, Jess Harris, Baron Dominic and

Baroness Vera von Schwertzell, who good naturedly assisted with logistics and sample

collection, notwithstanding pre-dawn starts and cramped conditions in tiny hides; Hans

Rehme of Lemgo, who generously provided access to his outstanding global collection of

cranes; Tom and Tanya Arnold (Miranda Downs cattle station), the Waddell family

(Woodleigh cattle station), Terry Trantor and Nick Reynolds, Marg Atkinson and Greg

Jenkins, the Godfrey family and Fay Moller-Nielsen, who all provided unfettered access to

their farms and cattle properties; and Angela and Peter Freeman of Cairns Tropical Zoo and

the North Queensland Wildlife Trust, who generously provided access to their captive cranes,

as well as funding for laboratory analyses.

Last but not least my heartfelt thanks to the scores of other ‘craniacs’ who helped me to learn

more about these fantastic birds than I ever thought possible.

Page 15: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

15

LIST OF TABLES

Page

Chapter 3

Table 3.1: Sources of Sarus Crane DNA used in analyses 54

Table 3.2: Details of sample collections for Sarus Crane 54

Table 3.3: PCR specifications and diversity of microsatellite loci 60

Table 3.4: Cluster analyses 61

Chapter 4

Table 4.1: Reference birds 80

Table 4.2: PCR specifications and diversity of microsatellite loci 81

Table 4.3: The number (and percentage) of cranes observed across 46 sites on the

Atherton Tablelands 2013-2016 84

Table 4.4. Total number of cranes and proportion of juvenile cranes at sites dominated

by either Brolgas or Sarus Cranes 86

Chapter 5

Table 5.1: Area and value of agricultural production on the Atherton Tablelands,

Queensland 98

Table 5.2: Population and density of crane species by land use type 108

Table 5.3: Probabilities from modelling of crop preferences of two crane species 108

Table 5.4: Density of crane species in crop types on the Atherton Tablelands,

Queensland, during surveys from May to December from 2014-2016 109

Page 16: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

16

Chapter 6

Table 6.1: Agricultural production on the Atherton Tablelands, Queensland, in 2015;

sorted by area for and within each major type of agricultural activity and the intensity

of use by Brolgas and Sarus Cranes 122

Table 6.2: Returns from land on the Atherton Tablelands, Queensland; with different

intensities of dry season crane usage 123

Table 6.3: Landholder ranking of social issues 125

Page 17: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

17

LIST OF FIGURES

Page

Chapter 2

Fig. 2.1: Global distribution of Brolga Antigone rubicunda; Sarus Crane

Antigone antigone; and field data collection sites in northern Queensland 41

Fig. 2.2: Atherton Tablelands region; showing location, major settlements,

land use and Key Biodiversity Area 42

Chapter 3

Fig. 3.1: Global distribution of the Sarus Crane A. antigone; showing populations

and subspecies 50

Fig. 3.2: Extant Sarus Crane subspecies 52

Fig. 3.3: Bayesian cluster analyses of microsatellite data 62

Fig. 3.4: Phylogenetic tree for three subspecies of the Sarus Crane A. antigone;

based on mtDNA sequences resulting from Bayesian analysis 64

Fig. 3.5: TCS network for five populations of the Sarus Crane A. antigone;

based on mtDNA sequences compared to the Brolga A. rubicunda 65

Chapter 4

Fig. 4.1: Typical Brolga with unfledged chick (Gulf Plains); showing small

red skull-cap type comb, dark wattle and grey legs, with scalloped plumage 72

Fig. 4.2: Typical Australian Sarus Cranes with first year juvenile (Atherton

Tablelands); showing pink legs, red comb extending down the neck,

whitish crown and darker grey colour 73

Fig. 4.3: Global distribution of Brolga Antigone rubicunda; Sarus Crane

Antigone antigone; and field data collection sites in northern Queensland 74

Page 18: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

18

Fig. 4.4: Heads and necks of typical species phenotypes and presumed hybrids 79

Fig. 4.5: Pie-charts showing general distribution and clustering of Brolga,

Sarus Crane and Sarolga; visually identified at 46 observation sites on the

Atherton Tablelands 2013-2016 85

Fig. 4.6: STRUCTURE plot with K = 2; showing admixture among a total

of 363 Brolgas and Sarus Cranes 87

Chapter 5

Fig. 5.1: Atherton Tablelands region; showing survey sites, irrigated

agriculture and the Atherton Tablelands Key Biodiversity Area 99

Fig. 5.2: Foraging site hotspot analysis; based on the percent of cranes that

were Brolgas or Sarus 105

Fig. 5.3: Species-specific crane densities at feeding sites; as a function of

distance to the nearest roost 106

Fig. 5.4: Mean density of Brolgas (a) and Sarus Cranes (b) at each survey site.

Roost locations are also presented, classified according to most prominent species

observed at each 106

Fig. 5.5: Regression tree habitat models for Sarus Crane and Brolga 107

Chapter 6

Fig 6.1: Brolgas feeding on volunteer peanuts in an irrigated winter wheat

crop at Innot Hot Springs, Atherton Tablelands, Queensland 117

Fig. 6.2: Pairs of Brolgas and Australian Sarus Cranes with cattle at Kaban,

Atherton Tablelands, Queensland 117

Fig. 6.3: Atherton Tablelands region; showing location, major settlements,

land use and Key Biodiversity Area 120

Page 19: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

19

Chapter 8

Fig. 8.1: Queensland Coat of Arms (with author endorsements) 149

Page 20: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

20

LIST OF ABBREVIATIONS

ABBBS Australian Bird and Bat Banding Scheme

ACBPS Australian Customs and Border Protection Service

AMOVA Analysis of Molecular Variance

CDU Charles Darwin University

CYP Cape York Peninsula

GIS Geographic Information System

ICF International Crane Foundation

IUCN International Union for the Conservation of Nature

KBA Key Biodiversity Area

NMNH National Museum of Natural History

PCR Polymerase Chain Reaction

QP&WS Queensland Parks and Wildlife Service

RIEL Research Institute for the Environment and Livelihoods

USA United States of America

Page 21: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

21

CHAPTER 1 – INTRODUCTION AND LITERATURE

REVIEW

1.1 Preamble

The 15 members of the bird family Gruidae - the cranes - are revered wherever they occur

through Asia, Africa, Europe and Australia (Von Treuenfels, 2006). Australia has two crane

species, the Brolga Antigone rubicunda and Australian Sarus Crane Antigone antigone gillae.

They are similar in appearance and sympatric, with the distribution of the Australian Sarus

Crane entirely overlapped by the much larger range of the Brolga (Menkhorst et al., 2017).

Both species breed in shallow swamps and feed in savanna grassland and on agricultural

lands, as well as occasionally interbreed. This raises questions about how they retain their

specific identities or whether they will continue to do so as the Australian environment

changes.

The Sarus Crane but not the Brolga also occurs in South and South-east Asia, where some

populations have been extirpated and others are declining. Globally the species is listed as

Vulnerable under the IUCN Red List (IUCN 2018). This means that the isolated Australian

population has substantial conservation significance. Not only might interactions with the

Brolga be affecting its prospects in Australia but so too might interactions with intensifying

agriculture. Most crane species have a close relationship with agriculture and the two

Australian cranes also used cultivated land when they encounter it.

In this thesis I throw light on some of these issues, including the relationship of the

Australian Sarus Crane to populations in Asia and its genetic relationship with the Brolga. I

analyse the interactions of both species with agriculture where they co-occur on the Atherton

Tablelands in north Queensland and explore the future of farming in the region through

Page 22: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

22

interviews with farmers. On the basis of this research, I consider the long-term prospects for

cranes and provide recommendations for their conservation.

I explore below the paleoenvironmental context of northern Australia, as well as the global

development of agriculture and its increasing impact on wildlife. I then discuss increasing

research interest in sustainable agriculture and food security, before moving on to review the

impact of agriculture on cranes and providing an introduction to Australia’s cranes.

1.2 A Paleoenvironmental Context

Up until the arrival of European agriculture in the 19th

Century, Australia’s two cranes, the

Brolga Antigone rubicunda and Australian Sarus Crane Antigone antigone gillae, were

confined to savanna woodland and grassland and adapted to the wet-dry conditions of

Australia’s monsoonal north. At the time of lowest sea level during the last Ice Age

(Lambeck et al. 2002), a savanna corridor extended both north and south of the equator

across the Sunda Plain (Bird et al. 2005) a feature soon lost by rising sea levels and rainforest

expansion. This corridor ended not far north of the Pleistocene Lake Carpentaria, around

which there would have been savannas inhabited by megafauna including grazing and

browsing marsupials, fundamentally influencing the extent and quality of crane habitat across

the region.

The maximum extent of global ice caps between 18,000 and 15,000 years ago marked the

greatest extent of land in the Sunda and Sahul Shelves (White, 1997) and the high point of

continental aridity. Deglaciation was underway by 14,000 years ago and global sea levels

rose rapidly from 130 metres below current levels (at the 14,000 year glacial maximum) to 35

metres below current levels 10,000 years ago. During this approximately 4,000 year period,

hydrological stress declined, temperature variations were less limiting to biota and wind

Page 23: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

23

regimes less forceful. The Carpentaria Basin was gradually invaded by rising seas between

12,000 and 8,000 years ago (Joseph et al. 2019), with full marine conditions established

about 8,000 years ago. The period 9,000 to 7,000 years ago was one of optimal rainfall and

temperature for the regional expansion of rainforest and diminishing wet savanna, with sea

levels peaking 5,500 to 5,000 years ago at between 1 and 2 metres above current levels.

Between 3,000 and 2,000 years ago a drier period, with reactivated hydrological stress,

characterised northern Australia (White, 1997). So, throughout the Holocene and the last half

of the Quaternary, Australia has been a land of drought-tolerant biota, which included

Australia’s two mobile and adaptable crane species.

These conditions would have especially determined the distribution of Brolgas, which

uniquely amongst cranes have a salt excreting gland (Hughes and Blackman, 1973); allowing

them to differentially exploit foraging opportunities in harsh brackish coastal environments

more hostile to Sarus Cranes, which do not posess salt glands. When observed in the same

area as Sarus Cranes, Brolgas favour deeper, generally coastal marshes (Walkinshaw, 1973).

Perhaps this is why the Australian population of Sarus Cranes, although long known to

Aboriginal people (Reardon, 2007), was not formally identified until 1964 (Gill, 1967) due to

their restricted distribution and small numbers. Perhaps recently expanding in numbers

because they had access to a suitably- modified savanna environment, biologically

engineered anew by a cattle-based grazing megafauna and the establishment of cultivation

and crops.

1.3 Agriculture and Biodiversity

The development of agriculture in Australia and its impact on biodiversity is recent and needs

to be considered against a global backdrop, with cultivation of domesticated cereal crops first

developing in Eurasia, Africa and North America from about 12,000 years ago (Roberts,

Page 24: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

24

2017). In addition to providing food for humans, these crops also provided accessible and

energy-rich food for many bird species, including cranes. As well as the gradual development

of cropping, pastoralists actively managed the Eurasian steppe and African savannas for their

livestock; extensive natural grazing assemblages using prairie and arid grasslands in North

America and Australasia. The subsequent ‘Columbian exchange’ of crops and technology

between the Americas, Eurasia, Africa and Australasia (Nunn and Qian, 2010) introduced

crops such as maize and potatoes to Eurasian and African cranes and wheat, rice, millet,

sorghum and cattle grazing to cranes in North America. Australasian cranes were only

exposed to all of these cropping changes from the 19th

Century onwards.

The transition from agrarian to industrial economies had its origins in Britain and coincided

with rising human populations and rapid declines in natural habitats (Deane, 1979). By the

1930s the developing world was also gradually intensifying agricultural production, with

increases in cropped area, yields and mechanical harvesting increasing foraging opportunities

for cranes. However, this expansion also led to the loss both of wetland and grassland

habitats, contributing to the decline and sometimes the disappearance of crane populations

(Austin et al., 2018). Another consequence of industrialisation and intensification was a

concomitant change from nomadic to sedentary lifestyles in traditionally grassland areas such

as Mongolia, also fundamentally reshaping the landscape (Ilyashenko and King, 2018).

Notwithstanding that more intensive farming and higher yields may have prevented the

transformation of more than 80m ha of natural habitat to agriculture between 1960 and 2000

(Alexandratos and Bruinsma, 2012), expansion of cultivation has accelerated in the 21st

Century (Phalan et al., 2013). The main facilitators of this expansion have been advances in

technology, mechanisation and crop protection measures; contributing to the expansion of

arable agriculture onto formerly marginal and less fertile land. During the 20th

Century,

Page 25: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

25

agriculture expanded globally by around 50%, from around 1,200m ha to around 1,800m ha

(Ramankutty et al. 2002). Alexandratos and Bruinsma (2012) estimated that 4m ha of arable

land were added to the global cultivated area each year between 1961 and 2007. Galluzzi et

al. (2011) estimated that by 2050 global arable land will have increased by around 70m ha.

Farming now depends less on the intrinsic capability and ecological function of farmland

than ever before and as urban growth and with other more intensive land uses encroaching on

regions with the best climate and soils for growing crops, agriculture is pushing into more

marginal areas and fragile wildlife habitat (Avery, 1997), accelerating an already substantial

decline in biodiversity (Bignal and McCracken, 2000). Moreover, notwithstanding

modification by transfer payments to agriculture from other sectors of the economy, the

inexorable impact of Engels Law (Timmer et al., 1983), is determining ongoing increases in

farm size, loss of small farm operations, crop specialisation, and increasingly concentrated

monocultures. All these factors lead to simplification of landscapes, reduction or elimination

of natural habitat patches or edges and hence the diversity and abundance of farmland

wildlife. Alongside this, more efficient harvesting equipment and intensive cultivation are

reducing the amount of waste grain and hence post-harvest food availability for wildlife

(Sherfy et al., 2011).

With recent advances in agronomic and harvesting efficiency leading to fewer post-harvest

foraging opportunities for cranes, the potential for feeding on newly-sown grain, uprooting

young plants, trampling vegetation or nest building increases conflict with farmers

(Ilyashenko and King 2018). Moreover, over-abstraction of water from rivers and

groundwater, interference and disturbance, collisions with powerlines (especially when

located between roost and feeding sites) and poisoning by agrochemicals also increase with

more intensive land use.

Page 26: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

26

1.4 Agricultural Sustainability

Against this concerning backdrop, recent research effort has been deployed to identify ways

in which these trends can be mitigated or reversed. Food security is not only linked to an

increasing global population, it is also a function of rising personal income and changing diet,

which together suggest a growth in demand for food which would be extremely challenging

to meet through sustainable production (Benton, 2016). Ray et al. (2013) predict that global

crop production needs to double by 2050 to meet the projected demands from rising

population, diet shifts, and increasing biofuel consumption. Although maize, rice and wheat

yields have been respectively increasing at 1.6%, 1.0%, and 0.9% a year respectively;

significantly less than the 2.4% per year growth rate required to double global food

production by 2050.

The balancing of food security against biodiversity loss potentially involves one of the key

societal trade-offs of the 21st Century. Tscharntke et al. (2012) argue that smaller farms rather

than capital intensive, large-scale agri-businesses, are more likely to minimise these trade-

offs, without significant compromise. This is supported by Hanspach et al. (2017), who

conducted a multivariate analysis of social–ecological data from 110 landscapes and found

that either trade-offs or mutual benefits between food security and biodiversity were possible

but high degrees of food security in landscapes with developed infrastructure, market access

and capital came at the expense of biodiversity.

Benton (2016) notes that emissions from the food sector are significant contributors to

climate change, perhaps more so than any other sector; and yet, probably less than half the

Page 27: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

27

world's calories are used directly for healthy diets (Kamp et al., 2015), with over half of

agricultural production lost or wasted, fed to animals or consumed in excess of healthy

requirements. It is this excess from which cranes have benefitted that will be explored in

section 1.5 below. Newell and Taylor (2018) argue that because of the corporate structure and

political influence of agribusiness, as well as its research focus on larger scale of enterprises,

‘Climate Smart Agriculture’ is extremely difficult to implement at the smallholder level.

Consistent with this, Mahon et al. (2018) noted that ‘Sustainable Intensification’ can be

criticised as being too narrow in scope; failing to address all aspects of sustainability and with

an overemphasis on environmental elements and insufficient weight given to social and

cultural dimensions. Whitfield et al. (2018) have sought to more clearly codify what Climate

Smart Agriculture could mean but believe that the firmly established paradigm of growing

more, cheaper food, without fundamentally changing the role of commodity-based trade and

markets, is a major hurdle.

Another thread in this debate is whether ‘land-sparing’, restricting farmland to smaller,

higher-yielding areas and strictly conserving natural capital elsewhere, can lead to better

biodiversity outcomes than ‘land sharing’. Based on modelling, Eisner et al. (2017) argue that

higher yielding agricultural production does not necessarily lead to higher environmental

impacts. Phalan et al. (2016) support this position, and Balmford et al. (2018) maintain that

field data suggest that impacts on wildlife would be greatly reduced through boosting

yields on existing farmland and sparing remaining natural habitats.

Although this debate continues, as yet no definitive position around ‘land sharing’ vs.

‘land sparing’ has emerged. In the meantime, as agriculture continues to change, the

significant agriculturally-related challenges faced by cranes are increasing.

Page 28: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

28

1.5 Agriculture and Cranes

All of the world’s crane species make use of agricultural land (Harris and Mirande, 2013).

Some crane populations have benefitted significantly from agricultural expansion by foraging

on arable land; especially on grain spilt or remaining after harvest, when high energy forage

is sought for migration and wintering (Meine and Archibald, 1996). However, the emerging

technological and agronomic factors explored above are particularly impacting on some crane

species.

Anteau et al. (2011) found that changes in agricultural practice such as reduction in cropping

intervals, as well as the application of modern technology, can lead to a reduction of available

forage for cranes. Improved harvesting efficiency leads to dramatic reductions in post-harvest

waste grain, forcing cranes to seek new food sources at greater distances from roost sites,

resulting in lower reproductive success and increased crop damage through switching to

newly-sown and growing crops (Pearse et al., 2010). How cranes can adapt to agricultural

cropping and land use change is therefore a key conservation issue and management

consideration for farmers, conservation land managers and policy makers in crane range areas

(Harris and Mirande, 2013; Austin et al., 2018).

Crane populations exhibit a range of responses to agriculture, with some species’

relationships with agriculture severely impacted by the accelerating uptake of mechanisation

and improvements in agronomy (Wright et al., 2012). During the early 20th

Century Sandhill

Cranes Antigone canadensis became rare in many American states as a result of agricultural

expansion and intensification but have since recovered as a result of protection and

adaptation to modern agricultural practices. Following agricultural intensification, increased

grain production provided a number of crane species with forage at resting and wintering

sites, and high-energy combinable crops along their migration routes led to range expansion

Page 29: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

29

and increased populations in many countries (Krapu et al., 2004; Nowald, 2012; Nowald and

Mewes, 2010; Prange, 2015; Salvi, 2015). However, intensification can lead also to fewer

roosting sites (Barzen et al., 2012). Under optimal conditions, cranes prefer to feed in fields

close to their roosts, with the distance between foraging and roost sites only rarely exceeding

30 km (Prange, 2010; Anteau et al., 2011, Nilsson et al. 2016, Nevard et al. 2019a). Limiting

accessibility to roosts may therefore lead cranes to concentrate on fewer and larger ones -

increasing conflict with farmers in close proximity (Bouffard et al., 2005; Nowald and

Mewes, 2010).

The political collapse of the former Soviet Union brought about a number of changes in crane

behaviour and distribution, and led to a westward shift in migration and dramatically

increased crane numbers along the two key European flyways (Prange, 2012; Prange, 2015).

In Ukraine, crane numbers at Askania-Nova Nature Reserve and in adjacent agricultural areas

increased from approximately 5,000 in the late 1980s to approximately 20,000 –45,000 in the

2010s; whilst numbers decreased markedly in areas of agricultural abandonment (Redchuk et

al., 2015).

In Western Europe, increased maize production has corresponded with increased numbers of

Eurasian Cranes at staging and resting areas and an increase in the migratory population from

about 40,000–45,000 in 1980s to 350,000 birds by the winter of 2014/2015 (Prange, 2015).

Sweden, Norway, Denmark and Estonia also saw dramatic increases in numbers (Sandvik,

2010; Lundgren, 2013; Tofft, 2013; Leito, 2014). In Germany the breeding range of Eurasian

cranes doubled in the thirty years to 2010 (Mewes, 2010). In France, cereal growing now

occurs over the majority of the country and has led to increases in the number of wintering

cranes to about 150,000 and a significant northward shift in crane wintering grounds (Salvi,

2015).

Page 30: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

30

A similar response to agricultural expansion and intensification took place in the Western

Cape of South Africa, where the former extent of fynbos vegetation limited Blue Crane

Anthropoides paradisea distribution until the 1990s; when conversion to cereal production

created favourable breeding and feeding conditions (Scott and Scott, 1996). Although this

area only represented 7.6% of the Blue Crane’s range, by the early 2000s cereal-growing on

former fynbos areas supported 47.4% of the South African population (McCann et al., 2007).

As outlined above, contemporary agronomic practices, such as integrated pest management,

can induce high agricultural productivity from almost any marginal soil and land type and in

the past have led to greater availability of residual post-harvest grain for cranes. However,

improvements in agronomy and harvesting technology are now significantly reducing

leftover grain. Krapu et al. (2004) found that, largely due to improved harvesting efficiency,

whilst maize yields in central North America increased by 20% from 1978 to 1998, post-

harvest leftover grain averaged only 1.8% of production by 1998; representing a 47%

reduction in available forage for cranes.

In Western Europe during the 1970s and 1980s approximately 3–5% of wheat, barley and

oats, and 5–10% of maize was left on the ground post-harvest. More recently, only 1–2% of

other cereals and 0–1% of maize are typical levels of residual post-harvest grain, leading to

increased conflicts between cranes and farmers, with cranes foraging more intensively

amongst newly-sown crops (Nowald, 2012).

Krapu et al. (2004) also found that, as a result of shorter cropping intervals, between 77–97%

of post-harvest residual grain became unavailable to foraging birds. With the advent of pre-

emergent herbicides, new varieties of grain and intensification, farmers have begun to

cultivate immediately following harvest, to control weeds and bury residual crop stubbles,

Page 31: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

31

leaving only a short time for cranes to feed (Prange, 2012). Cranes are hence moving to

newly-sown crops and coming into direct conflict with farmers at the critical time for crop

establishment (Nowald and Mewes, 2010).

Cropping changes can also negatively impact on crane breeding habitat. In India, replacement

of subsistence rice cropping with cash crops such as sugarcane, cotton and tobacco eliminated

crane breeding habitat in the areas where these crops are grown (Borad and Mukherjee,

2002). Some districts in Nepal have seen increased sugarcane production, with Indian Sarus

Cranes A. a. antigone no longer breeding in the Krishnagar region (Aryal et al., 2009).

1.6 Australian Cranes

Australia has two crane species. Although remaining relatively abundant in the north, the

Brolga Antigone rubicunda has disappeared from much of its range in southern Australia

(Arnol et al., 2004). The Sarus Crane Antigone antigone is primarily an Asian species (Indian

sub-continent, Myanmar, Thailand, Indochina and formerly the Philippines) and is declining

in significant parts of its range but the Australian Sarus Crane A. a. gillae is thought to have

increased significantly in numbers over the past 50 years (Archibald et al., 2003). Both

Australia’s crane species are iconic cultural symbols for both European and Aboriginal

people across Australia’s tropical north but are increasingly coming into conflict with the

agricultural sector.

The apparent population growth of the Australian Sarus Crane since its first formal

description in the 1960s (Gill, 1967) has occurred in the presence of the congeneric Brolga

and within the context of the establishment and more recent expansion of cropping in

northern Australia (Archibald et al. 2003). Not only is there now extensive Brolga-Sarus

sympatry and interaction on cropping lands, hybrids (‘Sarolgas’) have been recorded amongst

Page 32: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

32

mixed flocks (Lofdahl, 1979; Archibald, 1981, Nevard et al. 2019b) and are known to be

fertile (Gray, 1958; Nevard et al., 2019b).

We do not yet know the drivers for this introgression but Templeton et al. (1990) identified

that changes in landscapes can be associated with genetic consequences and a hybrid zone

may be establishing in north Queensland. Imprinting studies suggest that hybrid young may

introgress with one or both parental species (Immelmann, 1972), eventually leading to

character convergence and displacement, or competitive exclusion. As agriculturally-driven

sympatry is so recent, adaptation is therefore still likely to be happening and may be

accompanied by evolutionary consequences.

1.6.1 The Brolga

The Brolga is an Australasian (Australia and New Guinea) endemic and autochthonous form.

Archibald’s (1976) work on unison calls found the Brolga call to be distinct from all other

cranes but most related morphologically and behaviourally to Sarus and White-naped cranes

Antigone vipio. Notwithstanding the c.100 year probable allopatry of northern and southern

Brolga populations, recent research by Miller et al. (2019) has identified only small genetic

differences (using materials gathered by this research).

Known to early European settlers as the ‘Native Companion’ due to its apparent propensity to

be seen in company with Aboriginal people, the Brolga has been recorded throughout the

State of Queensland (Marchant and Higgins, 1993) but is now only common and abundant in

the north. During the dry season, large flocks, sometimes numbering several thousand, can

gather at key sites such as the Cromarty Wetlands south of Townsville (Marchant and

Higgins, 1993). While breeding occurs in suitable areas throughout their Queensland range

Page 33: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

33

(Marchant and Higgins 1993; Barrett et al. 2003), it appears to be focussed on the Gulf Plains

and Cape York Peninsula.

Brolgas have traditionally been identified as a species of freshwater and brackish wetlands;

widely dispersed when breeding but gathering in flocks during the dry season to feed on

sedge tubers in seasonally dry swamps (Herring, 2005; Chatto, 2006; Herring, 2007;

Franklin, 2008). In north Queensland they concentrate, to varying extents, on the Atherton

Tablelands and east coast marshes (such as near Cromarty, Giru, Townsville and Ingham).

Brolgas can use brackish waters, having a unique salt-gland (Hughes and Blackman, 1969)

and when observed with Australian Sarus Cranes near Normanton, Brolgas favoured deeper,

generally coastal marshes (Walkinshaw, 1973), also occurring in open habitats, forest and

woodland.

Despite early work on the ecology of the Brolga in Australia’s tropics, conducted by Lavery

and Blackman (1969), little is known about the details of key breeding and flocking sites,

although recent work by Sundar et al. (2018) has gone some way to clarify this in relation to

the Gulf Plains. In north Queensland, breeding sites are found at their highest densities in the

Normanton-Karumba area in the southeast of the Gulf of Carpentaria (G. Archibald pers.

comm.; Sundar et al., 2018) stretching inland and north to Dunbar and Kowanyama. Cape

York Peninsula is also thought to hold significant concentrations of breeding Brolgas

(Marchant and Higgins, 1993; Chatto, 2006).

Key known flocking sites in Queensland are the Gulf of Carpentaria coastal plain, Cape York

Peninsula, and coastal freshwater wetland areas along the east coast from south of Townsville

to inland areas of Cairns, including the Atherton Tablelands (Collins, 1995; Chatto, 2006).

Many of these areas have only been identified from casual observations and much of the

Page 34: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

34

information (apart from the Atherton Tablelands) is based on occasional counts. The Brolga

population in northern Australia is thought to be stable and estimated to comprise of 20,000-

100,000 individuals (Meine and Archibald, 1996). A total of 51,969 Brolgas were recorded

by Kingsford et al., (2012), with all but 135 of these being from northern Australia.

The behaviour and biology of Brolgas in southern habitats is well documented, where their

wetland habitat requirements have also been studied (White 1987; Arnol et al. 1984; Harding

2001; Herring 2007; Myers 2001) but in Queensland these are poorly understood. Recent

work by Sundar et al. (2018) has identified that their diet in their Gulf Plain breeding areas is

less diverse than that of Australian Sarus Cranes.

1.6.2 The Australian Sarus Crane

Work by Archibald et al. (2003) and Jones et al. (2005) sought to clarify the taxonomic

position of the Sarus Crane and its subspecies. This has since been more thoroughly explored

as part of this research (Nevard et al., 2019b and in press), confirming that a distinct genetic

separation exists between the Australian subspecies A. a. gillae and the clinal Asian

subspecies A. a. antigone and A. a. sharpii.

As noted, the population of Australian Sarus Crane appears to have increased relatively

rapidly from what is thought to have been small numbers in northern Australia following its

formal identification in 1964 (Gill, 1967). G. Archibald (pers. comm.) estimates that the

population was c.200 in 1972 and current estimates suggest 5,000-10,000 individuals (E.

Scambler pers. comm.) but these are relatively loose. More specific and reliable counts

undertaken on the Atherton Tablelands indicate c.1,200-3,000 individuals flocking in the dry

season from 1997-2008 (as well as more recent counts to 2017) but it is currently not known

what proportion of the total Australian Sarus population these birds represent.

Page 35: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

35

Sarus Cranes in Australia are almost entirely confined to Queensland, with known breeding

areas along the Gulf of Carpentaria coastal plain, at least as far north as New Mapoon and

central Cape York Peninsula (Marchant and Higgins, 1993; Sundar et al., 2018); Elinor

Scambler (pers.comm.); Lavery and Blackman (1969) and Author (pers. obs.) found that

Australian Sarus Cranes had a strong preference for ‘upland’ rather than wetland foraging

areas, where they appear to be confined to freshwater. Many feed in flocks during the dry

season on recently harvested cropping land on the Atherton Tablelands (in the company of

Brolgas), migrating to breed in wetlands on the Gulf Plains and Cape York Peninsula. Sundar

et al. (2018) found that Australian Sarus Cranes preferred Eucalyptus-dominated regional

ecosystems for breeding and had a more varied diet than sympatric Brolgas but fed across a

narrower range of trophic levels (Nevard et al. 2019a).

1.6.3 The ‘Sarolga’

Archibald (1976) has identified a common phylogenetic history between Brolgas and Sarus

cranes, diverging from a common ancestor c.3 million years ago, suggesting significant

behavioural and genetic divergence. However, with the apparent significant population

growth of Australian Sarus Cranes since the 1960s occurring in the presence of the

congeneric Brolga, one explanation put forward has been that the Australian Sarus Crane

population may have been augmented by Brolga genes. Hybrid ‘Sarolgas’ have been

recorded in the wild (Nevard et al., 2019b), and are known from captive observations to be

fertile (Gray, 1958). J. Grant, (pers. comm.) has noted that the majority of Sarolgas he has

observed appear to have been paired with Brolgas, in at least two instances accompanied by a

juvenile. Of the 44,554 cranes surveyed during this research, 2.58% were identified as

hybrids, which are backcrossing (Nevard et al., 2019b).

Page 36: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

36

1.6.4 The Conservation Status of Australian Cranes

The IUCN Red List of Threatened Species lists the Brolga as being of Least Concern, as it

has a global population of more than 10,000 individuals and it is neither declining rapidly nor

restricted to a small range. The Brolga is not listed as threatened under the Australian

Environment Protection and Biodiversity Conservation Act 1999; however, its conservation

status varies from State to State and is listed as Threatened under the Victorian Flora and

Fauna Guarantee Act (1988). Under this Act, an Action Statement for recovery and future

management has been prepared. It is also included in the 2007 advisory list of threatened

vertebrate fauna in Victoria where it is listed as Vulnerable. The Sarus Crane is classified as

Vulnerable on the IUCN Red List on the basis of rates of decline in Asia (BirdLife

International, 2018c).

The current legal status of Brolgas and Australian Sarus Cranes affords a degree of

protection, but within their range in Queensland there are only a few protected areas used by

cranes, including Rinyirru (Lakefield) and Errk Oyangand (Staaten River) National Parks. On

the crane-critical Atherton Tablelands, only Hastie’s Swamp National Park and the Mareeba

Wetlands Nature Refuge have protected status.

1.7 Research Questions and Thesis Structure

Set against the background set out above, my thesis is framed to explore the distributional,

resource use, phylogenetic and agricultural relationships of Brolgas and Australian Sarus

Cranes. Its geographic focus is mainly on and around the Atherton Tablelands in north

Queensland but also extends to other parts of Australia and Asia. The key research questions

and the hypotheses I test are outlined below, followed by an explanation of the structure of

the thesis designed to address them.

Page 37: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

37

1.7.1 Research questions

Question 1: What are the evolutionary relationships between Australian Sarus Cranes and

subspecies in Asia?

Null hypothesis: Australian Sarus Cranes are likely to have not divereged from Asian

subspecies and to form a cline with them. Under this null hypothesis I would expect genetic

differentiation to be minimal and not significant.

Alternative hypothesis: Australian Sarus Cranes form a genetically distinct subspecies. Under

this alternative hypothesis, I would expect Australian Sarus Cranes to have diverged

significantly from all Asian subspecies and that there would be a discontinuity in variation

coinciding geographically with phenological variation.

Question 2: What is the propensity for introgression between Sarus Cranes and Brolgas?

Null hypothesis: Given their long period of divergence from a common ancestor, there are

sufficient behavioural and genetic obstacles to prevent significant introgression. Under this

null hypothesis I would expect introgression to be minimal and not a significant evolutionary

pathway.

Alternative hypothesis: Introgression is occurring at a rate sufficient to influence the

evolutionary trajectory of both Australian crane species. Under this alternative hypothesis I

would expect introgression to be significant and have the potential to increase.

Question 3: What are the distribution and resource partitioning characteristics of cranes on

the Atherton Tablelands?

Null hypothesis: Crane distribution and resource partitioning are not determined by foraging

opportunities, crop stage and range of agricultural crops. Under this null hypothesis I would

expect no significant correlation between foraging opportunities, crop stage and cropping

patterns.

Page 38: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

38

Alternative hypothesis: Crane distribution and resource partitioning are related to foraging

opportunities, cropping patterns and crop stage. Under this alternative hypothesis, I would

expect a strong relationship with foraging opportunities, cropping pattern, stage and crane

distribution.

Question 4: What are the attitudes of farmers and pastoralists towards cranes?

Null hypothesis: Farmers and pastoralists are not comfortable with cranes using their crops.

Under this null hypothesis I would expect farmers and pastoralists to take active measures to

exclude cranes.

Alternative hypothesis: Farmers and pastoralists tolerate cranes amongst some of their crops

but believe that forthcoming agronomic and cropping changes will reduce crane foraging

opportunities. Under this alternative hypothesis, I would expect farmers to be open to the

development of humane crane repellent measures and economic opportunities around crane-

based tourism.

1.7.2 Thesis Structure

To address these questions and associated hypotheses, my thesis is structured as follows:

Chapter 1 – Introduction, Literature Review and Thesis Structure: Paleoenvironmental

and agricultural contexts. In Chapter 1 I have described the evolutionary context of cranes

and their relationship with past, present and future agricultural development; including a brief

overview of the influence of the impacts of specific political, cultural, social and agronomic

changes on cranes. I have also presented an introduction to the biology, ecology and

conservation status of Brolgas and Australian Sarus Cranes.

Chapter 2 - Study Area and Methodology: I describe the region in which the research was

conducted (the Atherton Tablelands and Gulf Plains). I also describe the rationale for the

Page 39: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

39

adoption of the methodological approaches used, providing an outline of the detailed methods

set out in each of Chapters 4, 5, 6 and 7.

Chapter 3 - The Australian Sarus Crane: Genetic and phylogenetic significance of

Australian Sarus Cranes. In this chapter, submitted for publication in PLOS ONE, I describe

the phylogenetic relationships between the Australian Sarus Crane and other extant Sarus

Crane subspecies. I provide a genetic analysis based on a large range of samples from most

Sarus Crane range areas, as well as from the extinct Philippine subspecies.

Chapter 4 – Extent and Potential Trajectory of Introgression: Genetic and visual

evidence for hybridisation between the Brolga and Australian Sarus Crane. In this chapter,

published online in Oryx (Nevard et al., 2019b), I confirm the level of introgression between

the Brolga and Australian Sarus Crane and confirm their migration between the Atherton

Tablelands and Gulf Plains; using a range of genetic sample sources, including a particularly

large sample of recently shed feathers gathered on the Atherton Tablelands and in the Gulf

Plains.

Chapter 5 - Resource Partitioning and Competition: Distribution and foraging of Brolgas

and Sarus Cranes on the Atherton Tablelands. In this chapter, published online in Pacific

Conservation Biology (Nevard et al., 2019a), I provide a statistical analysis of observational

data, gathered over a three year period on the Atherton Tablelands; with definitive results in

relation to species distribution, foraging preferences and population.

Chapter 6 – Farmer Attitudes: Relationships between cranes, farmers and pastoralists on

the Atherton Tablelands. In this chapter, published online in Pacific Conservation Biology

(Nevard et al., 2018), I provide an analysis of farmer attitudes towards cranes, as well as an

Page 40: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

40

overview of their future cropping and agronomic intentions. I also explore the potential for

the development of a crane-based tourism sector.

Chapter 7 – Discussion: An exploration of research findings in relation to contemporary

issues and postulation of various potential future scenarios relating to cranes in north

Queensland.

Chapter 8 - Conclusions and Recommendations: In this final chapter I reiterate my

research conclusions and make recommendations for the conservation management of

Brolgas and Australian Sarus Cranes.

Page 41: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

41

CHAPTER 2 – STUDY AREA AND METHODOLOGY

2.1 Study Area and Geographical Scope

This research was mainly conducted on the Atherton Tablelands in Far North Queensland but

also refers to other parts of Australia and Asia (see Fig. 2.1 below).

Fig 2.1 Global distribution of Brolga Antigone rubicunda and Sarus Crane Antigone antigone (A),

field data collection sites in northern Queensland (B, C and D). Distribution data derived from

BirdLife International and NatureServe Bird Species Distribution Maps of the World (2014) and the

Australian Bird Guide (2017). Gulf Plains Interim Biological Regionalisation Area (IBRA) derived

from Department of Environment and Energy (2012) and state roads from Department of Transport

and Main Roads (2016).

Page 42: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

42

Brolgas and Sarus Cranes breed in the wet season on the Gulf Plains and in Cape York

Peninsula, and flock in the dry season on the Atherton Tablelands (see Fig. 2.2 below), which

was the principal location for this research.

Fig. 2.2 Atherton Tablelands region showing location, major settlements, land use and Key

Biodiversity Area (data from Queensland Globe and Department of Environment & Energy, 2012). .

The Atherton Tablelands extend from north of Mareeba and south of Innot Hot Springs, with

an area of approximately 6,000 km2 of which about 500 km

2 is intensively farmed. Elevation

ranges from 399 to 1,070 m above sea level, with a tropical mid-elevation to upland climate

and strongly seasonal rainfall, ranging from 900 to 2,200 mm. The Tablelands encompass

three main landscape elements: (i) rainforests (mainly part of the Wet Tropics World

Heritage Area); (ii) areas of dry sclerophyll savanna woodland; and (iii) some of Australia’s

most fertile and productive farmland (Department of Agriculture and Fisheries, 2015). Cranes

are absent from rainforest but occur in the latter two landscapes, mainly during the dry season

flocking months of May to December.

Page 43: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

43

Typical agricultural soils in the northern and southernmost areas of the Atherton Tablelands

are derived from granite and metamorphics and have a slightly acid pH and low fertility,

compared to the well-drained soils in the central Tablelands derived from basalt (Department

of Agriculture and Fisheries, 2015). Water for the Mareeba-Dimbulah Irrigation Area is

supplied through an extensive network of channels and streams from a large water

impoundment on the Barron River, Tinaroo Dam. Farming enterprises in other parts of the

Tablelands (including intensive agricultural areas around Atherton, Malanda and Ravenshoe)

draw water from volcanic groundwater aquifers, natural watercourses and smaller dams

(Department of Agriculture and Fisheries, 2015).

The Atherton Tablelands provide a regionally important dry season refuge for both Brolgas

and Australian Sarus Cranes. In the case of Australian Sarus Cranes, the area is currently

their only formally-identified dry season site (Archibald and Swengel, 1987) and as the Sarus

Crane is listed as Vulnerable by the IUCN (Birdlife International, 2018a), these factors were

key reasons for the designation of the Atherton Tablelands Key Biodiversity Area (Birdlife

International, 2018b); where both crane species have been counted annually by local BirdLife

Australia members since 1996.

2.2 Methodology

In order to be useful in relation to the multifaceted conservation problems facing cranes in

northern Australia, a multidisciplinary approach is required. This thesis uses observational,

genetic, GIS, socioeconomic and statistical tools to address this. As detailed methodologies

are provided in each of Chapters 3, 4, 5, and 6 a précis is provided below, along with

reference to where published.

Page 44: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

44

Chapter 3 - Subspecies in the Sarus Crane Antigone antigone revisited; with particular

reference to the Australian population. [Submitted for publication to PLOS ONE].

To thoroughly investigate the genetic relationships between Sarus Crane subspecies and

maximise the resolution of their genetic history, large numbers of blood and tissue samples

were assembled from most Sarus Crane range states and analysed using techniques

established in published genetic work on cranes. Ten amplified microsatellite loci were used

that had been developed for the Brolga (Miller et al., 2019), the Sarus Crane (Jones et al.,

2005) and other crane species (Hasegawa et al., 2000; Meares et al., 2009). Statistical

analyses used to calculate gene diversity (Nei, 1987) and allelic richness (Petit et al., 1998)

included FSTAT 2.9.3.2 (Goudet, 1995) and GenePop 4.2 (Raymond and Rousset, 1995;

Rousset, 2008). For population differentiation, microsatellite data were analysed by Bayesian

clustering, using STRUCTURE 2.3.4 (Pritchard et al., 2000; Falush et al., 2007). Structure

Harvester 0.6.94 (Earl et al., 2012) was used to analyse the data, following Evanno et al.

(2005). The most likely assignment of individuals to clusters was inferred in CLUMPAK

(Kopelman et al., 2015). K-means clustering was applied to validate the Bayesian approach

using GenoDive 2.0b23 (Meirmans and van Tienderen, 2004), as it is free of population

genetic assumptions (in contrast to STRUCTURE). Relationships among mitochondrial

haplotypes were analysed using statistical parsimony/TCS (Clement et al., 2002) and

implemented in PopART (Leigh and Bryant, 2015) and MrBayes 3.2.6 (Ronquist et al.,

2012).

Chapter 4 – Extent and Potential Trajectory of Introgression: The Sarolga: conservation

implications of genetic and visual evidence for hybridization between the Brolga Antigone

rubicunda and the Australian Sarus Crane Antigone antigone gillae. Oryx.

doi:10.1017/S003060531800073X.

Page 45: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

45

In order to gain a clear understanding of the extent of Brolga and Australian Sarus Crane

introgression, both visual surveys and genetic sampling were conducted across their flocking

areas of the Atherton Tablelands (see Chapter 5). Large numbers of freshly-shed feathers

were also collected opportunistically from sites in the Gulf Plains and on the southern edge of

the Atherton Tablelands. Using techniques described in Chapter 4, microsatellite loci were

tested which had been successfully used in the Sarus Crane (Jones et al., 2005), Brolga

(Miller, 2016) and other species of crane (Hasegawa et al., 2000; Meares et al., 2009). As

there was the possibility that feathers of a particular bird were collected more than once,

identical genotypes were identified using Cervus 3.0.7 (Kalinowski et al., 2007). In order to

identify the number of genetic clusters and potential hybrid individuals, Bayesian clustering

was conducted using STRUCTURE 2.3.4 (Pritchard et al., 2000; Falush et al., 2007).

Structure Harvester 0.6.94 (Earl and von Holdt, 2012) was used to analyse the data, following

Evanno et al. (2005) and Structure Plot V2.0 (Ramasamy et al., 2014). More detailed

analyses were conducted based on the clusters identified by Bayesian clustering using

FSTAT 2.9.3.2 (Goudet, 1995), GenePop 4.2 (Raymond and Rousset, 1995; Rousset, 2008),

and Micro-Checker 2.2 (van Oosterhout et al., 2004). In order to test whether shed feathers

might have a wider application in monitoring crane populations, application of the Lincoln-

Petersen Index was also undertaken for ‘genetically re-trapped’ birds; based on the

methodology described by Southwood and Henderson (2000).

Chapter 5 - Resource Partitioning and Competition: Agriculture, brolgas and Australian

sarus cranes on the Atherton Tablelands, Australia. Pacific Conservation Biology,

doi:10.1071/PC18081.

An observational approach was used to investigate crane distribution and foraging on the

Atherton Tablelands. Forty-six crane feeding sites, along a 335 km driven traverse, were

Page 46: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

46

surveyed on 44 occasions. The use of more resource-intensive techniques such as stable

isotope analysis was considered but rejected, as analyses of observational data were

considered to be capable of providing definitive findings in relation to differential numbers,

distribution, and foraging.

To evaluate whether species used the Atherton Tablelands differently over time, Permanova

was used and the Euclidean distance measure was employed to untransformed counts. To

determine whether crane species used different sites on the Atherton Tablelands, the rank

correlation between species was first examined (separately using each of abundance and

density). To measure spatial association of sites Moran’s I statistic was calculated, based on

the percentage of Brolgas at each site and using the Getis-Ord Gi* statistic to identify clusters

of sites with values higher (hotspot) or lower (coldspot) in magnitude than expected by

chance (Getis and Ord, 1992).

To consider the possibility that choice of feeding sites might be constrained by distance to

roosts, the distance between them was evaluated as a function of species. Effects of distance

class, species, and the distance class were then evaluated by species interaction in Permanova

with site as a random effect; using the Euclidean distance measure, a Type III Sum of

Squares model, and 9999 permutations. As the interaction term was statistically significant,

post-hoc pairwise comparisons of distances classes were conducted separately for each

species.

A regression tree analysis was run to examine species foraging preferences. To identify crop

type preferences, each crane species was considered in a separate model. The analysis was

performed in Permanova using the Euclidean distance measure, a Type III Sum of Squares

model, and 9999 permutations. To evaluate seasonal patterns of crop type availability,

Page 47: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

47

observed frequencies were compared to expected frequencies across months (May to

December), the latter based on constant proportionality calculated across all site surveys and

evaluated using Chi-squared analysis.

Chapter 6 – Farmer Attitudes: Farming and cranes on the Atherton Tablelands, Australia.

Pacific Conservation Biology, doi:10.1071/PC18055. 2018.

Information on the relationship between cranes and agriculture on the Atherton Tablelands is

critical for conservation management. Attitudes and socioeconomic conditions under which

farmers and graziers operate were explored using an interview methodology established by

Garnett et al. (2016).

Forty-five farmers and graziers, whose land was known to be used by cranes, described their

farming experience; the composition of their enterprises; their views on nature conservation;

knowledge of and attitudes towards cranes; damage by cranes and other wildlife; potential

solutions for reducing crop damage; and attitudes towards crane tourism and crop damage

mitigation trials. They also expressed their intentions for their enterprises in the future. Two

forms of analysis were conducted based on the landholder survey: (i) descriptive statistics

based on ranking to summarise knowledge and attitudes; and (ii) textual sampling of

comments made by individual interviewees to convey the tone of the interviews and provide

context.

Page 48: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

48

CHAPTER 3 - THE AUSTRALIAN SARUS CRANE

Subspecies in the Sarus Crane Antigone antigone revisited; with

particular reference to the Australian population. [Submitted for

publication to PLOS ONE 26/07/19].

3.1 Abstract

Subspecies are less well-defined than species but have become one of the basic units for

protection in law. Evidence for the erection or synonymy of subspecies therefore needs to be

based on the best science available. It is shown that there is clear genetic differentiation in the

Sarus Crane Antigone antigone where previously the variation had appeared to be clinal.

Based on a sample of 76 individuals, analysis of 10 microsatellite loci from 67 samples and

49 sequences from the mitochondrial control region, this research establishes that the

Australian Sarus Crane A. a. gillae differs significantly from both A. a. antigone (South Asia)

and A. a. sharpii (Myanmar and Indochina). A single sample from the extinct Philippine

subspecies A. a luzonica clustered with A. a. gillae, hinting at a more recent separation

between them than from A. a. antigone and A. a. sharpii even though A. a. sharpii is closer

geographically. The results demonstrate that failure to detect subspecies through initial

genetic profiling does not mean discontinuities are absent and has significance for other cases

where subspecies are dismissed on the basis of partial genetic evidence alone. It is also

potentially important for sourcing birds for reintroduction to the Philippines.

3.2 Introduction

Species are diagnosed along a continuum from phenotypically distinguishable but with

common ancestry, through to biological incompatibility (de Quieroz, 2007); with over 30

Page 49: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

49

definitions currently in use (Zachos, 2018). Subspecies are even less well defined with

diagnoses applied unevenly among taxa. Broadly they represent geographically defined

populations that are potentially incipient species diagnosable by at least one heritable trait

(Wallin et al., 2017) but while there have been attempts to define subspecies mathematically

(Baker et al., 2002, Patten and Unitt, 2002), debate continues (Patten and Remsen, 2017).

The hope that genetic analysis would resolve ambiguities has not eventuated. For example,

while cetacean biologists are content to define subspecies quantitatively on the basis of

mitochondrial DNA control region sequence data alone (Martien et al., 2017), this approach

has been rejected for birds (McCormack and Maley, 2015), not least because there is often

discordance between mitochondrial and nuclear DNA (Toews and Brelsford, 2012).

The definitions matter. This leaves open the point at which the best scientific information is

reached. A failure to recognise subspecies can mean they are lost before being recognised as

warranting conservation attention (Gippoliti et al., 2017). On the other hand, over-splitting

increases the probability of genetic problems among the necessarily smaller populations

identified (Frankham et al., 2012). Moreover, subspecies have become, with species, the

common currency of threatened species conservation in most jurisdictions (Garnett and

Christidis, 2007) with erection or synonymy of subspecies having legal, financial and social

consequences. For example, had the US Fish and Wildlife Service followed Zink et al. (2013)

and decided that the California gnatcatcher (Polioptila c. californica) did not warrant

subspecies status, 80,000 ha of its critical coastal sage scrub habitat would have been released

for development (Frost, 2015). In the end they decided otherwise, on the basis that the best

available scientific information did not support synonymy (US Fish and Wildlife Service,

2016).

Page 50: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

50

The Sarus Crane Antigone antigone [Linnaeus, 1758] exemplifies how further investigation

of variation has the potential to change both taxonomic treatment and the value given to

different populations, having geographically separate populations in southern Asia and

Australia (see Fig. 3.1).

Fig. 3.1 Global distribution of the Sarus Crane A. antigone, showing populations and subspecies

[Distribution data derived from BirdLife International and NatureServe Bird Species Distribution

Maps of the World (2014), the Australian Bird Guide (2017) and author contributions].

As it is extinct in the Philippines and declining in some of its Asian range, particularly in

Myanmar and Indochina (Meine and Archibald, 1996; Archibald et al., 2003), it is classed as

Vulnerable by the IUCN (BirdLife International, 2018c). Intraspecific variation within the

species has been the subject of ongoing debate. Blyth and Tegetmeier (1881) initially erected

the Indian and Myanmar birds as distinct species, based on plumage (the Indian Sarus Crane

has a white upper neck and tertials) and body size. Sharpe (1894) retained this distinction but

Page 51: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

51

shortly afterwards Blanford (1895) combined them into one species with two sub-species,

Grus antigone antigone and Grus antigone sharpii respectively, a classification which has

since endured. Hachisuka (1941) described the (then extant) Philippine population as Grus

antigone luzonica and distinct from both G. a. antigone and G. a. sharpii. Although the

taxonomic nomenclature A. a. sharpii has been followed for the South-east Asian subspecies

adopted by del Hoyo and Collar (2014), it is noted that the latinisation of Sharpe’s name as A.

a. sharpii is not grammatically correct and is more properly A. a. sharpeii.

In 1966, Sarus Cranes were discovered in Australia (Gill, 1967). Initially assigned to G. a.

sharpei (sic) because of a lack of white in the plumage, they were subsequently described by

Schodde (1988) as a new sub-species G. a. gilliae (sic), on the basis of distinct plumage and a

larger ear patch. Archibald (personal observation) notes that A. a. gillae also has different

unison calls from both A. a. antigone and A. a. sharpii, helping to differentiate it from the

sympatric Brolga A. rubicunda. Although he did not separate it by plumage, Hachisuka’s

(1941) description of the Philippine Sarus Crane population as the subspecies A. a. luzonica

was made on the basis of its significantly smaller size (in a range of measurement parameters

compared to A. a. antigone and A. a. sharpii). However, most recently, del Hoyo and Collar

(2014) have recognised only three subspecies (see Fig. 3.2): A. a. antigone (India), A. a.

sharpii (South-east Asia) and A. a. gillae (Australia); considering A. a. luzonica to have been

indistinguishable from A. a. sharpii.

Page 52: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

52

Fig. 3.2 Extant Sarus Crane subspecies: (a) antigone South Asia; (b) sharpii Myanmar and

Indochina; (c) gillae Australia. [Photographs T, Nevard (a) and (c); Robert van Zalinge (b)]

These relatively settled sub-specific arrangements, largely indicated by morphology, have not

hitherto been strongly supported by genetic analyses. Application of molecular techniques to

understand the subspecific arrangements of Sarus Cranes (Dessauer et al.,1992; Wood and

Krajewski, 1996; Jones et al., 2005) suggested that colonisation of Australia by Sarus Cranes

was relatively recent and that there had been little differentiation of populations across their

range (Jones et al., 2005).

The potential ancestral relationship with other populations is relevant for two reasons. The

first of these is as a contribution to the debate about using genetics as the sole basis for

synonymising or retaining subspecies; particularly whether the different subspecies should

continue to be treated as separate, valued management units - although urged regardless of

the taxonomy (Jones et al., 2005). Second, reintroduction of Sarus Cranes to the Philippines

is being considered (J. C. Gonzalez, pers. comm.), so an appropriate potential founding stock

needs to be identified.

Page 53: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

53

3.3 Methods

3.3.1 Samples

In order to measure the degree of differentiation between Sarus Crane populations, material

was secured opportunistically from all four putative subspecies and most range countries (see

Tables 3.1and 3.2 below). Sample material varied from naturally shed and deliberately

plucked feathers, to blood taken from live birds (both wild and captive) and tissue harvested

from natural mortalities and museum specimens.

Only samples from captive birds in Germany and Australia (Lemgo Crane Collection and

Cairns Tropical Zoo) and feathers from crane flocking sites in Australia and museum

specimens in the United States were specifically collected for this project. All other samples

assembled were derived from sets previously collected as part of other projects in Myanmar

(captive zoo and monastery birds), Thailand (captive zoo birds), and Cambodia (wild-caught

birds).

The blood sample collection protocol for captive German and Australian birds was for a

three-person restraining team (all with significant experience in crane restraint and sampling)

to catch the bird using a landing net; followed by immediate hooding and drawing ≤1mm of

blood from the brachial vein (placing this immediately into 100% ethanol). In all cases

restraint lasted less than 2 minutes. In Myanmar and Thailand, although sample collection

was not part of this project, the collection protocol was consistent. In Cambodia, birds were

wild caught using alpha chloralose, as part of a previous project and sampled as above, with

an additional oral swab. One tissue (brain) sample from Cambodia was from a bird that had

died recently from natural causes.

Page 54: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

54

Shed feathers visibly free of soil and/or faecal contamination were gathered from crane

flocking sites in Australia using tweezers and re-sealable plastic bags then refrigerated.

Lightly-plucked chest feathers (3 to 5, ≤ 25mm) were obtained from restrained captive birds

in Germany and Australia and refrigerated.

Table 3.1 Sources of Sarus Crane DNA used in analyses.

Subspecies Country

Sample

type

No. DNA samples

Nuclear Mitochondrial

antigone India Blood 5 3

Toe pad 3 1

gillae Australia Feather 25 20

luzonica Philippines Toe pad 1 -

sharpii Cambodia Brain 1 -

Blood 12 6

Oral swab 1 -

Myanmar Blood 11 11

Thailand Blood 8 8

Total 67 49

Table 3.2 Details of sample collections for Sarus Crane (Ssp: subspecies; M: microsatellites; S:

sequences)

Code Ssp Country Locality

Collection

number

DNA

source M/S Aus01 gillae Australia Gulf Plains B1974 feather M/S

Aus02 gillae Australia Gulf Plains B1976 feather M/--

Aus03 gillae Australia Gulf Plains B1980 feather M/S

Aus04 gillae Australia Gulf Plains B1986 feather M/--

Aus05 gillae Australia Gulf Plains B1989 feather M/--

Aus06 gillae Australia Gulf Plains B2168 feather M/--

Aus07 gillae Australia Gulf Plains B2216 feather M/S

Aus08 gillae Australia Gulf Plains B2220 feather M/--

Aus09 gillae Australia Gulf Plains B2225 feather M/--

Aus10 gillae Australia Gulf Plains B2228 feather M/--

Aus11 gillae Australia Gulf Plains B2233 feather M/--

Aus12 gillae Australia Gulf Plains B2234 feather M/--

Aus13 gillae Australia Gulf Plains B2239 feather M/--

Aus14 gillae Australia Gulf Plains B2241 feather M/S

Aus15 gillae Australia Gulf Plains B2243 feather M/--

Aus16 gillae Australia Gulf Plains B2245 feather M/--

Aus17 gillae Australia Gulf Plains B2247 feather M/--

Aus18 gillae Australia Gulf Plains B2248 feather M/--

Aus19 gillae Australia Gulf Plains B2249 feather M/--

Aus20 gillae Australia Gulf Plains B2254 feather M/S

Aus21 gillae Australia Gulf Plains B2255 feather M/--

Aus22 gillae Australia Gulf Plains B2258 feather M/S

Page 55: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

55

Code Ssp Country Locality

Collection

number

DNA

source M/S Aus23 gillae Australia Gulf Plains B2261 feather M/--

Aus24 gillae Australia Gulf Plains B2352 feather M/--

Aus25 gillae Australia Gulf Plains B2380 feather M/--

Aus26 gillae Australia Gulf Plains B1922 feather --/S

Aus27 gillae Australia Gulf Plains B1926 feather --/S

Aus28 gillae Australia Gulf Plains B1927 feather --/S

Aus29 gillae Australia Gulf Plains B1983 feather --/S

Aus30 gillae Australia Gulf Plains B2008 feather --/S

Aus31 gillae Australia Gulf Plains B2009 feather --/S

Aus32 gillae Australia Gulf Plains B2015 feather --/S

Aus33 gillae Australia Gulf Plains B2196 feather --/S

Aus34 gillae Australia Gulf Plains B2198 feather --/S

Aus35 gillae Australia Gulf Plains B2218 feather --/S

Aus36 gillae Australia Gulf Plains B2224 feather --/S

Aus37 gillae Australia Gulf Plains B2250 feather --/S

Aus38 gillae Australia Gulf Plains B2251 feather --/S

Aus39 gillae Australia Gulf Plains B2327 feather --/S

Cam01 sharpii Cambodia

Siem Reap, Angkor Centre for

Conservation of Biodiversity KHL15-ACCB-

006-Br-RNA brain M/--

Cam02 sharpii Cambodia Tonle Sap floodplains

KHL16-

ZALINGE-002 blood M/--

Cam03 sharpii Cambodia Tonle Sap floodplains

KHL16-

ZALINGE-003 blood M/--

Cam04 sharpii Cambodia Mekong delta region

KHL16-

ZALINGE-005 blood M/--

Cam05 sharpii Cambodia Mekong delta region

KHL16-

ZALINGE-006 blood M/--

Cam06 sharpii Cambodia Tonle Sap floodplains

KHL16-

ZALINGE-008 blood M/S

Cam07 sharpii Cambodia Tonle Sap floodplains

KHL16-

ZALINGE-009 blood M/--

Cam08 sharpii Cambodia Tonle Sap floodplains

KHL16-

ZALINGE-011 blood M/S

Cam09 sharpii Cambodia

Siem Reap, Angkor Centre for

Conservation of Biodiversity

KHL16-

ZALINGE-KH003 blood M/S

Cam10 sharpii Cambodia Tonle Sap floodplains

KHL16-

ZALINGE-R/W blood M/S

Cam11 sharpii Cambodia Tonle Sap floodplains

KHL16-

ZALINGE-001 blood M/S

Cam12 sharpii Cambodia

Siem Reap, Angkor Centre for

Conservation of Biodiversity

KHL16-

ZALINGE-KH001 blood M/--

Cam13 sharpii Cambodia

Siem Reap, Angkor Centre for

Conservation of Biodiversity

KHL16-

ZALINGE-KH002 blood M/S

Cam14 sharpii Cambodia

Siem Reap, Angkor Centre for

Conservation of Biodiversity ACCB-A0058-1

oral

swab M/--

Ind01 antigone India

Private collection:

Lemgo/Germany B1920 blood M/--

Ind02 antigone India

Private collection:

Lemgo/Germany B1921 blood M/--

Ind03 antigone India

Private collection:

Lemgo/Germany B2192 blood M/S

Ind04 antigone India

Private collection:

Lemgo/Germany B2822 blood M/S

Ind05 antigone India

Private collection:

Lemgo/Germany B2823 blood M/S

Page 56: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

56

Code Ssp Country Locality

Collection

number

DNA

source M/S Ind06 antigone India NMNH, Washington DC USNM64453 toe pad M/S

Ind07 antigone India Cincinnati Zoo CinB299851 toe pad M/--

Ind08 antigone India Cincinnati Zoo CinB299851 toe pad M/--

Mya01 sharpii Myanmar Minbya, Rakhine State IPMB64768 blood M/S

Mya02 sharpii Myanmar Maubin, Ayeyarwady Region IPMB64769 blood M/S

Mya03 sharpii Myanmar Maubin, Ayeyarwady Region IPMB64770 blood M/S

Mya04 sharpii Myanmar Maubin, Ayeyarwady Region IPMB64771 blood M/S

Mya05 sharpii Myanmar Einme, Ayeyarwady Region IPMB64772 blood M/S

Mya06 sharpii Myanmar Einme, Ayeyarwady Region IPMB64773 blood M/S

Mya07 sharpii Myanmar Nay Pyi Taw Zoo, IPMB64774 blood M/S

Mya08 sharpii Myanmar Nay Pyi Taw Zoo, IPMB64775 blood M/S

Mya09 sharpii Myanmar Yadanapon Zoo IPMB64776 blood M/S

Mya10 sharpii Myanmar Minbya, Rakaine State IPMB64780 blood M/S

Mya11 sharpii Myanmar Minbya, Rakaine State IPMB64781 blood M/S

Phi01 luzonica Philippines NMNH, Washington DC USNM256982 toe pad M/--

Tha01 sharpii Thailand Korat Zoo 275 blood M/S

Tha02 sharpii Thailand Korat Zoo 280 blood M/S

Tha03 sharpii Thailand Korat Zoo 282 blood M/S

Tha04 sharpii Thailand Korat Zoo 283 blood M/S

Tha05 sharpii Thailand Korat Zoo 288 blood M/S

Tha06 sharpii Thailand Korat Zoo 292 blood M/S

Tha07 sharpii Thailand Korat Zoo 294 blood M/S

Tha08 sharpii Thailand Korat Zoo 295 blood M/S

3.3.2 Extraction

3.3.2.1 Nuclear DNA

DNA was extracted using the SDS/salting-out protocol of Miller et al. (1988). Dithiothreitol

and Roti-PinkDNA (Carl Roth, Karlsruhe, Germany) were added in order to increase the

yield. For the Cambodian blood and tissue samples, QIAGEN’s RNeasy Mini Kit was used,

for an oral swab the QIAamp Viral RNA Kit. For the Thailand blood samples, DNA was

extracted by using QIAGEN’s RNeasy Mini Kit. We amplified the ten microsatellite loci

(Gamµ3, 18, 24,101b; GjM8, 13, 15, 48b; GR22, 25) used in our analysis of hybridization of

the Brolga (Antigone rubicunda) and Australian Sarus Crane (A. antigone gillae) (Nevard et

al., 2019b) which have been developed for other crane species (Hasegawa et al., 2000;

Meares et al., 2009), the Sarus Crane (Jones et al., 2005) and the Brolga (Miller et al., 2019),

respectively. PCRs conducted in a volume of 10 µl contained 1 µl DNA (10-25 ng), 1 µl of

10 x NH4-based Reaction Buffer, 1.5-2.25 mM MgCl2 Solution (Table 2), 0.25 mM of each

primer, 0.2 mM of dNTP, 0.04 µl of BioTaq DNA Polymerase (5 U/µl), 0.6 µl of 1 % BSA

Page 57: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

57

and sterile ddH2O. If not successful with this first protocol, the MyTaq mix (all products from

BIOLINE, London, UK) was used. The PCR profile started with an initial denaturation at

94°C, followed by 36 cycles of denaturation at 94°C, primer specific annealing (Table 2) and

extension at 72°C each for 30 s, and a final elongation at 72°C for 30 min. Microsatellite

alleles were separated on a 3130xl Genetic Analyzer using the GeneScan 600 LIZ Size

Standard 2.0. Fragment sizes were determined manually in GeneMapper 4.0 (all three

products from Applied Biosystems, Waltham, USA) as automatic calling with arbitrarily

predefined bin width may give inconsistent results. In order to maximize accuracy of size

determination, we repeated PCRs of samples with initially weak signals or which had rare

variants. Eventually, PCR samples peculiar to different runs had to be loaded on the same

plate to improve comparability.

3.3.2.2 Mitochondrial DNA

Where DNA quality allowed, we also sequenced large parts of copy 2 of the mitochondrial

control region (Akiyama et al. 2017) using primers L16707 and H1247 (Krajewski et al.,

2010) spanning a fragment of c. 1000 bp in A. antigone and 1150 bp in three specimens of A.

rubicunda; one from the Gulf plains and two from the Atherton Tablelands (see Nevard et al.,

2019b), which we used as outgroup in the phylogenetic analyses. In some specimens we had

to target a shorter fragment using the forward primer L514 instead resulting in lengths of

c.610 bp. PCRs were conducted using the MyTaq mix. The temperature profile for the long

fragment comprised: 95°C for 3 min, 5 cycles of 95°C/15 s, 65°C/20 s and 72°C/25 s, 5

cycles 95°C/15 s, 60°C/20 s and 72°C/25 s, 30 cycles 95°C/15 s, 55°C/20 s and 72°C/25 s,

and a final extension at 72°C for 5 min. For the short fragment the profile was similar and

had 4, 4 and 32 cycles with respective annealing temperatures of 60°C, 55°C and 50°C. PCR

products were cleaned using an exonuclease I/shrimp alkaline phosphatase mix. Cycle

Page 58: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

58

sequencing was then performed in 10 µl using the PCR primers and ABI’s Big Dye

Terminator Ready Reaction Mix 3.1 of which 50% were replaced by halfBD (Merck). The

thermal cycler profile followed the manufacturer’s suggestions except that the annealing

temperature was lowered to 48° C. HighPrep DTR magnetic beads (Biozym) were used for

purification of the sequencing reactions. The sequences were read on an ABI 3130xl Genetic

Analyzer (Thermo Fisher Scientific Inc., Waltham, MA, USA). Raw sequences were edited

in Geneious 10 (www.geneious.com) and BioEdit 7.0.5.3 (Hall, 1999), respectively, and

aligned using the web version of MAFFT 7 (Katoh et al., 2017).

3.3.3 Statistical analysis

3.3.3.1 Nuclear DNA

FSTAT 2.9.3.2 (Goudet, 1995) as well as GenePop 4.2 (Raymond & Rousset, 1995; Rousset,

2008) were used to test the microsatellite data for Hardy-Weinberg equilibrium and to

calculate gene diversity (Nei, 1987) and allelic richness (Petit et al., 1998) of subspecies. For

population differentiation, the microsatellite data were analysed in two ways, (i) in a divisive

approach without a priori designation of subspecies by Bayesian clustering using

STRUCTURE 2.3.4 (Pritchard et al., 2000; Falush et al., 2007); and (ii) by estimating

differentiation of the nominal subspecies (except the single individual of A. a. luzonica),

calculating pairwise FST values using FSTAT. STRUCTURE was run with K (number of

clusters) ranging from one to six and ten replicates assuming the admixture model since the

sequence analyses suggested that there had been admixture. We modelled both, uncorrelated

and correlated allele frequencies as it was unclear which approach provided a better fit for the

biological context. The Markov chains ran for 1 million generations after a burn in of

100,000. Structure Harvester 0.6.94 (Earl et al., 2012) was used to analyse the data following

Evanno et al. (2005). Integrating the results of the replicated runs in STRUCTURE, the most

Page 59: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

59

likely assignment of individuals to clusters was inferred in CLUMPAK (Kopelman et al.,

2015). K-means clustering was applied to validate the Bayesian approach using GenoDive

2.0b23 (Meirmans and van Tienderen, 2004) because it is free of population genetic

assumptions in contrast to STRUCTURE. Individuals were clustered based on their allele

frequencies according to the pseudo-F-statistic of Calinski and Harabasz (1974) as described

in Meirmans (2012). Finally, we estimated gene flow among subspecies based on FST and the

private alleles approach of Barton and Slatkin (1986) using GenePop (see Slatkin and Barton,

1989; Whitlock and McCauley, 1999).

3.3.3.2 Mitochondrial DNA

Relationships among mitochondrial haplotypes were analysed using statistical

parsimony/TCS (Clement et al., 2002) implemented in PopART (Leigh and Bryant, 2015)

and MrBayes 3.2.6 (Ronquist et al., 2011), respectively. MrBayes was run using GTR+I+G

identified as best fitting substitution model by jModeltest 2.1.4 (Darriba et al., 2012) with

default settings over 2 million generations with a 25% burnin. Effective sample sizes were >

700, potential scale reduction factors equalled 1.000 or 1.001, and the standard deviation of

split frequencies was < 0.006 indicating convergence of parameter estimates and both parallel

runs.

3.4 Results

The nominate subspecies had the highest diversity despite the lowest sample size, while A. a.

gillae had comparatively lower diversity than the nominate subspecies (Table 3.3). A. a.

antigone had four private alleles, two of them rare (only in one individual each and only

heterozygous), A. a. gillae five, four of them rare (each in not more than 2 specimens and

only heterozygous), and A. a. sharpii eight.

Page 60: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

60

Table 3.3 PCR specifications and diversity of microsatellite loci. The subspecies are abbreviated by

the first three letters (ant: A. a. Antigone; gil: A. a. gillae; sha: A, a. sharpii)

N alleles Gene diversity Allelic richness

Locus/dye

MgCl2

[mM]

T

[°C] ant gil sha ant gil sha ant gil sha

Gamµ3/FAM 1.5 59 5 2 5 0.839 0.115 0.768 4.741 1.570 4.364

Gamµ18/HEX 1.5 53 2 1 1 0.321 0.000 0.000 1.993 1.000 1.000

Gamµ24/HEX 2.25 60 7 5 5 0.567 0.227 0.534 4.000 2.290 3.291

Gamµ101b/HEX 2 59 8 2 7 0.821 0.393 0.799 5.399 1.985 5.048

GjM8/FAM 2 59 5 2 3 0.125 0.115 0.425 1.750 1.570 2.526

GjM13/HEX 2 60 5 2 4 0.875 0.488 0.591 4.849 1.999 3.209

GjM15/FAM 2 59 6 4 4 0.696 0.623 0.494 2.999 3.289 3.225

GjM48b/HEX 2 56 4 2 4 0.393 0.040 0.458 1.999 1.240 2.966

GR22/HEX na 60 5 2 4 0.491 0.487 0.412 2.749 1.999 2.980

GR25/Cyanine3 na 60 3 3 2 0.339 0.486 0.496 1.993 2.237 1.999

Mean 5.0 2.5 3.9 0.547 0.297 0.498 3.247 1.918 3.062

Of these, five were rare (each in not more than three individuals and four only heterozygous).

Three of the private alleles occurred only in Myanmar and another three in both Cambodia

and Thailand. The single A. a. luzonica sampled had one allele that did not occur in any other

subspecies. Deviations from the Hardy-Weinberg equilibrium at several loci in A. a. antigone

and A. a. sharpii suggested that these subspecies are probably not panmictic. The same was

true analysing the samples by country (data not shown). This was confirmed from the results

of analysis using STRUCTURE and K-means clustering.

According to Evanno et al.’s (2005) ΔK criterion and assuming correlated allele frequencies,

STRUCTURE identified three clusters, modelling independent allele frequencies only two

(Fig. 3.3; cluster composition as summarized by CLUMPAK Table 3.4).

Page 61: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

61

Table 3.4 Cluster analyses (codes from Table 3.2 above).

Bayesian clustering assuming population admixture

k-means clustering Correlated allele frequencies

Independent allele

frequencies Aus01 Cam05 Cam01 Aus01 Cam01 Aus01 Aus08

Aus02 Cam06 Cam02 Aus02 Cam02 Aus02 Aus20

Aus03 Cam08 Cam03 Aus03 Cam03 Aus03 Cam01

Aus04 Cam13 Cam04 Aus04 Cam04 Aus04 Cam02

Aus05 Cam14 Cam07 Aus05 Cam05 Aus05 Cam03

Aus06 Ind01 Cam09 Aus06 Cam06 Aus06 Cam04

Aus07 Ind02 Cam10 Aus07 Cam07 Aus07 Cam05

Aus08 Ind03 Cam11 Aus08 Cam09 Aus09 Cam06

Aus09 Ind04 Cam12 Aus09 Cam10 Aus10 Cam07

Aus10 Ind05 Mya03 Aus10 Cam11 Aus11 Cam09

Aus11 Ind06 Mya06 Aus11 Cam12 Aus12 Cam10

Aus12 Ind07 Tha01 Aus12 Cam13 Aus13 Cam11

Aus13 Ind08 Tha03 Aus13 Cam14 Aus14 Cam12

Aus14 Mya01 Tha04 Aus14 Ind02 Aus15 Cam13

Aus15 Mya02 Tha05 Aus15 Ind04 Aus16 Cam14

Aus16 Mya04 Tha06 Aus16 Ind05 Aus17 Ind01

Aus17 Mya05 Aus17 Ind06 Aus18 Ind02

Aus18 Mya07 Aus18 Ind07 Aus19 Ind03

Aus19 Mya08 Aus19 Ind08 Aus21 Ind04

Aus20 Mya09 Aus20 Mya01 Aus22 Ind05

Aus21 Mya10 Aus21 Mya02 Aus23 Ind06

Aus22 Mya11 Aus22 Mya03 Aus24 Ind07

Aus23 Tha02 Aus23 Mya04 Aus25 Ind08

Aus24 Tha07 Aus24 Mya07 Cam08 Mya01

Aus25 Tha08 Aus25 Mya08 Phi01 Mya02

Phi01 Cam08 Mya09 Mya03

Ind01 Mya10 Mya04

Ind03 Mya11 Mya05

Mya05 Tha01 Mya06

Mya06 Tha02 Mya07

Phi01 Tha03 Mya08

Tha04 Mya09

Tha05 Mya10

Tha06 Mya11

Tha07 Tha01

Tha08 Tha02

Tha03

Tha04

Tha05

Tha06

Tha07

Tha08

Page 62: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

62

Fig. 3.3 Bayesian cluster analyses of microsatellite data (STRUCTURE/CLUMPAK). A, B: assuming

admixture and correlated allele frequencies; y axis, membership coefficient from 0 (bottom) to 100

(top); in A, individuals are sorted by subspecies and country of origin; in B, samples are ordered by

cluster, subspecies within clusters and membership coefficient Q (y-axis). For A. a. sharpii, the

country of origin is indicated by the first letter (C = Cambodia, M = Myanmar, T = Thailand). C:

assuming admixture and independent allele frequencies, samples ordered by subspecies and country

of origin within A. a. sharpii. Cluster 1 consists of all A. a. gillae as well as specimens labelled with 1.

In both analyses, all A. a. gillae fell into one cluster together with the A. a. luzonica

specimen. Assuming independent allele frequencies, the cluster with these subspecies also

contained three specimens of A. a. sharpii and two Indian individuals. For both models, a

solution with four clusters had the highest likelihood but the composition of the clusters was

less meaningful, apart from grouping all Australian individuals together. K-means clustering

also divided the sample set into two clusters (Table 3.4), one consisting of 23 A. a. gillae, one

Page 63: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

63

A. a. sharpii, and the single A. a. luzonica, and the other of two A. a. gillae, all A. a. antigone

from India, and the remaining A. a. sharpii. Both Bayesian clustering (assuming admixture

and independent allele frequencies) as well as K-means clustering converged to very similar

solutions. The STRUCTURE bar plots also reflect the higher genetic diversity in the Asian

subspecies as summarized by the standard population genetic parameters above and in Table

3.2.

Differentiation among subspecies based on FST estimates revealed that A. a. antigone and A.

a. sharpii were considerably closer to each other (FST = 0.086) than either were to A. a. gillae

(FST = 0.282 and 0.168, respectively). These FST values translated into gene flow estimates of

2.66 migrants per generation between A. a. antigone and A. a. sharpii, 0.64 between the

nominate subspecies and A. a. gillae, and 1.24 between A. a. sharpii and A. a. gillae. The

private alleles approach estimated 0.71, 0.44, and 0.53 migrants, respectively. This again

emphasises the somewhat isolated position of the Australian subspecies.

The alignment of the control region 2 sequences comprised 1155 positions. A major

difference between A. rubicunda and A. antigone were two indels comprising 46 and 95

positions, respectively, which were present in the former and absent in the latter species.

Apart from these, A. rubicunda differed by at least 28 mutations from A. antigone, rendering

the latter monophyletic in the Bayesian tree reconstruction (Fig. 3.4), which is also illustrated

by the TCS network (Fig. 3.5).

Page 64: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

64

Fig. 3.4 Phylogenetic tree for three subspecies of the Sarus Crane A. Antigone; based on mtDNA

sequences resulting from Bayesian analysis. Posterior probabilities are given if > 0.70 (for sample

codes see Ch. III Table 2; outgroup (A. rubicunda) pruned off.).

Page 65: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

65

Fig. 3.5 TCS network for five populations of the Sarus Crane A. antigone based on mtDNA sequences

compared to the Brolga A. rubicunda (for sample codes see Table 3.2)

However, both tree and network agreed that no subspecies of A. antigone was monophyletic.

Both reconstructions suggested an ancestral polymorphism and/or repeated introgression,

meaning that there had been at least limited gene flow among the subspecies. Given the

overall low differentiation across A. antigone, resulting in low posterior probabilities (i.e.

node support) and the low sample size of the nominate subspecies, inferring evolutionary

trajectories is not possible.

Page 66: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

66

3.5 Discussion

Our analyses differ from earlier work (Woods and Krajewski, 1996; Jones et al., 2005) by

having a larger sample size and in sequencing a highly variable part of the mitochondrial

control region (Krajewski et al.,2010) instead of protein coding genes (Woods and

Krajewski, 1996), thereby providing better phylogenetic resolution. Similarly to Woods and

Krajewski (1996), we found that Sarus Crane subspecies and populations were not

monophyletic (probably due to an ancestral polymorphism and/or introgression) and

microsatellite variation in A. a. antigone and A. a. sharpii overlapped significantly (Jones et

al. 2005). However, we have established that A. a. gillae is far more distinct from A. a.

antigone and A. a. sharpii than previously thought, irrespective of the clustering method and

the model assumptions used in Bayesian clustering. This was also confirmed by F-statistics

and gene flow estimates. We have also shown that the single A. a. luzonica specimen we have

hitherto been able to sample was more similar to A. a. gillae than the geographically closer A.

a. sharpii. The first finding has potential implications for definitions of subspecies, the

second in relation to better understanding the phylogeography of the species and potential

sourcing of birds for any Philippine reintroduction. We are well aware how problematic any

conclusions based on a single specimen might be but given that Philippine Sarus Cranes are

extinct and the scarcity of museum material, no alternative approach is available.

Given there are now attempts to define subspecies under law (Mallett et al., 2018), there is a

need for far greater understanding of just how much weight should be given to genetic data,

particularly where genetic variation appears to be lacking. While patterns of crane

morphological variation have not been reflected in marked genetic differences between

populations (Rhymer et al., 2001; Haase and Ilyashenko, 2012; Mudrik et al., 2015), the

findings of Jones et al. (2005) could have been used to suggest that the variation in Sarus

Page 67: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

67

Cranes is clinal without distinct breaks in genetic variability. Our results suggest that simply

by looking at a slightly different part of the genome with a larger sample size a different

conclusion would have been drawn. This is relevant for current Australian policy, which has

not been consistent. For example, Schodde and Mason (1999) diagnosed new subspecies of

Southern Emu-wren Stipiturus malachurus and Eastern Bristlebird Dasyornis brachypterus

on the basis of morphological discontinuities. Despite genetic differences in the emu-wren

failing to match morphology (Donellan et al., 2009), threatened subspecies continue to be

recognised under legislation (Department of Environment & Energy, 2018). A similar level

of variation in the Eastern Bristlebird (Roberts et al., 2011) has meant that there has been no

recognition of the northern form of Eastern Bristlebird D. b. monoides (Schodde and Mason,

1999), of which 40 individuals are thought to survive (Stone et al., 2018) but for which

conservation effort has been inconsistent (Guerrero et al., 2017). Were rigid definitions of

subspecies enforced by law, as now being argued in the USA (Mallett et al., 2018), with the

level of knowledge previously available from Jones et al.(2005), Sarus Crane subspecies

might not have been eligible for conservation as separate subspecies. Our results confirm the

position of Patten and Remsen (2017) that synonymising subspecies can be highly

problematic without testing hypotheses using multiple data sources, as advocated by

integrative taxonomy (Dayrat, 2005). We therefore wish to stress that our findings are not a

final verdict on phylogeographic differentiation in the Sarus Crane, which requires further

morphological and genetic work to complement our analyses.

Although first formally noted in Australia in the 1960s (Gill, 1967), Sarus Cranes have been

in the country long enough to have been given a Wik (Cape York Aboriginal language group)

name, meaning ‘the Brolga that dipped its head in blood’ (Reardon 2007). Wood and

Krajewski (1996) have suggested that Sarus Cranes first arrived in Australia 37,500 years

Page 68: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

68

ago, when sea levels were 40 m lower than today (Zong et al., 2002), permitting development

of a savanna corridor that extended both north and south of the equator across the Sunda plain

(Bird et al., 2005), much of which would have been lost by rising sea levels at the start of the

Holocene (about 10,000 years ago). In an Australian context, this corridor (Joseph et al.,

2019) would have ended not far north of the Pleistocene Lake Carpentaria, around which it is

thought there would have been savannas structurally similar to those used by Sarus Cranes in

northern Australia. It is also true that lower sea levels would have narrowed the distance

between the Philippines, Borneo/Peninsula Malaya (via Palawan) and Indochina (Voris,

2000) and hence potentially also brought A. a luzonica and A. a. sharpii into closer

proximity. However, especially given the shifting of the courses of the Mekong (Attwood and

Johnston, 2001) it is likely that facing coastal regions of both Indochina and the Philippines

would have been mainly comprised of closed forest, making sub-specific contact more

difficult.

3.6 Conclusions and Recommendations

We have shown that A. a. gillae differs significantly from the A. a. antigone and A. a. sharpii

genetic cline described by others. Where once A. a gillae might have been considered part of

this cline, more detailed analysis has revealed greater structure. This has relevance to the

wider debate about subspecies, suggesting that the level of genetic analysis required before

subspecies are dismissed needs to be carefully considered, and wherever feasible triangulated

with information gleaned from other character traits.

That the single sample from A. a luzonica clustered with A. a. gillae hints at the potential for

a close evolutionary relationship. Should reintroduction of Sarus Cranes to the Philippines be

deemed desirable and viable, subject to further research on the genetic affinities of A. a.

luzonica, Australia might be an appropriate source of birds.

Page 69: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

69

Whilst Hachisuka (1941) found that Philippine birds were significantly smaller than those on

the south-Asian mainland, the general case for insular dwarfism is equivocal (van der Geer et

al., 2010). As we had access to only one individual of A. a. luzonica, further genetic work on

samples from Philippine museum specimens could help to clarify the status of this subspecies

and its potential to shed further light on the phylogeography of Sarus Cranes.

Page 70: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

70

CHAPTER 4 – THE EXTENT AND TRAJECTORY OF

INTROGRESSION

The Sarolga: conservation implications of genetic and visual

evidence for hybridization between the Brolga Antigone

rubicunda and the Australian Sarus Crane Antigone antigone

gillae. [Oryx, doi:10.1017/S003060531800073X].

4.1 Abstract

In order to investigate the extent of suspected hybridisation between the Brolga, Antigone

rubicunda, and Australian Sarus Crane, A. antigone gillae, first noted in the 1970s, the

genetic diversity of 389 feathers collected from breeding and flocking areas in north

Queensland, Australia was analysed. These were compared with 15 samples from birds of

known identity or which were phenotypically typical. Bayesian clustering based on ten

microsatellite loci identified nine admixed birds (Q = 0.1-0.9) confirming that Australian

cranes do hybridize in the wild. Four of these were backcrosses, also confirming that wild

Australian crane hybrids are fertile. Genetic analyses identified ten times more hybrids than

our accompanying visual field observations.

The analyses also provided the first definitive evidence that both Brolgas and Sarus Cranes

migrate between the Gulf Plains, the principal breeding area for Sarus Cranes, and major non-

breeding locations on the Atherton Tablelands. It is suggested that genetic analysis of shed

feathers might potentially provide a cost-effective means to provide ongoing monitoring of

this migration.

Page 71: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

71

The first observations of hybrids coincided with significantly increased opportunities for

interaction when foraging on agricultural crops, which have developed significantly in the

Atherton Tablelands flocking area since the 1960s. As the Sarus Crane is declining in much

of its Asian range, challenges to the genetic integrity of Australian Sarus Crane populations

have international conservation significance.

4.2 Introduction

4.2.1 Background

Hybridisation between what would appear to be well differentiated species is a commonly

observed phenomenon and widely studied in the context of speciation and reproductive

isolation among species (Arnold, 1997; Pennisi, 2016). Hybridisation has a range of

evolutionary consequences depending on the fitness of the hybrids, which may range from

sterility and reduced fertility, to hybrid vigour or heterosis, where hybrid individuals perform

better and out-compete parental species (Burke and Arnold, 2001; Pennisi, 2016). As a

consequence, if species are rare and locally confined (Rhymer and Simberloff, 1996), genetic

mixing can lead to extinction of one or even both parental species (Todesco et al., 2016).

Eventually, hybridisation can lead to speciation where two parental species gives rise to a

new species (Dowling and Secor, 1997; Soltis and Soltis, 2009), highlighting the relevance of

studying hybridisation from the perspective of conservation biology. Allendorf et al. (2001)

noted that taxa which have arisen through natural hybridisation clearly warrant protection.

However, they also noted that increased anthropogenically-related hybridisation is causing

extinction of many taxa (species, subspecies and locally adapted populations) by both

replacement and genetic mixing; going on to argue that hybrid taxa resulting from

anthropogenic causes should therefore be protected only in exceptional circumstances. It is

not yet known if Brolga/Sarus Crane hybrids have arisen as a result of anthropogenic habitat

Page 72: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

72

change nor if agricultural and conservation policy has the potential to curtail this; but it is

clearly an important matter for careful consideration. The starting point for this is gaining an

appreciation of its scale. Depending on the interaction of parental alleles and their dominance

patterns as well as epistatic effects, hybrid individuals need not be phenotypically

intermediate between their parents (Rieseberg et al., 1999). Introgression of alleles from one

species into the other also need not be symmetrical, as demonstrated in many cases of

mitochondrial introgression (Toews and Brelsford, 2012). On the nuclear level, genomic data

indicate that hybridisation and introgression interacting with recombination and selection

result in much more complex pictures of genetic diversity (Payseur and Rieseberg, 2016).

Secondary contact of taxa which are well differentiated in allopatry may result in

hybridisation (Grant and Grant, 2016; Vijay et al., 2016), even among non-sister species

(Koch et al., 2017). Such a situation is encountered for the two Australian cranes Antigone

rubicunda, the Brolga (see Fig. 4.1) and A. antigone gillae, the Australian subspecies of the

Sarus Crane (see Fig 4.2).

Fig. 4.1 Typical Brolga with unfledged chick (Gulf Plains); showing small red skull-cap type comb,

dark wattle and grey legs; with scalloped plumage [Photograph: T. Nevard]

Page 73: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

73

Fig. 4.2 Typical Australian Sarus Cranes with first year juvenile (Atherton Tablelands), showing pink

legs, red comb extending down the neck, whitish crown and darker grey colour. [Photograph: T.

Nevard]

Krajewski et al. (2010) calculated that the Brolga, which is endemic to Australasia (Australia,

New Guinea), probably split from a common ancestor with the Sarus Crane between 3.3 and

4.9 million years ago (mya). Sarus Cranes, which have a range that includes South and South-

east Asia (see Fig 4.3 below), were only noticed in Australia in the 1960s (Gill, 1967), but

were thought to have been isolated from Asian populations for up to 37,500 years (Wood and

Krajewski 1996); whereas Nevard et al. (2019b) calculate that this isolation was probably

closer to 28,000 years.

Page 74: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

74

Fig. 4.3 Global distribution of Brolga Antigone rubicunda and Sarus Crane Antigone antigone (A),

field data collection sites in northern Queensland (B, C and D). Distribution data derived from

BirdLife International and NatureServe Bird Species Distribution Maps of the World (2014) and the

Australian Bird Guide (2017). Gulf Plains Interim Biological Regionalisation Area (IBRA) derived

from Department of Environment and Energy (2012) and state roads from Department of Transport

and Main Roads (2016).

Hybridisation between the Brolga and Sarus Crane was first noted in French aviculture in

1934 (Gray, 1958), when fertile hybrids between the Brolga and the Southeast Asian

subspecies of the Sarus Crane A. a. sharpii, were produced. In 1972, when observing flocking

behaviour of Brolgas and Sarus Cranes on the Atherton Tablelands in Queensland, Archibald

(1981) identified at least two and possibly six potential hybrids (for which he coined the

Page 75: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

75

name ‘Sarolga’) and speculated that, if hybridisation were to increase, introgression could be

significant in the conservation of cranes in Australia. Similar concerns led Jones et al. (2005)

to recommend monitoring of the genetic integrity of Australian Sarus Cranes. This research is

the first to confirm and quantify the extent of natural hybridisation between the Brolga and

Australian Sarus Crane.

4.2.2 The Brolga

In Australia, Brolgas occur across northern Australia in Queensland, the Northern Territory

and Western Australia, and extend as far south as Victoria, where they are threatened

(Marchant and Higgins, 1993; Fig. 4.3A). In Queensland, they have been recorded

throughout the state (Marchant and Higgins, 1993) but are now only widespread and

abundant in the north. While breeding occurs throughout their Queensland range (Marchant

and Higgins, 1993; Barrett et al., 2003), it appears to be focused on the grasslands and open

savanna woodland of the Gulf Plains, south-east of the Gulf of Carpentaria and in Cape York

Peninsula. Scattered breeding occurs elsewhere, with flocking (non-breeding) regions

including the Atherton Tablelands and south of Townsville on the north-east coast of

Queensland (Lavery and Blackman 1969).

The Brolga population in northern Australia is generally considered stable and was estimated

to comprise of 20,000-100,000 individuals (Meine and Archibald 1996). A total of 51,969

Brolgas were recorded by Kingsford et al. (2012); all but 135 of these being in northern

Australia. This number should be treated with caution, as the count would also have included

Sarus Cranes and as cranes occur away from wetlands, these would not have been well-

counted. The only longitudinal counts are from north-east Queensland (Atherton Tablelands

and Townsville regions); with an annual average of approximately 800 individuals recorded

Page 76: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

76

by community-based naturalists over 21 years between 1997 and 2017; with a peak of

approximately 2,000 in 2005. (E. Scambler, pers. comm.).

The dispersal patterns of Brolgas and movements across their range are poorly understood,

including Queensland (Marchant and Higgins, 1993). Brolgas undertake seasonal movements

between breeding and non-breeding (flocking) habitats in response to rain and flooding

(Marchant and Higgins, 1993). Flocks are formed during the dry season and pairs disperse to

breeding sites in the wet season. Although a body of work is beginning to build in the south

of its range, especially in Victoria (Harding, 2001; Herring, 2005, 2007), knowledge of the

behaviour and ecology of tropical Brolgas remain largely confined to earlier studies such as

those of Hughes and Blackman (1969). Brolgas can use brackish waters, having a unique salt-

gland, and when observed with sarus cranes near Normanton, Brolgas favoured deeper,

generally coastal marshes (Walkinshaw, 1973). However, they do also occur in open forest

and woodland, grassland and cultivated land (Lavery and Blackman, 1969).

4.2.3 The Australian Sarus Crane

While the Sarus Crane is primarily a south Asian species, it is extinct in the Philippines,

declining in much of its Asian range, particularly in Burma, Thailand and Indochina (Meine

and Archibald, 1996; Archibald et al., 2003) and is classed as Vulnerable by the IUCN. The

Australian Sarus Crane population is therefore of international conservation significance and

hybridisation with the Brolga could potentially be a threat to the genetic integrity of the

Australian subspecies.

Sarus Cranes in Australia are almost entirely confined to Queensland (Fig. 3A, with known

breeding areas along the Gulf Plains and on the Normanby River floodplains in Eastern Cape

York Peninsula (Marchant and Higgins, 1993; Barrett et al., 2003; Franklin, 2008). Most

breeding records are from the Gulf Plains and concentrated in the Gilbert and Norman River

Page 77: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

77

catchments, where cranes make use of both natural and man-made wetlands on cattle

properties (Barrett et al., 2003). The only known major dry season flocking area for Sarus

Cranes is the Atherton Tablelands and its immediate vicinity. Counts of dry season flocks on

the Atherton Tablelands provide the only current estimates of Sarus Crane recruitment rates

available (Marchant and Higgins, 1993, Grant, 2005). An unknown proportion of the Sarus

Crane population remains on the Gulf Plains and the adjacent inland area in the dry season

and therefore utilizes different landscapes and food types during this period, as there are no

significant cropped areas. Migration routes followed by Sarus Cranes to the Atherton

Tablelands from their breeding areas and their habitat usage along the way are currently

unknown. The importance of migration routes is becoming increasingly understood, as

knowledge of breeding and non-breeding locations alone is insufficient to secure the

protection of migratory species (Runge et al. 2014).

Estimates of total numbers of Australian Sarus Cranes have varied from 5,000-10,000

individuals (E. Scambler, pers. comm.), but these have low reliability (Garnett and Crowley,

2000). Annual (unpublished) counts undertaken by local Birdlife Australia members on the

Atherton Tablelands (E. Scambler, pers. comm.) range between 800 and 1,500 individuals

flocking in dry seasons from 1997-2017 but it is currently not known what proportion of the

total Australian Sarus Crane population these birds represent. Estimates of current population

trends in Australia do not yet exist (E. Scambler, pers. comm.), but it is thought that Sarus

Crane numbers have risen from perhaps as few as 200 on the Atherton Tablelands in the late

1960s (G. Archibald, unpubl. obs.).

4.2.4 Aims

Although the presence of putative hybrid cranes on the Atherton Tablelands has been noted

by experienced observers for many years, it was hypothesised that it was possible that these

Page 78: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

78

were non-hybrid birds; exhibiting intrinsic phenotypic variability. Testing the degree of

introgression (if any) in the crane population, as well as providing a benchmark for its future

measurement both visually and genetically was therefore important. To achieve this, it

needed to be understood whether birds with the visual characteristics that had drawn

observers to believe that they were hybrids (Sarolgas) supported this assumption.

Against this background, we aimed to: (i) establish the extent of hybridisation between

Australian Sarus Cranes and Brolgas using molecular methods; (ii) seek evidence from these

genetic data for the presumed connectivity of breeding and flocking areas in the Gulf Plains

and Atherton Tablelands; (iii) describe the distribution of individuals with hybrid

characteristics on the Atherton Tablelands; and (iv) discuss the results, primarily with respect

to their relevance to biological conservation of the two species in Australia.

4.3 Methods

4.3.1 Field observations and sample collection

Over a three year period from July 2013 to June 2016, observations of cranes were made on

fifty-one 334.7 km circuits that encompassed 46 foraging sites distributed across all known

crane flocking areas of the Atherton Tablelands (Fig. 4.3D and Fig. 4.5) and readily visible

from the road. At each site counts were made of adult and first year juvenile Brolgas and

Sarus Cranes, using a pair of 10x40 binoculars; with potential hybrids (see Fig. 4.4) visually

examined with a 50x magnification spotting scope. As individuals could not be distinguished,

birds were probably counted on multiple occasions and at different sites. However, enough

cranes were counted that there is no reason to suspect systematic bias in the estimate of the

proportion of visually distinguishable hybrids amongst the wider crane population.

Page 79: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

79

Fig. 4.4 Heads and necks of typical species phenotypes and presumed hybrids. A: typical Brolga; B:

Sarolga (this bird was congruent with a typical Brolga, except that it had pink legs, slightly more red

on the nape of the comb than is typical of Brolgas, and a very small, almost non-existent wattle. This

combination of characters had not hitherto been recorded); C: typical Sarolga, with much shorter neck

comb and slight wattle; D: typical Australian Sarus Crane. [Photos A, C and D: T. Nevard; Photo B:

Brian Johnson].

The initial identification of hybrids was based on the presence of at least four readily-

observable characteristics (including at least two of (i), (iv) and (vi)) in Archibald’s (1981)

typology of Sarolgas: (i) larger and heavier than either parent species; (ii) an irregular border

to the comb; (iii) notches in the cap; (iv) intermediate or mixed-colour legs, (iv) a scalloped

mantle; (v) brighter yellow eyes than typical Sarus Cranes (which are more orange); and (vi)

a small wattle. Although Archibald’s (1981) Sarolga typology was helpful, in overall

appearance the Sarolgas he describes are more like atypical Sarus Cranes than atypical

Brolgas, whereas putative hybrids looking more like atypical Brolgas were also observed

(Fig. 4.4B).

Page 80: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

80

In May 2014, freshly-shed feathers were collected opportunistically from Miranda Downs

cattle station (17°19.755’S/141°52.770’E, elevation 55 m) on the Gulf Plains and in

September 2015 from Innot Hot Springs (17°43.749’S/145°16.302’E, elevation 628 m) on

the southern edge of the Atherton Tablelands (Fig. 4.3C). Only feathers with minimal

contamination from soil and vegetation were selected and picked-up from dry ground with

forceps, then immediately placed in sealed plastic bags and refrigerated prior to analysis.

4.3.2 Molecular analysis

A total of 389 feathers were collected and prepared for genotyping. This included 187 from

the Gulf Plains, 202 from the Atherton Tablelands, and 15 samples (blood, feathers) from

reference birds, which were either phenotypically typical or for which a potential history of

contact with the other species could be excluded or minimised (3 A. a. antigone, 7 A. a.

gillae, 5 A. rubicunda (see Table 4.1); note that including the non-Australian A. a. antigone in

the reference collection makes no difference to the veracity of the comparison with Brolgas).

Table 4.1 Reference birds. Samples were either taken from living and recently dead wild birds which

were phenotypically typical or from birds whose lineages of descent were known not to have had

contact with the other species. N: number of specimens.

Species N Locality Notes

A. a. antigone 3

Lemgo Avicultural

Collection, Germany

Breeding records show no interbreeding

with other Sarus Crane subspecies since

their ancestors were collected from the wild

in the 1970s.

A. a. gillae 5 Lemgo Avicultural

Collection, Germany

Breeding records show no interbreeding

since their ancestors were collected from the

wild in the 1970s.

A. a. gillae 2 Kairi, Queensland

Wild birds; phenotypically typical Sarus

crane.

A. rubicunda 3 Cairns Tropical Zoo,

Queensland

Breeding records show no interbreeding

since collected from a wild population with

no history of contact with Sarus Cranes.

A. rubicunda 2 Herbert River, Atherton

Tablelands, Queensland Wild birds; phenotypically typical Brolga.

Page 81: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

81

DNA was extracted using the slightly modified SDS/salting-out protocol of Miller et al. (in

press [accepted for publication in Avian Biology Research 21/08/2018]). In order to increase

the yield, dithiothreitol and Roti®-PinkDNA (Carl Roth®, Karlsruhe, Germany) were added.

Initially, the 22 microsatellite loci were tested which had been successfully used in other

species of crane (Hasegawa et al., 2000; Meares et al., 2009) as well as the Sarus Crane

(Jones, 2005) and the Brolga (Miller, 2016). Of these, ten could be consistently amplified and

scored (Table 4.2). PCRs were conducted in a volume of 10 µl and contained 1 µl DNA (10-

25 ng), 1 µl of 10 x NH4-based Reaction Buffer, 1.5-2.25 mM MgCl2 Solution (see Table 2

below), 0.25 mM of each primer, 0.2 mM of dNTP, 0.04 µl of BioTaq DNA Polymerase (5

U/µl), 0.6 µl of 1 % BSA and sterile ddH2O. For two loci, GR22 and GR25, we used the

MyTaqTM

mix (all products from BIOLINETM

, London, UK).

Table 4.2 PCR specifications and diversity of microsatellite loci. Dye, fluorophore at 5’-end

of forward primer; MgCl2, concentration of magnesium chloride (mM); N alleles, number of

alleles; T, annealing temperature (°C); gene diversity was estimated according to Nei (1987)

and allelic richness following Petit et al. (1998). Loci that could not be consistently amplified

or scored: Gamµ2, Gamµ5, Gamµ6, Gamµ7, Gamµ9, Gamµ12, Gamµ21, Gamµ102, GjM11,

GjM34, GjM40, GR17 (Hasegawa et al. 2000; Meares et al. 2009; Miller, 2016).

N alleles Brolga Sarus Brolga Sarus

Locus/dye

MgCl

2 T Brolga Sarus Gulf Table Gulf Table Gulf Table Gulf Table

Gamµ3/FAM 1.5 59 7 6 0.618 0.654 0.162 0.229 6.994 6.965 3.183 2.875

Gamµ18/HEX 1.5 53 3d 1 0.224 0.363 0.000 0.000 2.943 3.000 1.000 1.000

Gamµ24/HEX 2.25 60 5 5d 0.578 0.628 0.204 0.338 4.943 4.998 3.732 4.000

Gamµ101b/HEX 2 59 4 3 0.173 0.203 0.349 0.326 3.000 2.566 2.273 2.999

GjM8/FAM 2 59 3 3 0.159 0.265 0.157 0.125 2.996 2.910 2.254 2.000

GjM13/HEX 2 60 5d 3

d 0.722 0.557 0.498 0.042 5.000 3.815 2.276 1.875

GjM15/FAM 2 59 9 5 0.794 0.757 0.552 0.539 8.994 8.251 4.422 4.000

GjM48b/HEX 2 56 8d 6

d 0.559 0.571 0.160 0.362 6.999 7.853 3.039 3.862

GR22m/HEX na 60 11

d 5 0.625 0.566 0.448 0.434 7.937 8.430 2.575 2.875

GR25m/Cyanine3 na 60 5 3 0.727 0.741 0.512 0.442 5.000 5.000 2.996 2.000

d heterocygote deficiency;

m multiplexed

Page 82: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

82

The PCR profile comprised an initial denaturation at 94 °C, 36 cycles including denaturation

at 94 °C, primer specific annealing (Table 4.2) and extension at 72 °C each for 30 s, and a

final elongation at 72 °C for 10 min. Microsatellite alleles were separated on a 3130xl

Genetic Analyzer, together with the GeneScanTM

600 LIZ® Size Standard 2.0. Fragment

sizes were determined manually in order to avoid the inconsistencies in automatic calling due

to the arbitrariness of bin width definitions using GeneMapper® 4.0 (all three products from

Applied Biosystems®, Waltham, USA). In order to control accuracy of size determination,

PCRs of samples were repeated which initially gave weak signals or had rare variants. In

other cases, problematic PCR samples that were initially typed in different runs, so possibly

difficult to calibrate, were loaded on the same plate to improve comparability.

As there was the possibility that feathers of a particular bird were collected more than once,

identical genotypes were identified using Cervus 3.0.7 (Kalinowski et al., 2007).

Consequently, 41 samples (16 Gulf Plains, 24 Atherton Tablelands, 1 captive reference A. a.

gillae that was identical to its sibling) were excluded from the following analyses. In eleven

of these cases, the same genotype was found in both localities. In these, the sample from the

Atherton Tablelands was arbitrarily excluded. As only 16.2% of the birds had missing data

and only two of the nine potential hybrids lacked data (at one locus each) this was not

considered problematic. Given the moderate number of markers, it was not possible to

distinguish siblings from non-siblings because genetic variance would be too large.

In order to identify the number of genetic clusters and potential hybrid individuals, Bayesian

clustering was conducted, using STRUCTURE 2.3.4 (Pritchard et al., 2000; Falush et al.,

2007) with K (number of clusters) ranging from one to five and ten replicates assuming the

admixture model and, in different runs, either uncorrelated or correlated allele frequencies.

The Markov chains ran for 550,000 generations including a burn in of 50,000. Pilot runs with

Page 83: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

83

longer Markov chains yielded identical results. Structure Harvester 0.6.94 (Earl and von

Holdt, 2012) was used to analyse the data following Evanno et al. (2005) and Structure Plot

V2.0 (Ramasamy et al., 2014) and Adobe Illustrator CS4 (Adobe Systems Inc., San José,

California) to render the graphical output. Individuals were defined with admixture Q

between 0.1 and 0.9 as hybrids following Devitt et al. (2011). These thresholds appear

considerably more conservative in the analysis, as ten loci were used and not only three as by

Devitt et al. (2011). However, this restriction was necessary, as the loci they used had species

specific alleles whereas the two species of cranes compared shared some of their alleles.

There was no attempt made to identify different classes of hybrids (F1, F2 and their

backcrosses) because this would have required genotyping at least 50 loci (Fitzpatrick, 2012),

whereas only 10 were available. The Bayesian clustering approach was validated by K-means

clustering, which, in contrast to STRUCTURE, is free of population genetic assumptions,

using GenoDive 2.0b23 (Meirmans and van Tienderen, 2004). Individuals were clustered

based on their allele frequencies according to the pseudo-F-statistic of Calinski and Harabasz

(1974; see Meirmans, 2012).

More detailed population genetic analyses were conducted based on the clusters identified by

Bayesian clustering including descriptive statistics, tests for Hardy-Weinberg-Equilibrium,

linkage disequilibrium, null-alleles and population differentiation, using FSTAT 2.9.3.2

(Goudet, 1995), GenePop 4.2 (Raymond and Rousset, 1995; Rousset, 2008), and Micro-

Checker 2.2 (van Oosterhout et al., 2004). Gene flow was estimated using the formula Nm =

[(1/FST)-1]/4 (Nm, number of effective migrants) as well as based on private alleles, i.e.

alleles occurring only in one subpopulation (Barton and Slatkin, 1986), the latter in GenePop.

Probabilities that two individuals were identical were calculated with Cervus, assuming either

Page 84: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

84

that the two birds were unrelated or, more conservatively, that they were full siblings (Waits

et al., 2001). In multiple tests, α was Bonferroni corrected.

4.3.3 Lincoln Petersen Index

The Lincoln-Petersen Index was applied to ‘genetically re-trapped’ birds (nGulf x

nTablelands/nboth), based on methodology described by Southwood and Henderson (2000).

4.4 Results

4.4.1 Field observations

Crane observations on the Atherton Tablelands are summarized in Table 4.3 and Fig 4.5.

Table 4.3 The number (and percentage) of cranes observed across 46 sites on the Atherton

Tablelands 2013-2016.

Life history

stage Brolga

Sarus

Crane Sarolga Unidentified¹

Total

cranes

Adults 24,870 (55.8) 12,245 (27.5) 84 (0.2) - 37,199 (83.5)

First year

juveniles 2,491 (5.6) 1,136 (2.5) 0 (0) - 3,627 (8.1)

Total 27,361 (61.4) 13,381 (30) 84 (0.2) 3,728 (8.4) 44,554 (100)

¹Due to poor weather/light conditions

Page 85: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

85

Fig. 4.5 Pie-charts showing general distribution and clustering of Brolga (blue), Sarus Crane (red) and

Sarolga (open yellow dots) visually identified at 46 observation sites on the Atherton Tablelands

2013-2016. Grey quadrants represent unidentifiable cranes due to poor light and weather conditions.

Out of 40,826 cranes identifiable to species level (out of a total of 44,554 observations), only

84 were classified as morphologically intermediate Sarolgas (0.205 %). Variation among

Page 86: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

86

these intermediate birds meant it was not possible to assess the extent of potential

hybridisation (see Fig. 4.4).

Observations of visually identifiable hybrids on the Atherton Tablelands were concentrated in

a relatively small number of areas (see Fig. 4.5), such as Innot Hot Springs, Tumoulin/Kaban,

Hastie’s Swamp and the Mareeba Wetlands. Figure 4.5 also demonstrates that although they

occur in mixed flocks, Brolgas and Sarus Cranes are differentially distributed across the

Atherton Tablelands; with higher Sarus Crane numbers focused on more fertile volcanic soils

closer to Atherton itself and both Brolgas and putative hybrids predominantly visually

identified on areas of lower intrinsic fertility.

At sites with higher Brolga or Sarus Crane counts, the species with the lesser numbers in the

flock tended to have proportionally more first year juveniles (Table 4.4; Chi2 tests: p <

0.0001 in both cases), potentially increasing opportunities for interactions between species,

whilst juveniles are at a critical age for pair-bonding.

Table 4.4 Proportion of juveniles at sites dominated by either Brolgas or Sarus Cranes.

Brolga Sarus Crane

Total no. % Juvenile Total no. % Juvenile

Brolga-dominated sites 25,510 8.8 2,273 14.5

Sarus-dominated sites 1,851 13.3 11,308 8.9

Total 27,361 9.1 13,581 9.8

4.4.2 Population genetics

In all STRUCTURE analyses, the highest ΔK resulted assuming two clusters and this was

confirmed by k-means clustering. These clusters corresponded well to the species Brolga and

Sarus crane, as indicated by the reference birds (Fig. 4.6) As allele frequencies apparently

differed considerably between the species, we based the subsequent analyses on the

Page 87: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

87

STRUCTURE model assuming uncorrelated frequencies. While the majority of birds were

either pure Brolga or pure Sarus Crane, some were admixed (Fig. 4.6).

Fig. 4.6 STRUCTURE plot with K = 2 showing admixture among a total of 363 Brolga (blue) and

Sarus Cranes (red) ordered along x-axis by decreasing Q (membership coefficient; y-axis) with Brolga

as reference. Black rectangle includes nine hybrids with Q ranging between 0.9 and 0.1. Bar

underneath indicates origin of birds and reference specimens (green, Atherton Tablelands; orange,

Gulf Plains; blue, A. rubicunda; magenta, A. a. antigone; red, A. a. gillae)

According to the criteria used, 239 samples belonged to Brolgas (90 Gulf Plains, 149

Atherton Tablelands) and 101 to Sarus Cranes (77 Gulf Plains, 24 Atherton Tablelands). Nine

individuals (2.58 %; four Gulf Plains, five Atherton Tablelands) had Q scores ranging from

0.125 to 0.888 (reference: Brolga) and were identified as hybrids (Fig. 6), a proportion about

ten times greater than in field observations. Four hybrids had alleles occurring only in

Brolgas at one or two loci indicating that hybrids are fertile and can interbreed at least with

Brolgas or other hybrids.

After exclusion of reference birds and hybrids, the species were analysed separately,

subdividing both into two subpopulations corresponding to the localities. Genetic diversity is

summarized in Table 4.2. Gene diversity and allelic richness were both considerably higher

in Brolgas. Brolgas also exhibited 26 private alleles, between one and six per locus, whereas

Sarus Cranes had only eight private alleles and four loci did not have any. This translates into

much lower probabilities in Brolgas that two individuals will carry the same genotype than in

Sarus Cranes. The probabilities to find two unrelated individuals with identical genotype

Page 88: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

88

were 7.1x10-7

for Brolgas and 0.0012 for Sarus Cranes. Assuming the individuals were full

siblings, these probabilities rose to 0.0023 and 0.0395, respectively. In both Brolgas and

Sarus Cranes respectively, four and three loci deviated from the Hardy-Weinberg-

Equilibrium and the same loci had indications of null-alleles. However, this was probably due

to a fairly high number of rare alleles (frequency < 5%) in both species and was therefore not

considered problematic for the analyses. In any case, the general results of the STRUCTURE

analysis did not change after exclusion of these loci. Linkage disequilibrium did not play a

role in either species. In both species, subpopulations were hardly differentiated with FST =

0.010 (confidence interval: 0.003 - 0.020) in Brolgas and 0.047 (-0.006 - 0.136) in Sarus

Cranes. Sample size limited the significance of the result for Brolgas and is probably not

biologically relevant. Estimates of migration based on the private allele method were high in

both species with 6.56 and 7.75 migrants per generation for Brolgas and Sarus Cranes,

respectively. Transforming FST into estimates of migration using the formula Nm = [(1/FST)-

1]/4, 24.75 and 5.07 migrants respectively were inferred.

4.4.3 Lincoln-Petersen Index

In eleven cases, the same genotypes were collected at both Miranda Downs in the Gulf Plains

and Innot Hot Springs in the southern Atherton Tablelands. Of these, nine were attributable to

Brolgas, and two to Sarus Cranes. According to the probabilities of encountering two

individuals carrying identical genotypes by chance with the sample sizes, only the two Sarus

Cranes might be cases of chance identity, provided the respective birds were siblings (see

above). Application of the Lincoln-Petersen Index (nGulf x nTablelands/nboth; Southwood and

Henderson, 2000) to these results implies that at least 1473 Brolgas and 923 Sarus Cranes

could have moved between the Gulf and Atherton Tablelands.

Page 89: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

89

4.5 Discussion and Conclusions

Since Archibald’s first observation of hybrids in 1972 (Archibald, 1981), local observers

have continued to note the presence of apparent hybrids amongst dry season, non-breeding

crane flocks on the Atherton Tablelands (Matthiessen, 2002), as well as elsewhere. This

research has shown that natural hybridisation between Brolgas and Sarus Cranes is happening

and occurs more widely than is visually detectable. As hybridisation is still relatively

uncommon and Brolgas and Sarus cranes are differentially distributed on the Atherton

Tablelands (Fig. 4.5), its evolutionary influence could potentially be relatively limited.

However, a closer look at the allele-combinations in the hybrids suggests that introgression

could become more widespread. Four of the birds had alleles that occur only in Brolgas at

one or two loci. While one of these birds had a membership index of 0.5, suggestive of being

an F1-hybrid, these pure Brolga combinations indicate that the birds are hybrids of a later

generation with a hybrid parent that back-crossed with a Brolga or another hybrid. Sarolgas

therefore appear to be fertile at several generational re-combinations.

Longitudinal research on Darwin’s Finches Geospiza spp. has shown that they evolved from

a common ancestor in the Galápagos archipelago in the last two million years (Grant and

Grant, 2008) with both natural selection and introgressive hybridisation key processes in their

evolution. At this stage it is not possible to predict the trajectory of Brolga/Sarus Crane

introgression but Lamichhaney et al. (2016) have demonstrated that introgression at only a

single locus may have important evolutionary consequences. Interaction of introgression and

selection may be complex and locality dependent, as shown for European Crows Corvus

corone (Vijay et al., 2016), where genetic and phenotypic differentiation varies significantly

across its range; with some boundaries firm and others mobile. Ongoing Brolga/Sarus Crane

introgression may provide alleles to either species that could allow them to adapt to

Page 90: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

90

environmental changes brought about by such stresses as climate change or agricultural

intensification. Whether the larger body size of hybrids identified by Archibald (1981)

confers a selective advantage over the parental species, potentially leading to their

displacement, cannot yet be determined.

The current Australian population of Sarus Cranes appears to have increased relatively

rapidly from what is thought to have been small numbers in northern Australia following its

formal identification in the 1960s (Gill, 1967). Archibald (1981) estimated the Sarus Crane

population was approximately 200 on the Atherton Tablelands in 1972, when he identified at

least two and possibly six potential hybrids between Brolgas and Sarus Cranes (Archibald,

1981). Crops have been grown in northern Australia since the late 19th Century and

expanded substantially on the Atherton Tablelands in the 1960s, soon after Gill’s first Sarus

Crane observation. Templeton et al. (1990) have identified that changes in landscapes can be

associated with genetic consequences, including introgression that threatens species integrity

(e.g. Kakariki Cyanorhamphus spp (Triggs and Daugherty, 1996), Black-eared Miners

Manorina malanotis Clarke et al., 2001). Landscape features and heterogeneity can also be

reflected in population genetic structure within and across species (Miller and Haig, 2010;

Safner et al., 2011) and often changes in landscape are associated with habitat loss,

contributing to population decline and possibly resulting in changes in the genetic diversity of

remaining populations (Keyghobadi, 2007; Templeton et al., 1990).

That the species with the lesser numbers in a mixed-species flock has proportionally more

first year juveniles means that young Brolgas and Sarus Cranes are more exposed to one

another other at a time when imprinting is important (Horwich, 1996). This may make pairing

with the other species more likely. The enduring pair bond and longevity in both species

(Johnsgard, 1983) suggests that, once formed, hybrid pairs could contribute numerous hybrid

Page 91: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

91

offspring to the population. Although we cannot yet predict the trajectory of crane

introgression in northern Australia, there are clear conservation implications. The Sarus

Crane is thought to be declining in much of its Asian range, particularly in Burma, Thailand

and Indochina (Meine and Archibald, 1996; Archibald et al., 2003). The maintenance of

Australian Sarus Crane populations and their genetic integrity may therefore provide security

for the species should populations continue to decline in Asia. However, there would appear

to be no practical way of eliminating Australian crane introgression in the wild, as: (i) hybrids

are not necessarily visually distinguishable in the field; (ii) Australian cranes occur in remote

and inaccessible areas; and (iii) the underlying behavioural stimuli are not yet understood. If

related to anthropogenically-driven land use change, Brolgas and Sarus Cranes may be

brought into ever-closer and more frequent foraging and flocking proximity, potentially

leading to more hybridisation opportunities. Therefore, should the maintenance of

genetically pure (screened) populations of Australian Sarus Cranes and northern Australian

Brolgas be considered desirable for conservation purposes, one or more well-managed

extralimital populations may need to be established. As far as is known, the only captive

flock of Australian Sarus Cranes is at Lemgo in Germany, whose ancestry dates from a wild

capture in the 1970s. This could potentially form a nucleus for a captive Australian Sarus

Crane captive metapopulation, spread across several institutions. Establishing a flock

managed to maintain its genetic variability at this stage could also serve as insurance against

climate change impacts should these eventuate – with some models predicting a potential

91% decline in climatic suitability for the Australian Sarus Crane by 2085 (Garnett et al.,

2013).

As feathers have been collected from what appear to be the same birds in both the Gulf Plains

and Atherton Tablelands, considering the low probabilities of encountering genetically

Page 92: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

92

identical individuals, this has confirmed what has hitherto only been assumed by

birdwatchers in North Queensland – that cranes migrate between breeding areas in the Gulf

Plains and the Atherton Tablelands, a distance of about 500 km. As one would expect, there

appears to be no material genetic difference between Gulf Plains and Atherton Tablelands

cranes, indicated by estimates of population differentiation and migration after exclusion of

the duplicate genotypes. The conservation implications of this are significant, as both areas

will need to be targeted with appropriate conservation measures to ensure the survival of

north Queensland’s Brolgas and Sarus cranes. There are currently no laws protecting species

that migrate within Australia (Runge et al., 2017).

The identification of identical birds in both the Gulf Plains and Atherton Tablelands using

shed feathers highlights the potential of genetic analyses to explore migration patterns and

potential trends, not only in Australian cranes but also in other species. Although collection

of feathers was opportunistic rather than deliberately random and estimates using the

Lincoln-Petersen Index therefore necessarily indicative, they are of the same order as other

estimates of crane populations on the Atherton Tablelands (Brolgas 800-2000 counted/1473

calculated; Sarus Cranes 800-1500 counted/973 calculated). Similarly to the use of non-

invasive genetic monitoring techniques elsewhere, such as for large carnivores in Europe

(Kaczensky et al., 2012), genetic analysis of feathers could contribute to the assessment of

population trends and movements of cranes within the Atherton Tablelands and between the

Atherton Tablelands and other locations where either species is found.

Although feathers have the advantage of being a non-invasive means of gathering genetic

data on individuals, as noted by Hou et al. (2018); the probability of detecting hybrids was

over 10 times greater using genetic analysis. However, the boundaries between pure and

hybridised birds are unclear and are probably indefinable, given that hybrids are fertile;

Page 93: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

93

meaning that introgression may have reached the point where it has no clear thresholds. A

subject of future research could therefore be the definition of the relationship between

genetics and phenological expression, using the feathers of known birds with hybrid

phenotypes; using data from more than the 10 loci available (Fitzpatrick, 2012). In addition,

genomic approaches targeting coding loci (Payseur and Rieseberg, 2016) would provide a

deeper insight into the evolutionary consequences of the hybridisation, as the example of

Darwin’s Finches has shown (Lamichhaney et al., 2016). However, to obtain DNA in the

necessary quantities and quality would require catching many more cranes to obtain tissue

and blood samples.

Whilst acknowledging that the proportion of apparent hybrids identifiable by observers is

likely to be an order of magnitude lower than are actually present in the crane population,

both visual observation and analysis of shed feathers could provide valuable data to help

monitor the trajectory of Australian crane introgression.

Page 94: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

94

CHAPTER 5 – RESOURCE PARTITIONING AND

COMPETITION

Agriculture, Brolgas and Australian Sarus Cranes on the

Atherton Tablelands.

[Accepted for publication in Pacific Conservation Biology 24/12/2018]

5.1 Abstract

Flocks of Brolgas Antigone rubicunda and Australian Sarus Cranes A. antigone gillae

congregate in cropping areas of the Atherton Tablelands in north Queensland, Australia,

during the non-breeding months of May to December each year and sometimes come into

conflict with farmers. The central part of the region has been declared a Key Biodiversity

Area, largely because it is the only well-known non-breeding area for the Australian Sarus

Crane. The spatial and temporal patterns of use of this landscape were investigated for

foraging by the two species to determine how they might be affected by changes in cropping.

Abundances of the species were positively correlated with each other over both time and

space. Sarus Cranes were nevertheless markedly more abundant on the fertile volcanic soils

of the central Tablelands, whilst Brolgas were more abundant on a variety of soils in outlying

cropping areas close to roost sites, especially in the south-west of the region. Both species

used a wide variety of crops and pastures but occurred at highest densities on ploughed land

and areas from which crops (esp. maize) had been harvested. In addition, Brolgas were also

strongly associated with early-stage winter cereals with volunteer peanuts from the previous

crop. We conclude that maize and peanut crops are important as foraging sites for both

species during the non-breeding season, a situation that requires management in the interest

Page 95: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

95

of both cranes and farmers, especially as cropping patterns intensify and agricultural

technology changes. However, the mobility and adaptability to agricultural change of most

crane species suggests that Brolgas and Sarus Cranes are also likely to be able to take

advantage of newly created cropping areas.

5.2 Introduction

Most crane species make use of agricultural land for foraging (Harris and Mirande, 2013).

As global food production has expanded to meet rising demand from a growing human

population and increasing personal wealth, some crane populations have benefitted

significantly from the expansion of mechanical harvesting, by foraging on arable land,

especially on grain spilt or remaining after harvest (Meine and Archibald, 1996). However,

emerging technological and agronomic factors are likely to have an impact on some of the

crane species that have adapted to contemporary agricultural practices, potentially leading to

changes in their population, distribution, migration patterns, flocking and breeding success.

Changes in agricultural practice, such as reduction in cropping intervals and improved

harvesting efficiency can lead to dramatic reductions in post-harvest waste grain (Anteau et

al. 2011). In some sites cranes have been forced to seek new food sources at greater distances

from roost sites, resulting in lower reproductive success and increased crop damage (Pearse et

al., 2010, Austin, 2012, Barzen and Ballinger, 2017). Birdlife International (2018a) have

identified that agriculture is the biggest threat to globally threatened birds, with cranes having

a Red List Index (RLI) of 0.71 in 2016; significantly below the 0.91 RLI for threatened birds

as a whole. How cranes adapt to agricultural change is therefore a key conservation issue and

a management consideration for farmers and policy makers in crane range areas (Harris and

Mirande 2013).

Page 96: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

96

Both Australian crane species interact with agriculture. The Brolga Antigone rubicunda has

disappeared from much of its historical range in southern Australia as a result of agricultural

development and persecution (Arnol et al., 1984) but remains relatively common in

Australia’s tropics (Kingsford et al., 2012) where it has come into conflict with the

agricultural sector in north Queensland since at least the 1930s (Townsville Daily Bulletin,

1935). The Australian Sarus Crane A. antigone gillae breeds primarily on remote pastoral

areas in north Queensland but, along with the Brolga, migrates 500 km to spend the dry, non-

breeding season on the Atherton Tablelands (Archibald and Swengel, 1987; Nevard et al.

(2019b), one of Australia’s most intensively farmed areas. Because the Atherton Tablelands

are so important for the Sarus Crane, which is listed as Vulnerable by the IUCN (Birdlife

International, 2018b), they were a key factor in designation of the Atherton Tablelands Key

Biodiversity Area (Birdlife International, 2018c).

Ecologically Brolgas have traditionally been characterised as a species of freshwater and

brackish wetlands; widely dispersed when breeding but gathering in flocks during the dry

season to feed on sedge tubers in seasonally dry swamps (Herring, 2005; Chatto, 2006;

Herring, 2007; Franklin, 2008). In north Queensland they concentrate, to varying degrees, on

the Atherton Tablelands and east coast marshes (such as near Cromarty, Giru, Townsville and

Ingham), with their unique salt-gland (Hughes and Blackman, 1973) allowing them to exploit

brackish waters. Australian Sarus Cranes tend to avoid coastal wetlands, appearing to prefer

well-drained rather than wetland foraging areas (Lavery and Blackman, 1969). Where the

breeding ranges overlap, Brolgas favoured deeper, generally more coastal, marshes than

Sarus Cranes (Walkinshaw, 1973). However, both species also occur in open forest and

woodland, grassland and cultivation (Lavery and Blackman, 1969).

Page 97: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

97

Of 51,969 Brolgas recorded nationally by Kingsford et al. (2012), all but 135 were in

northern Australia. Since 1997 between 800 to 1500 Sarus Cranes have been counted

annually on the Atherton Tablelands (E, Scambler, pers. comm.), a number that may have

risen from a comparatively low base in the late 1960s (G. Archibald, pers. comm.), possibly

because arable agricultural development and changing crop patterns increased the availability

of residual unharvested crops.

Based on the experience of other crane ranges, the future of cranes on the Atherton

Tablelands is likely to be influenced by two key elements: a) land use change - through

expansion of cultivation and perennial crops and intensification of cultivation and water

management (Nevard et al. 2019); and b) technological development - through increased

mechanization, automation and agronomic sophistication. Both have the potential to affect

the availability of forage for cranes, especially as they recover from migration to dry

season/wintering habitats and prepare for the following breeding season.

This research seeks to build a platform from which to develop appropriate crane conservation

measures under conditions of changing agronomic practice. The spatial and temporal patterns

of occurrence of cranes are reported at their feeding sites over three non-breeding seasons to

answer three main questions: (i) what are the key habitat parameters, including crop

associations, of crane species; (ii) how do crane species differ in their usage of the study area;

and (iii) what are the implications for crane conservation management?

5.3 Methods

5.3.1 Atherton Tablelands Study Area

The study area encompassed known crane foraging habitat on the Atherton Tablelands (see

Fig. 5.1 below) from north of Mareeba south to Innot Hot Springs. Elevation of the surveyed

Page 98: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

98

area ranged from 399 to 1,070 m above sea level and with a tropical mid-elevation to upland

climate and strongly seasonal rainfall, ranging from 900 to 2,200 mm. The underlying

geology is granitic and metamorphic in the north and west and volcanic in the centre and

south, with patches of alluvial loam associated with major river systems and peat in dormant

volcanic craters.

Fig. 5.1 Atherton Tablelands region, showing survey sites, irrigated agriculture (Queensland

Government 2016) and the Atherton Tablelands Key Biodiversity Area. Inset shows location of study

region and general distribution of Brolgas and Sarus Cranes in Australia.

As a result of this topographic, climatic and geological variation, as well as its tropical

location, the Atherton Tablelands supports one of Australia’s most diverse biotas (Nix and

Switzer, 1991). It also sustains one of Australia’s largest ranges of both irrigated and non-

irrigated crops (see Table 5.1), contributing gross revenue of over AU$552 million to the

regional economy. Tree crops, are the largest component of the agricultural sector by value,

followed by field crops, animal industries and horticulture (Department of Agriculture and

Page 99: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

99

Fisheries 2015, Nevard et al., 2018). There is extensive public and private sector irrigation

infrastructure and a relatively dense rural population, many small family farms and a high

diversity of cropping.

Table 5.1 Area and value of agricultural production on the Atherton Tablelands, Queensland

(Department of Agriculture and Fisheries (2015) and Nevard et al. (2018).

Agricultural activity Area (ha)

Gross revenue per hectare

per year (AU$ '000)

Livestock 558,872 0.2 Beef cattle 550,000 0.1

Dairy cattle 8,800 3.9

Pigs and poultry 72 550.4

Field Crops and Horticulture 23,427 4.6

Sugar cane 10,956 3.6

Maize 4,719 2.4

Hay 3,020 1.2

Grass seed 1,195 4.0

Potato 972 16.2

Legume seed 968 3.1

Peanut 874 5.5

Minor field crops 723 35.2

Tree crops and forestry 11,297 27.8

Forestry plantations 3,600 5.6

Minor tree crops 2,497 34.5

Mango 2,400 21.1

Banana 1,850 49.2

Avocado 950 87.3

Miscellaneous 151 577.2

Total 593,747

Water for the Mareeba-Dimbulah Irrigation Area within the Atherton Tablelands region is

supplied from Tinaroo Dam (which has important crane roosting sites along its shoreline)

through an extensive network of channels and streams. Farming enterprises in other parts of

the Tablelands (including intensive agricultural areas around Atherton, Malanda and

Ravenshoe) draw water from natural watercourses, bores and small dams, many of which

provide drinking and roosting opportunities for cranes.

Page 100: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

100

5.3.2 Field surveys

A 335 km driven traverse, with 46 readily visible crane feeding sites spread widely across the

Atherton Tablelands, was established and data recorded from 44 traverses between July 2013

and December 2015. Traverses were undertaken in the months of May to December

inclusive, encompassing the time cranes are present on the Atherton Tablelands. A mean of

43.0 sites were assessed per traverse, a total of 1,892 site surveys, some site surveys being

missed or curtailed due to weather and the need to vary the timing of site visits. At each site

survey, cranes were counted and identified as Sarus Crane, Brolga, a hybrid of the two

species (Nevard et al. 2019a) or unidentified crane, using 10x40 binoculars and a 50x

spotting scope; with land use recorded as one of 32 crop classes, including current state (e.g.

ploughed, growing, harvested). On a one-off basis for each site, the following parameters

were recorded: area at each site visible from the road and used by cranes, elevation, distance

from dwellings or other buildings with human activity (3 classes), surface geology

(Cainozoic volcanic, metamorphic or alluvial), land use class (cropping, grazing, wetland)

and irrigation status (irrigated or not). All known roosts, identified during annual counts

conducted by Birdlife North Queensland (E. Scambler, pers. comm.), were visited, cranes

counted and then classified as either Brolga or Sarus Crane-dominated. Classification was by

simple majority but most roosts were strongly skewed to one species.

5.3.3 Data analysis

Most cranes were identified to species level but those that were unidentified were attributed

to a species in proportion to the sum of identified cranes counted at that site. The very few

hybrids identified visually (<0.3% of cranes counted) were ignored.

Sites were characterised by slope, distance from the nearest known crane roost and mean

annual rainfall. Slope was identified in three classes (<2%, 2–9%, >9%). Distances from the

Page 101: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

101

nearest roost to each feeding site were calculated to 0.1 km using a Geographic Information

System (GIS). Mean annual rainfall was determined by interpolation in three classes (<1,000

mm, 1,000–1,200 mm, 1,200–1,600 mm) from the records of the Bureau of Meteorology

(www.bom.gov.au). Field observations of site usage were converted to area measurements in

a GIS, and the density of each crane species calculated for each site survey.

Correlations and MARSpline and regression tree analyses were performed in Statistica 13.2

(Dell Inc. 1984-2016). Complex models were evaluated with the Permanova module in the

Permanova+ add-on to Primer 6 (Anderson et al., 2008). Permanova+ is permutational

software, meaning that probabilities are calculated by permutation rather than based on the

normal distribution. This is particularly appropriate where there are frequent zeroes in the

dataset, as was the case in this study. Spatial analyses were undertaken with ArcGIS v10.4.1

(ESRI, 2016).

To evaluate whether crane species used the Atherton Tablelands differently over time,

Permanova was used to model crane numbers using crane species and traverse number as

fixed effects, site as a random effect included to reflect the repeated measures structure of the

dataset, and the species by traverse interaction. A significant interaction term would provide

evidence of different frequencies across time. A total of 29 complete traverses for 43 sites

were used in this analysis. The Euclidean distance measure was employed to untransformed

count totals, a Type III Sum of Squares model, and 9999 permutations. Additionally, counts

were summed for each traverse and examined the temporal correlation between Sarus Crane

and Brolga numbers (log (n+1)-transformed).

To determine whether crane species use different sites on the Atherton Tablelands, the rank

correlation between species was first examined (separately using each of abundance and

density) for the 46 sites averaged across all site surveys. To measure spatial association of

Page 102: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

102

sites Moran’s I statistic was calculated, based on the percentage of Brolgas at each site and

the Getis-Ord Gi* statistic used to identify clusters of sites with values higher (hotspot) or

lower (coldspot) in magnitude than expected by chance (Getis and Ord, 1992).

To consider the possibility that choice of feeding sites might be constrained by distance to

roosts, the distance between them was evaluated as a function of species. Distances to the

nearest roost were categorised into four classes with approximately equal number of foraging

sites (0-1.5 km, n=10; 1.5-4.0 km, n=10; 4.0-8.0 km, n=13; 8.0-18 km, n=13). Site usage was

quantified as the fourth-root of species-specific mean crane density. Effects of distance class

were then evaluated by species interaction in Permanova with site as a random effect, using

the Euclidean distance measure, a Type III Sum of Squares model, and 9999 permutations.

As the interaction term was statistically significant, post-hoc pairwise comparisons of

distances classes were conducted for each species separately.

To examine foraging habitat preferences of the species based on fixed attributes of feeding

sites (i.e. not including crop type, which changed over time at many sites), a regression tree

analysis was run for each species mean site density after first culling variables using

MARSplines (non-parametric multiple regression models) on mean site density, to avoid

serious overfitting. For the regression trees, only groups containing at least 10 sites were

split, and the final trees pruned to remove incoherent or uninterpretable lower branches.

To identify crop type preferences, each crane species was considered in a separate model,

using the 31 feeding sites at which cropping was a frequent occurrence. The 32 crop

classes/states were aggregated into ten with similar characteristics. The response variable was

the density of the crane species during each site survey, with aggregated crop type as the key

predictor and site and traverse as random effects to control for non-independence of

individual surveys in space and time. The analysis was performed in Permanova using the

Page 103: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

103

Euclidean distance measure, a Type III Sum of Squares model, and 9999 permutations. Given

significant effects of crop type, preferences were then evaluated by post-hoc pair-wise

comparisons.

To evaluate seasonal patterns of crop type availability, the 10 crop classes were further

aggregated into four based on similar characteristics: ploughed land, growing grains,

harvested grains and others. Observed frequencies were compared to expected frequencies

across months (May to December), the latter based on constant proportionality calculated

across all site surveys and evaluated using Chi-squared analysis.

5.4 Results

In total, 28,548 crane observations were used in analyses, a mean of 649 per traverse, 15.1

per site survey and 50.4 per site at the 569 (30.1%) site surveys at which one or more cranes

were present. These were 62.1% Brolgas, 37.7% Sarus Cranes and <0.3% hybrids. When

cranes were present, site surveys yielded from one to 880 cranes and densities from 0.008 to

23.7 ha-1

(mean 1.6 ha-1

). Both species were present in 241 site surveys, Brolgas alone in 159

site surveys, Sarus Cranes alone in 166 site surveys and hybrids alone (one individual on each

occasion) in three surveys. Cranes were present from 2.4% to 90.3% of the times that a

particular site was surveyed (mean 30.6%). Twelve roosts were dominated by Brolgas and

30 by Sarus Cranes.

5.4.1 Temporal patterns

Though the number of cranes varied significantly between species (P = 0.002) and site (P <

0.001), the difference between traverses was not significant (P = 0.09) and there was no

significant interaction between species and traverse (P > 0.99). Furthermore, the abundance

of Sarus Cranes and Brolgas was strongly correlated across traverses (r = 0.84, n = 29, P

Page 104: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

104

<0.001). There is therefore no evidence of differential temporal patterns of use at a whole-of-

Tablelands scale for the months studied (May to December).

5.4.2 Spatial patterns

Among sites averaged across traverses, both the rank abundance and rank density of Brolgas

were positively but weakly correlated with those of Sarus Cranes (both tests, Spearman’s R =

0.39, n = 46, P = 0.007). There was significant clustering of sites based on the % of cranes

that were Brolgas (Moran's I = 0.84, Z = 9.1, P<<0.001), as well as significant hotspots for

each species (see Fig. 2 below). Brolga hotspots were located in the south (near Tumoulin

and Innot Hot Springs) and Sarus Crane hotspots were located in the central Tablelands (near

Atherton and Tinaroo Dam), with no significant clustering in the north (near Chewko and the

Mareeba Wetlands).

Page 105: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

105

Fig. 5.2 Foraging site hotspot analysis based on the percent of cranes that were Brolgas or Sarus.

Non-significant sites had an admixture of the two. Roost locations are also presented, classified

according to most prominent species observed at each.

Mean density of cranes did not differ between either distance to nearest roost class nor

species (P = 0.53 and 0.58 respectively) but the distance class by species interaction was

significant (P = 0.038). Post-hoc pair-wise analysis revealed no significant difference

between distance classes for Sarus Cranes (P all > 0.27) but for Brolgas there was one with a

significant difference (P = 0.01) and three with differences of P < 0.06. Brolgas were at much

higher density at feeding sites close to roosts than far from them (see Fig. 5.3 below). The

non-significant apparent reverse trend for Sarus Cranes is largely a product of a single high-

density feeding site 14.9 km from the nearest known roost (see Figs. 5.4 a and b below).

Page 106: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

106

Fig. 5.3 Species-specific crane densities at feeding sites as a function of distance to the nearest roost.

Fig. 5.4 Mean density of Brolgas (a) and Sarus Cranes (b) at each survey site. Roost locations are also

presented, classified according to most prominent species observed at each.

Page 107: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

107

5.4.3 Habitat use

For Sarus Cranes, MARSpline analysis (model R2 = 0.39) identified a positive association of

crane density with volcanic soils, negative association with elevation, and a preference for

moderate slopes. The positive association with higher fertility volcanic soils and strongly

diminishing density with increased elevation was also evident in the regression tree analysis

(Fig. 5.5A), but an effect of slope was not supported. For Brolgas, MARSpline analysis

(model R2 = 0.27) identified that Brolgas were negatively associated with distance from roost

(i.e. fed at sites closer to roosts) and were negatively associated with disturbance. Regression

tree analysis illustrates the effect of distance from roost on Brolga density (see Fig. 5.5B

below) but did not provide support for a directional effect of disturbance.

Fig. 5.5 Regression tree habitat models for Sarus Crane and Brolga. Values in boxes are the number

of sites and mean density of the crane species in the group, while values outside boxes are habitat

parameters. Higher densities are shown to the left.

Page 108: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

108

Brolgas were more often present at grazing and wetland sites than cropping sites but at higher

numbers and densities at cropping and grazing sites (see Table 5.2 below). Sarus Cranes were

more often present at wetlands, but most abundant and at highest densities at cropping sites.

Table 5.2 Population and density of crane species by land use type. Data are for site surveys. Sample

sizes: cropping – 1,300 site surveys across 31 sites; grazing – 423 site surveys across 11 sites; wetland

169 site surveys across 4 sites, with ‘% presence’ being the percentage of site surveys at which the

species was recorded.

Median (quartiles) when present

Species Land use % presence Population Density (ha-1

)

Brolga cropping 16.5 15 (4–51) 0.5 (0.1–1.6)

grazing 31.7 14 (4–27) 0.6 (0.1–1.4)

wetland 30.2 4 (3–9) 0.3 (0.1–0.3)

Sarus Crane cropping 20.8 19 (7–49) 0.7 (0.2–2.0)

grazing 20.3 4 (3–13) 0.1 (0.1–0.4)

wetland 30.2 3 (2–4) 0.1 (0.05–0.2)

The density of both crane species varied significantly among crop types after controlling for

the confounding effects of Site and Traverse. The effect of crop type was stronger for Sarus

Cranes than Brolgas (see Table 5.3).

Table 5.3 Probabilities from modelling of crop preferences of two crane species.

Pair-wise tests suggested three broad classes of crop preference for each species (see Table

5.4), with both species at highest densities where grain had been harvested and on ploughed

land with Brolgas also at high densities where grain was growing. This result was because the

Brolgas favoured a newly-sown site where they could feed on self-sown peanuts from the

previous crop.

Explanator Sarus Crane model Brolga model

Crop type <0.0001 0.003

Site (random effect) <0.0001 <0.0001

Traverse (random effect) <0.0001 0.017

Page 109: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

109

Table 5.4 Density of crane species in crop types on the Atherton Tablelands, Queensland, during

surveys from May to December from 2014-2016.

Mean crane density (birds/ha)

Crop (field) type Sarus Cranes Brolgas

Grains – harvested 0.95 0.66

Ploughed land 0.53 0.26

Maize – sown 0.33 0.09

Peanuts 0.25 0.03

Grains – growing 0.20 0.77

Potatoes 0.07 <0.01

Grazing and hay 0.05 0.01

Miscellaneous 0.02 0.00

Sugarcane 0.02 0.02

Maize – ripe <0.01 0.03

5.4.4 Crop seasonality

All four aggregated crop types (ploughed fields, growing grains, harvested grains, other)

were available in all months (May to December). Although relative crop frequency varied

significantly across months (Chi-square = 83.9, d.f. = 21, P << 0.001), there was no clear

seasonal trend in availability.

5.5 Discussion

Mixed flocks of crane species occur on all continents during the non-breeding season

(Johnsgard, 1983); although co-occurrence may mask minor differences in ecology between

species (Watanabe, 2006). On the Atherton Tablelands the differences in habitat use between

co-occurring Brolga and Sarus Crane were similarly subtle. There was no temporal separation

between species, both occurred at their highest densities and numbers where grain had been

harvested and on ploughed land and both species were found most frequently at wetland sites

to which they retreated when no cropping land suitable for crane foraging was available.

However, Brolgas (not Sarus Cranes) were most common close to their roost sites and were

found most commonly on grazing areas, while Sarus Cranes favoured the intrinsically more

fertile volcanic soils of the central Tablelands. The reasons the Sarus Cranes favoured this

Page 110: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

110

habitat and most Brolgas did not cannot be determined from a correlative analysis but the

differences are most readily explained by some form of low level competition between the

two species. In a captive feeding situation, Brownsmith (1978) found that Sarus Cranes

dominated Brolgas. While in the wild the reverse may be true in at least some circumstances

(T. Nevard, unpublished data), the largest Australian Sarus Cranes are larger than the largest

Brolgas (Marchant and Higgins, 1993), so there could be active exclusion if resources are

limiting, especially as Sarus Cranes, being larger, may then have higher nutritional

requirements. This could also explain not only the selection of the better soils by Sarus

Cranes but also differences in roost site proximity. While, for neither species, is the distance

travelled from the roost site unusual compared to other cranes (e.g. Eurasian Crane Grus

grus; Alonso et al., 2004), Brolgas may nevertheless be able to obtain their dietary

requirements more readily within a short distance of roost sites than Sarus Cranes, which

could be willing to travel further to feed. More speculatively, nutritional limitations may also

help explain the restricted distribution of Sarus Cranes in Australia where most soils have

intrinsic low fertility (Flannery, 1994) and more fertile soils until recently tend to have been

heavily vegetated. Grasslands on seasonally flooded alluvial soils near the Gulf of

Carpentaria, to which Sarus Cranes had been confined until the advent of arable cropping,

would have been an exception to this.

There are two main consequences for the choice of favoured agricultural habitat by

Australian Sarus Cranes: (i) the vulnerability of this habitat to agronomic trends; and (ii) the

novelty of the habitat.

Interviews with Atherton Tablelands farmers on whose land cranes have been recorded

(Nevard et al., 2018) indicate that there is a strong intention to shift towards tree crops that

give a return on investment an order of magnitude greater than the field crops favoured by the

Page 111: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

111

cranes. As has happened in India (Borad and Mukherjee, 2002) and Nepal (Aryal et al.,

2009), such shifts in crop preference can cause rapid declines in crane populations. Even for

those farmers intending to continue with field crops, some gains in yield are likely to arise

from reducing losses during harvesting, leaving little leftover grain available for cranes. This

would be consistent with findings in north America, where maize yield in the central Great

Plains increased by 20% from 1978 to 1998 by reducing post-harvest leftover grain to only

1.8% of production; a 47% reduction in the amount available for cranes (Krapu et al., 2004).

Similarly, in Western Europe, the amount of grain left behind dropped from 5–10% of maize

and 3–5% of wheat, barley and oats in the 1970s and 1980s to only 0–1% of maize and 1–2%

of other cereals today (Nowald, 2012). Further reductions in grain availability for cranes have

arisen as a result of adopting pre-emergent herbicides and new varieties and shortening

cropping intervals (Krapu et al., 2004; Prange, 2012).

The consequence of habitat loss is that cranes are likely to concentrate their foraging on the

small areas of suitable habitat remaining and potentially shift from foraging from waste food

to newly-sown seed. In North America and Europe this has resulted in conflict with farmers

(Nowald and Mewes, 2010). The results of farmer interviews carried out on the Atherton

Tablelands by Nevard et al., (2018) indicate that cranes can already cause economically

significant crop damage, especially where they forage in large numbers close to their roosts.

This has sometimes resulted in unlawful killing of cranes (Australian Broadcasting

Corporation, 2013).

The second conclusion to be drawn from the preference of Sarus Cranes for the better soils is

that this preferred habitat is almost entirely relatively new. The tropical savanna woodland

preferred by Brolgas, and most of the swamps beside which they roost, were present at the

time of European arrival in the area in the 1880s (albeit partially modified since by cattle

Page 112: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

112

grazing) and probably for thousands of years, given the environmental history of the area

under Indigenous occupation (Mooney et al., 2017). The fertile volcanic soils, however, were

almost entirely covered with rainforest until the first half of the 20th

Century. Such habitat

would have been unsuitable for cranes. As it is, Sarus Cranes were only observed on the

Atherton Tablelands for the first time in 1967, at which time numbers overall were thought to

be far lower than today (G. Archibald pers. obs.). This implies that the migration to the

Tablelands has been occurring for just a few generations, given that cranes commonly live for

several decades (Johnsgard, 1983).

In ungulates, migratory pathways established through social processes take 8-9 generations to

become fully established among naïve animals, even when they are joining an established

migration pathway (Jesmer et al., 2018). The rapidity with which Australian Sarus Cranes

have been able to locate and exploit the newly formed habitat on the Atherton Tablelands

suggests that, spatially, they are highly adaptable and opportunistic. This would be consistent

with our own observations of Brolgas and Sarus Cranes foraging in grain crops grown on

newly-developed arable areas in Cape York Peninsula (250 km north if the Atherton

Tablelands) and near Georgetown (300 km west) in May 2018; as well as near Emerald in

2014-5, some 900 km south (W. Cooper, pers. comm.).

5.6 Conservation Management

In areas with economically significant crop damage on the Atherton Tablelands, Nevard et al.

(2018) found that the majority of farmers interviewed were prepared to adopt appropriate

agronomic practices, such as crane repellents, to avoid damage. While encouraging, the

consequence of this could be similar to a progressive change in agronomic practice, i.e.

concentration of cranes on fewer and fewer farms, where the damage they would cause would

become increasingly severe. However, as detailed in Nevard et al. (2018), this may provide

Page 113: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

113

positive opportunities for ecotourism on the Atherton Tablelands, where an Annual Crane

Week celebrates the presence of large numbers of the birds. To succeed, this as yet immature

initiative needs coordination with agronomic practices in surrounding farmland, particularly

the spread of benefits so that farmers incurring economic damage from cranes receive

adequate compensation.

To achieve this would require detailed evaluation of the crane-related financial losses of a

representative sample of farming businesses, ideally in partnership with farming

organisations and local conservation organisations. This could then provide the basis for the

development of a package of measures for discussion with all levels of government and

potential sponsors. Nevard et al. (2018) identified that due to its proximity to the Cairns

tourism market, about AU$125,000 of annual gross regional product could potentially be

generated from crane watching on the Atherton Tablelands. However, the temporal window

for this opportunity could be curtailed if cranes adapt to habitat constrictions on the Atherton

Tablelands by moving to more remote newly-developing agricultural lands.

5.7 Conclusions and Recommendations

Arable cereal crops (especially maize) and peanuts are important to dry season crane flocks

on the Atherton Tablelands. While we have not established whether the area suitable for

crane foraging is currently limiting their abundance, constrictions are likely if changes to

arable cropping on the Atherton Tablelands reflect global agronomic trends towards

increased efficiency in harvesting and agronomy of field crops or their replacement with

higher value perennial crops.

However, the relative novelty of crane migration to the Tablelands, particularly for Sarus

Cranes, suggests that they may seek out and exploit other agricultural lands if, as appears to

Page 114: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

114

be happening, these are created elsewhere in north Queensland. If not, there are opportunities

for adaptation by combining the adoption of crane repellent techniques with the provision of

food or sacrifice crops in association with an expansion in crane-related tourism.

Page 115: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

115

CHAPTER 6 - FARMER ATTITUDES

Farming and Cranes on the Atherton Tablelands, Australia

[Accepted for publication in Pacific Conservation Biology 01/07/2018]

6.1 Abstract

Brolgas Antigone rubicunda and Australian Sarus Cranes A. antigone gillae, form dry season

flocks on the Atherton Tablelands in north Queensland, Australia; where they forage almost

exclusively amongst planted crops or intensively managed grassland. The long-term

relationship between cranes and farmers is therefore critical to their conservation, especially

as cranes can sometimes cause significant economic damage to crops. Farmers were

interviewed to explore their current attitudes to cranes and their intentions for land use that

might affect the birds. Most farmers were found to tolerate cranes, particularly when they fed

amongst stubble. Most, however, are increasing the efficiency of their agronomic practices,

harvesting combinable crops such as maize and peanuts in ways that are beginning to reduce

post-harvest crop residues. There is also a rapid trend away from field crops to perennial and

tree crops that have a higher return per unit area. Both trends may reduce foraging

opportunities for the cranes and, unless managed effectively, are likely to increase the

potential for damage and conflict with farmers in the field crops that remain.

6.2 Introduction

When the importance of agricultural lands for biodiversity is discussed, it is usually the

complexity of the matrix that is emphasised (Fahrig et al., 2011). Relatively few bird species

become synanthropic and benefit from agricultural land. In Australia, where agriculture has

been a dominant part of the landscape for little over 100 years (Garnett and Crowley, 2008),

Page 116: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

116

there are particularly few synanthropic native species. Among birds, for example, 60 native

species are recorded as localised pests of fruit crops (Tracey et al., 2007) and just eight

species in grain crops (Bomford and Sinclair, 2002); all with extensive ranges outside the

crops in which they cause problems.

Internationally, agricultural specialists are highly susceptible to changes in agronomic

practice. This has been most apparent in Eurasia, where a broad suite of farmland birds has

been disadvantaged by both intensification of cultivation systems and shifts from cultivation

to pasture (Robinson et al., 2001). In both Eurasia and North America, however, a small

number of large granivorous and herbivorous species have flourished in cultivated landscapes

on which they are increasingly dependent (Amano et al., 2008; Le Roy, 2010; Sugden et al.,

1988). Such is their abundance that they cause significant economic losses (Heinrich and

Craven, 1992; Lane et al., 1998; Lorenzen and Madsen, 1986).

Agricultural opportunists include members of the crane family (Gruidae), with nearly all of

the 15 species relying on agricultural lands during at least part of the year (Harris, 2010). The

extent to which the two crane species that occur in Australia, the Sarus Crane Antigone

antigone and the Brolga A. rubicunda (see Figure. 6.1 and Figure 6.2), rely on agriculture is

not fully understood but both species do use agricultural lands extensively during the non-

breeding season on the Atherton Tablelands in north Queensland, consuming grain, insects

and even small vertebrates (Meine and Archibald, 1996).

Page 117: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

117

Fig 6.1 Brolgas feeding on volunteer peanuts in an irrigated winter wheat crop at Innot Hot Springs,

Atherton Tablelands, Queensland. [Photograph: T. Nevard]

Fig. 6.2 Pairs of Brolgas and Australian Sarus Cranes with cattle at Kaban, Atherton Tablelands,

Queensland. [Photograph: T. Nevard]

Page 118: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

118

The Brolga population in northern Australia is generally considered stable and has been

estimated to comprise of 20,000-100,000 individuals (Meine and Archibald, 1996). Of 51,969

Brolgas recorded during a national census of waterbirds, all but 135 were in northern

Australia (Kingsford et al., 2012). The only longitudinal counts of Brolgas in tropical

Australia have been from north-east Queensland (Atherton Tablelands and Townsville

regions); with an annual average of approximately 800 individuals recorded by community-

based naturalists over 21 years between 1997 and 2017 (E. Scambler pers. comm. and Nevard

et al., 2019a and 2019b).

The 800-1,500 Sarus Cranes (sources as above) counted annually on the Atherton Tablelands

are globally important for a species classified as Vulnerable in the IUCN Red List (BirdLife

International, 2018c) and are the reason the Atherton Tablelands are registered as a Key

Biodiversity Area (BirdLife International, 2018b). A substantial proportion of the total

population of the endemic Australian subspecies A.a.gilliae migrates annually to the

Tablelands from their breeding grounds 400 km north-west near the Gulf of Carpentaria

(Nevard et al., 2019b). As with cranes in Eurasia, Africa and North America

(Andryuschenko, 2011; Prange, 2012 and Nowald and Mewes, 2010), foraging among crops

has caused damage on the Atherton Tablelands. There has also been one case where a

farming company was found guilty of poisoning (Australian Broadcasting Corporation,

2013).

Conflict between cranes and agriculture outside Australia has been managed in four ways that

do not entail lethal control: (i) compensation to farmers, an approach adopted in Europe and

North America with some success although the cost is rising (Nilsson et al., 2016). Surveys

suggest that there is guarded support for such programs among farmers in north Queensland

(Greiner, 2015); (ii) a crane repellent which makes cranes nauseous is available in the United

Page 119: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

119

States, but has not yet been tested in Australia; (iii) cranes can be diverted from farms by

providing grain elsewhere in the landscape near their major roost sites such as Lake

Hornborgasjön in Sweden and Lac du Der in France (Lundgren, 2013; Salvi, 2015); (iv)

cranes can become an economic asset as a focus for tourism, such as at Tsurui in Japan and

Lac du Der in France. The Wildlife Conservancy of Tropical Queensland and the Tablelands

Regional Council have established Atherton Tablelands Crane Week (www.craneweek.org)

to raise the profile of cranes in north Queensland and explore potential farm-based crane-

viewing opportunities for both domestic and international visitors. However, none of these

solutions can be implemented without first understanding farm practice, farmer attitudes and

their intentions (Manfredo, 1989). Such research has not been conducted in Australia. The

aim of this research was therefore to provide an insight into the views of Atherton

Tablelands’ farmers and the interaction of cranes with their livelihoods as a prelude to

identification of crane management options for the Atherton Tablelands.

6.3 Study Area

The study was conducted in agricultural lands on the Atherton Tablelands, an area of

approximately 6,000 km2 of which about 500 km

2 is under intensive agricultural production.

The region is highly diverse with three main landscape elements: (i) rainforests, which are

mainly part of the Wet Tropics World Heritage Area; (ii) areas of dry sclerophyll savanna

woodland; and (iii) some of Australia’s most fertile and productive farmland. Cranes are

absent from rainforest but occur in the latter two land types, mainly during the dry season

flocking months of May to December (see Fig. 6.3).

Page 120: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

120

Fig. 6.3 Atherton Tablelands region showing location, major settlements, land use and Key

Biodiversity Area (data from Queensland Globe and Department of Environment & Energy, 2012).

Typical agricultural soils in the northern and southernmost areas are derived from granite and

metamorphics and have a slightly acid pH and low fertility, compared to the well-drained

soils in the central Tablelands derived from basalt. The latter areas contain some of the most

expensive arable land in Queensland, with land and associated water allocations in 2018

being sold for up to AUD80,000/ha (J. Suffield, Elders Real Estate, pers.comm.)

Water for the Mareeba-Dimbulah Irrigation Area is supplied from a large water impoundment

on the Barron River, Tinaroo Dam, through an extensive network of channels and streams.

Farming enterprises in other parts of the Tablelands (including intensive agricultural areas

around Atherton, Malanda and Ravenshoe) draw water from volcanic groundwater aquifers,

natural watercourses and small dams.

Page 121: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

121

6.4 Methods

Information on the relationship between cranes and agriculture on the Atherton Tablelands

was obtained from two main sources.

1. Data on agricultural systems in the study area were obtained from Queensland

Department of Agriculture and Fisheries (Department of Agriculture and Fisheries,

2015). Based on this cropping data and farmers’ experience, we have rated the

intensity of use by cranes as low, moderate or high.

2. Attitudes and socioeconomic conditions under which farmers and graziers operate on

the Atherton Tablelands were explored using informal landholder interviews

(approved by the Charles Darwin Univerity Human Ethics Committee), similar to the

methodology previously used successfully by Garnett et al. (2016) to test landholder

support for Night Parrot Pezoporus occidentalis conservation.

All forty-five of the 260 farmers and graziers listed in the Yellow Pages from the Atherton

Tablelands (Sensis, 2015) on whose land identified as being used by cranes were contacted

for interview. Each of those 31 who agreed to be interviewed (14 declined or did not respond)

described their farming experience, the composition of their enterprises, their views on nature

conservation, knowledge of and attitudes towards cranes, damage by cranes and other

wildlife, potential solutions for reducing crop damage and attitudes towards crane tourism

and crop damage mitigation trials. They also expressed their intentions for their enterprises in

the future.

Interviews based on questions approved by the Charles Darwin University Human Ethics

Committee took place between February 2015 and March 2016. Before commencing, each

interviewee was first assured of the anonymity of their responses and location of their land.

Each interview was conducted in person and transcribed during the interview, with transcripts

Page 122: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

122

read back at the end of each interview and opportunity given for additional comments or

correction.

Two forms of analysis were conducted based on the landholder survey: (i) descriptive

statistics based on ranking to summarise knowledge and attitudes; and (ii) textual sampling of

comments made by individual interviewees to convey the tone of the interviews and provide

context.

6.5 Results

6.5.1 Primary Production on the Atherton Tablelands

The Queensland Department of Agriculture and Fisheries (2015) identified 40 agricultural

land uses in the Atherton Tablelands region (see Tables 6.1 and 6.2).

Table 6.1 Agricultural production on the Atherton Tablelands, Queensland, in 2015 sorted by area for

each major type of agricultural activity and the intensity of use by Brolga and Sarus Cranes [Source:

Department of Agriculture and Fisheries (2015) and this research].

Agricultural activity Area (ha)

Gross revenue

per hectare/yr

(AUD '000)

Crane use

intensity

Livestock 558,872 0.2

Beef cattle 550,000 0.1 Low

Dairy cattle 8,800 3.9 Moderate

Pigs and poultry 72 550.4 Low

Field Crops and Horticulture 23,427 4.6

Sugar cane 10,956 3.6 Moderate

Maize 4,719 2.4 High

Hay 3,020 1.2 Moderate

Grass seed 1,195 4.0 Moderate

Potato 972 16.2 Moderate

Legume seed 968 3.1 Moderate

Peanut 874 5.5 High

Minor field crops 723 35.2 Low

Tree crops and forestry 11,297 27.8

Forestry plantations 3,600 5.6 Low

Minor tree crops 2,497 34.5 Low

Mango 2,400 21.1 Low

Banana 1,850 49.2 Low

Page 123: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

123

Agricultural activity Area (ha)

Gross revenue

per hectare/yr

(AUD '000)

Crane use

intensity

Avocado 950 87.3 Low

Miscellaneous 151 577.2 Low

Total 593,747 910.4

Table 6.2 Returns from land on the Atherton Tablelands, Queensland, with different intensities of dry

season crane usage [Source: Department of Agriculture and Fisheries (2015)]

Potential for crane usage Area (ha)

Gross revenue per

hectare (AUD '000)

High (maize and peanuts) 5,593 2.9

Moderate 25,911 3.9

Low (intensive use) 12,243 31.8

Low (beef cattle) 550,000 0.1

Beef cattle production was the dominant agricultural land use over the wider region,

accounting for 93% of the total agricultural area; largely on low fertility savanna woodlands.

The main agricultural use in areas of high fertility soils was cropping (e.g. sugar cane, maize

and hay), which accounted for more than half of the intensely farmed areas; followed by

dairy cattle and tree crops (e.g. forest plantations, minor tree crops, mango and banana).

Returns per hectare on tree crops, were an order of magnitude higher than for field crops or

dairy cattle.

6.5.2 Farmer intentions and attitudes

Of the 45 landholders within the areas of the Atherton Tablelands visited by cranes in the dry

season, 31 participated in interviews. This sample represents approximately 12% of all

farmers in the region.

The properties that interviewees managed were widely distributed across the Atherton

Tablelands and ranged from c. 30 to c. 13,500 ha. They were located on irrigated and non-

irrigated land and engaged in farming livestock (dairy and beef), cropping (sugar cane, maize,

Page 124: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

124

sorghum, peanut, potato, soya, canola, Rhodes grass, wheat, barley, oats, pumpkin) and/or

tree crops (avocado, banana, lychee, coffee). Many were mixed enterprises.

Interviewees had a median of 38 years of experience in primary industry (20-69 years) with a

median of 25 years on their property (9-55 years). Thirteen interviewees indicated that they

thought their children would be likely to take on the property and 18 owned or managed more

than one property. Nineteen were partnerships or sole traders with 12 having a company

structure. The median percentage of income derived from farming was 80%, with the average

contribution of primary industry to total income being 68%.

All interviewees were questioned about their cropping and livestock husbandry practices.

Eighteen were involved in annual cropping, 13 had livestock, nine had tree crops and ten

grew sugar cane. When asked to speculate on the future of farming on the Atherton

Tablelands, all interviewees felt that more tree and perennial crops would be planted and nine

indicated that more efficient crop management, harvesting and sowing was already being

adopted. Two quotes from the interviews underline this: “High value crops such as

avocadoes, bananas and sugar are increasing, and we’re almost the only large-scale maize

and peanut growers left in our area” and “As cropping is becoming specialised and needs new

efficient equipment to be profitable, there’s a general move towards avocadoes, sugar and

bananas on the better crop land and we’re planting avocadoes”.

To provide a context for opinions, all interviewees were asked to rank the following six

socioeconomic issues in order of importance to themselves: economic growth; nature

conservation; healthcare; education; employment; and immigration. The majority of

respondents ranked economic growth as the most important. Education was second, with

health care and employment joint third, and immigration and nature conservation least

important (see Table 6.3).

Page 125: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

125

Table 6.3 Landholder ranks of social issues (ranked from 1 (highest) – 6 (lowest); n=30)

Policy issue Average rank No. ranking highest No. ranking last

Economic growth 2.1 18 0

Education 1.9 11 0

Conservation 4.7 0 8

Healthcare 2.8 2 6

Employment 3.2 0 5

Immigration 5.0 1 15

Most interviewees (23) felt that landholders should generally look after wildlife but a quarter

did not. A typical response was: “Society in general has a responsibility for wildlife, not just

farmers and government. Consumers should be paying an ‘environmental premium’ for food,

as we have to get the marketplace to pay”. Some financial responsibility for managing

wildlife was conceded by 21 interviewees, with ten disagreeing “Landholders should bear

some burden in the sense that profitable businesses pay more tax but it is essentially a

taxpayer responsibility”. Only six interviewees thought that government was looking after

threatened species well in or off National Parks. One interviewee said: “They’re making a

mess of it and sitting in offices rather than spending time in the bush”.

Interviewees were roughly equally split on whether the local Council should play a bigger

role in helping to conserve wildlife on private land (15 for and 16 against). Typical responses

from both sides were: “Council could (play a bigger role) but it doesn’t have the resources,

especially when our roads are falling apart” and “it’s not their job”.

Nearly all (27) interviewees could distinguish between Brolga and Sarus Crane species, with

11 demonstrating knowledge of their breeding and foraging habits. The median length of time

Page 126: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

126

that interviewees judged that they’d had cranes on their property was 25 years (10-50 years),

for example one respondent indicated that they’d had cranes “since the Mareeba Wetlands

was created around 20 years ago”.

Twenty-eight interviewees said that cranes first arrived each year in May and June (12 and 16

respondents respectively) and 26 said that the cranes left in November and December (6 and

20 respectively). A smaller number of interviewees (5), all of whom had predominantly

Brolgas on their land, said that the cranes left in January and February.

Twenty-four respondents indicated that cranes foraged on their land and six that they both

foraged and roosted. Only one indicated that fewer cranes foraged on their land due to

cropping changes: “They feed on our farm but don’t roost. As our agronomy and harvesting

gear has got more efficient, we get less Brolgas and Sarus Cranes than we used to a few years

ago”. When asked if their stock interacted with Brolgas and Sarus Cranes, 13 indicated that

this happened, for example “The cranes seem to like the company of stock and will spend a

lot of time walking among them”.

When interviewees were asked to describe how Brolgas and Sarus Cranes used their land, 11

indicated that a mix of maize, peanuts and recently cultivated land had the highest number of

crane visits, followed by a mix of maize and peanuts (eight respondents); a mix of livestock

and grass, sugar cane, and cultivations were each nominated by three respondents. Just over

half (16 respondents) reported no damage from cranes with 14 reporting damage to peanuts

and 11 to maize. When asked about the critical time for damage, 20 nominated after sowing,

two pre-harvest and nine had no damage so offered no opinion.

When asked to nominate other animals causing damage, 19 interviewees grouped Magpie

Geese, Whistling Ducks and Cockatoos together as causing damage, six indicated only

Page 127: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

127

Magpie Geese, three (all graziers) named dingoes and two feral pigs. Twelve respondents

took no measures to prevent damage, eight concentrated on human intervention such as riding

a motorbike in the paddock, five combined mechanical measures such as bird scarers with

human intervention, for example “scaring-off by driving a ute and beeping the horn and gas

guns” and four graziers poisoned dingoes and feral dogs. Agronomic measures, such as

repellent seed coatings, and human intervention were considered potentially effective in

reducing bird damage by seven and six interviewees respectively, one comment being:

“We’ve heard that they use a repellent for seed corn and potatoes to stop cranes and geese in

the US”. Eleven interviewees thought that no measures are effective in reducing damage.

Nineteen indicated that they would be interested in participating in trialling measures to

reduce crop damage by Brolgas and Sarus Cranes, a typical comment being: “yes but we

don’t have much time”.

Interviewees were asked about their knowledge of or involvement in crane conservation or

research activities, to ascertain regional interest in cranes amongst both farmers and graziers.

Twenty-four said that they had heard of the annual Atherton Tablelands crane count and 23

said that they had heard of Crane Week, a local celebration of crane abundance. Six

interviewees felt that crane-based tourism could potentially provide a source of income for

their property, including one individual with a degree in hospitality and tourism. Twenty said

they would not welcome birdwatchers on their property: “we’d rather not have them”.

6.6 Discussion and Recommendations

Established global trends to reduce harvesting losses through machinery and agronomic

efficiency improvements and minimise fallow periods tend to reduce crane foraging

opportunities (Prange, 2012). During the dry season, cranes on the Atherton Tablelands were

observed to feed primarily on unharvested grain and nuts in the stubble of maize and peanut

Page 128: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

128

crops, and to a lesser extent on insects extracted from among hay crops, potatoes, legumes,

seed crops and, post-harvest in sugar cane, grass crop and dairy fields. As agronomic and

harvesting efficiencies are taken-up, these opportunities are likely to become far fewer.

Alongside agronomic and harvesting efficiencies, in many areas where production costs are

high, perennial and tree crops are replacing combinable crops eaten by cranes (grains,

oilseeds and peanuts), which have significantly lower returns per hectare than those with low

crane suitability (see Table 6.2). This is illustrated on the Atherton Tablelands by the banana

industry, which had the highest gross income for any crop in 2015 (AUD90.9 million)

followed by avocadoes (AUD82.9 million). The area under both crops has therefore

expanded substantially in recent years (Department of Agriculture and Fisheries, 2015), as

have other high return crops such as citrus (AUD31.3 million) and pawpaw (AUD15.1

million).

The pattern of cropping is also potentially likely to shift as, in cooperation with the

Queensland State Government, the Tablelands Regional Council is investigating the

establishment of further irrigated cropping areas (10-15,000 ha) in the southern Atherton

Tablelands; south and west of Ravenshoe (Dr Hurriyet Babacan, CEO of Tablelands

Regional Council, pers. comm.). Although it is anticipated that annual crops will continue to

be grown in this area, rises in the price of water could have the potential to tip the balance in

favour of higher value tree crops.

Given this background, the results highlight three concerning factors for the future of cranes

on the Atherton Tablelands:

(i) Gross returns per hectare from crops that cranes use frequently (e.g. maize and

peanuts) are an order of magnitude lower than crops such bananas and avocadoes,

Page 129: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

129

with low intensity cranes use. As anticipated by the farmers interviewed, with

rising labour, land, machinery, input and irrigation costs there will be substantial

economic pressure to convert many areas of combinable crops to higher value

agricultural activities that tend to be unsuitable for crane foraging.

(ii) There are growing economic imperatives to increase harvest and agronomic

efficiency in combinable crops. Many farmers commented that cranes favoured

post-harvest and cultivated fields because maize and peanuts are lost in the

harvesting process and remain available to the cranes. More efficient farming

techniques and changing cropping patterns will reduce the availability of this post-

harvest grain residue (Krapu et al., 2004; Austin, 2012; Prange, 2015; Nilsson et

al., 2016). Indeed one farmer from the Atherton Tablelands commented that his

increased efficiency had meant cranes no longer visited his property. Increased

harvest efficiency has been linked to a reduction in food availability and

population declines/range shifts for cranes in many countries (Andryuschenko,

2011; Prange, 2012; and Nowald and Mewes, 2010).

(iii) Although the scale of poisoning remains unknown on the Atherton Tablelands,

van Velden et al. (2016) found that farmers who had experienced large flocks on

their farms and farmers who ranked cranes as more problematic than other bird

pests were more likely to perceive cranes to be damaging their livelihoods.

Biographical variables and crop profiles were not correlated with perceptions of

damage, indicating the complexity of this human-wildlife conflict. However,

farmers’ need for management alternatives was related to the perceived severity of

damage. For ethical reasons, poisoning was not discussed during interviews

because it is illegal and could have either implicated the interviewee or made the

interviewer complicit in the illegal activity. Nevertheless, many farmers

Page 130: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

130

mentioned that their crops had been damaged by cranes immediately after sowing

and that they felt crop losses to be severe. They were also well-informed about the

recent prosecution for poisoning cranes.

The interview process found that farmers on the Atherton Tablelands are well aware of

cranes, with every interviewee having some knowledge of the timing of arrival and departure

on their properties. Importantly, some interviewees valued the cranes in their own right

beyond economic and financial considerations. For example, one believed that “as

landholders, farmers should recognise their responsibility for the wildlife on their farms and

contribute”. However there was also a general undercurrent of caution amongst farmers (less

obvious amongst graziers, who do not suffer damage) in relation to talking to outsiders about

cranes on their land. Given the potentially adverse trajectory of cropping change for cranes

on the Atherton Tablelands, there is a need for well-considered management options

developed cooperatively with farmers. However, farmer intentions revealed through

interviews suggest that changes in crop selection and agronomic practice do not auger well

for the populations of Brolga and Sarus Cranes on the Atherton Tablelands Key Biodiversity

Area, notwithstanding the differences between stated intentions and actual action (Fishbein

and Ajzen, 2011).

6.6.1 Crane management options

All of the following four options for managing conflict between cranes and agriculture could

be relevant to the Atherton Tablelands:

(i) Compensation to farmers. While compensation paid for damage caused by cranes

would be unlikely to overcome the discrepancy in returns between tree and

combinable crops, it could reduce the market incentive to switch out of maize and

Page 131: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

131

peanuts. However, because neither Brolga nor Sarus Cranes are listed as

threatened under Commonwealth or State legislation, compensation may not be

available from government, as funding tends to be given only to listed taxa.

Compensation funds would therefore have to be drawn from private philanthropic

funds.

(ii) Crane repellents. The State and Commonwealth Governments could facilitate

research to test commercially available crane repellents in Australia, especially as

they are already licensed for use in the United States, South Africa and parts of

Europe. The repellents protect newly emerging grain (Barzen and Ballinger, 2017)

which farmers noted was the most susceptible to damage by foraging cranes.

Attempts by private individuals to import the repellent for trial over the last ten

years have been thwarted by Commonwealth agrochemical regulations that

require expensive Australian trials prior to approval for release (G. Harrington

pers. comm.).

(iii) Diversion to alternate foraging areas. Active feeding of cranes at particular sites

could divert cranes and reduce damage on valuable agricultural land, and sites

could be selected to coincide with tourism opportunities. Such schemes exist in

Europe with annual contributions of up to 100,000€ towards the cost of

supplementary food and site management being provided by regional government

authorities (Le Roy, 2004; Nowald, 2012; Nilsson et al., 2016). The Tablelands

wintering population of cranes (Brolga and Sarus) is currently about 3,000

individuals. Blackwell et al. (2002) found that in captive situations, a pair of

Sandhill Cranes Antigone canadensis consumed 231.4g of maize a day. The

average weight of male and female Sandhill Cranes is 3,550 g, Brolgas 6,359 and

Australian Sarus Cranes 7,800 g (Johnsgard, 1983). This equates to an

Page 132: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

132

approximate Australian crane/Sandhill Crane weight ratio of approximately 2:1

(7,078g/3,550g). As food consumption is unlikely to be directly related to weight

and Blackwell’s findings are not from a wild situation (where other food sources

are available), for the purposes of calculating an order of magnitude of maize that

could be required to replace crane crop raiding on the Atherton Tablelands, a wild

Australian crane/captive Sandhill Crane maize consumption ratio of 1:1 has been

adopted. Assuming that the approximately 3,000 wintering cranes (1,500 ‘pair

equivalents’) remain on the Atherton Tablelands for an average of 180 days (June-

November) they would consume about 62,000 kg of maize a year. At current

prices this would cost AUD 22,000 p.a.

(iv) Ecotourism. Birding is a growing sector in the Australian tourism market (Steven

et al. 2015) and, in the absence of government-sponsored agri-environmental

measures, nature-based tourism has been identified as a potential ecosystem

service for which transfer payments can be made to assist with conservation

(Balmford et al. 2009). Whilst most interviewees were not interested in having

birdwatchers on their land, there was some interest in using cranes to generate

income from tourism among interviewees, especially by those landholders close

enough to Cairns to benefit from day visitors. The Atherton Tablelands Crane

Week aims particularly to take advantage of the large numbers of Japanese and

Chinese tourists who visit north Queensland. In both countries cranes are

venerated and have a number of crane-led visitor destinations (e.g. Tsurui in

northeast Hokkaido, Japan) with numerous farm-based/council-funded crane-

viewing opportunities (Y. Momose, pers. comm. and Matthiessen, 2014). Other

important north Queensland tourism markets, such as the United States, France

and Germany, also have well-developed crane-based tourism (Nowald, 2012),

Page 133: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

133

potentially creating the opportunity for a global network of crane viewing

opportunities. In the Hula Valley of northern Israel, large wetland areas have been

drained for agricultural expansion and the land used for grain crops. Migratory

Eurasian Cranes Grus grus started wintering, with numbers building to 100,000

and crop damage becoming substantial. Feeding stations were established to draw

cranes away from growing crops, to which visitors were transported in portable

viewing stations. Funds generated by current visitor numbers of up to 400,000 a

year contribute to crane food and on-site management. (Shanni et al., 2012 and

Yossi Leshem, pers. comm.). On the Atherton Tablelands returns to the local

economy from, for example, an extra 5,000 tourists p.a. could exceed the cost of

paying to feed visiting cranes by an order of magnitude (e.g. a local restaurant

meal of AUD25 per head could equate to AUD125,000 of additional gross

regional product).

6.6.2 Consequences of failing to conserve cranes on the Atherton Tablelands

For Sarus Cranes at least, there is some evidence that numbers have greatly increased since

the 1960s, possibly because of the abundance of food in the dry season provided by

agriculture (Archibald, 1987). Given that the population of Sarus Cranes using the Atherton

Tablelands is thought to be only 1,500 at the most (Elinor Scambler, pers. comm.), the loss of

this population would mean that the only birds remaining would be those that do not migrate

to the Atherton Tablelands from their breeding grounds on the Gulf Plains, 500 km to the

west. The size of the non-migratory population is unknown but is likely to be much smaller

than the Tablelands population, meaning that the subspecies of Sarus Crane in Australia

could be likely to meet IUCN Red List Criteria as vulnerable.

Page 134: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

134

However, were habitat no longer available on the Atherton Tablelands, the situation could

revert to that which occurred before the land was cleared, the tree crops keeping cranes off

fertile regional soils, just as rainforest once did.

Should this transpire, the cranes of the Atherton Tablelands, which could potentially be

globally significant for conservation and form the basis of an incipient tourism industry,

might not endure. This research provides the basis for discussions between primary producers

and government, conservation and tourism interests about potential measures for retaining

cranes by reducing crop damage and expanding opportunities for tourism. Identification of

sources of funding for physical measures and compensatory payments would need to be part

of these discussions.

Page 135: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

135

CHAPTER 7 – DISCUSSION

7.1 Introduction

Cranes are iconic birds and have close cultural relationships with people, forged over

thousands of years. Eulogised in storytelling (such as the Australian Aboriginal Buralga

myth), dance, poetry and song; depicted in art; celebrated in traditional cultures; used as

contemporary symbols of trust and fidelity; they are prominent flagships for wetland

ecosystems and unsurpassed symbols of natural heritage. However, they are becoming

increasingly dependent upon modified agricultural and pastoral landscapes, requiring a

sustainable balance to be struck between the needs of cranes and those of farming.

A key challenge to achieve this is that landscapes important for cranes are also generally

those most suitable for farming. Global agriculture is intensifying rapidly, with its trajectory

set towards feeding a burgeoning global human population, which, with the accelerants of

increasing personal wealth and dietary change (explored in Chapter 1), may need 50% more

food by 2030, doubling by 2050. As a result, especially where unmoderated by cultural

values or viable economic alternatives, farmers and cranes are coming into conflict.

Balmford et al. (2019) argue that whilst land-sparing approaches to meeting human

agricultural demand remain the least detrimental to biodiversity, practical difficulties in

implementing land sparing could still have significant impacts on biodiversity. Bull et al.

(2017) have also explored this in the context of offsets, raising concerns about potential gaps

between their theoretical and practical effectiveness. As cranes are synanthropic, with most

species having at least some level of dependency on agriculture, practical crane-specific land-

sharing measures are required within the landscapes they share with farmers. Strategies

which explicitly link yield increases could therefore be extremely valuable in helping to

Page 136: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

136

offset increased crane damage, brought about by agronomic change and increased

efficiencies.

Notwithstanding recent progress in the debate around ‘sustainable intensification’ it is not yet

seen as mature enough at either policy or agronomic levels to more than marginally affect the

focus of commercial agricultural development; which remains firmly production-oriented.

Whilst increasing yields once underpinned significant increases in the populations of some

crane species (Nowald, 2012), harvesting efficiencies are now inexorably increasing pressure

on crane populations (Krapu, et al., 2004) and practical measures to protect cranes within

agricultural landscapes are urgently required.

Against this backdrop, the ICF (Austin et al., 2018) has identified inter alia a range of

research that should be undertaken to support the complex decision-making required to

address the changes at the agriculture/crane interface. This includes: (i) understanding the

crane-related dietary and bioenergetic characteristics of grains and other crops; (ii) formal

delineation and protection of flyways and important stopover areas; (iii) understanding

habitat connectivity and patterns of habitat loss; (iv) identifying spatial and temporal risks

from factors such as drought, fire, crop protection, power lines and avian diseases; and (v)

better understanding the key socioeconomic factors at the agriculture/crane conservation

interface. To create workable models across such a complex matrix of issues requires

adaptive, multidisciplinary approaches; founded on sound research and monitoring, and

generating high quality data to underpin appropriate practical and policy-related conservation

interventions.

Page 137: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

137

7.2 Key Findings

The global backdrop against which my research findings and their implications have been

discussed is set out in each of Chapters 1, 3, 4, 5, and 6 above. For completeness, the

following summary discusses five key findings of the research and goes on to postulate a set

of potential future scenarios that could affect cranes in tropical Australia.

Key Finding 1: The Australian Sarus Crane is a distinct subspecies; significantly different

from Asian mainland subspecies, underlining the conservation importance of the Australian

population. It also seems likely that the Australian and extinct Philippine subspecies share a

close evolutionary relationship.

For the formulation of robust and durable crane conservation strategies and framing of

appropriate supporting legislation, it is first essential to confirm the taxonomic status of

Australia’s cranes. Miller et al. (in press) have done this for the Brolga; finding that the large

northern population is slightly different genetically from the small and potentially threatened

southern population but not differentiated sufficiently for subspecies status. This research

(see Chapter 3) has explored the genetic differences between the Australian Sarus Crane and

extant Asian subspecies; finding that it differs substantially and potentially clusters

genetically with the extinct Philippine subspecies.

Typically, genetic drift occurs in small populations, where rare alleles are more likely to be

lost. The result of this process is reduced genetic diversity and increased genetic

differentiation (Hedrick, 2011). Compared to the two extant Asian subspecies, the genetic

diversity of Australian Sarus Cranes is significantly lower, as it may have passed through a

genetic bottleneck around 28,000 years ago, soon after its colonisation of Australia (Jones et

al. 2005). The implications of this for any captive breeding program are significant. Should it

Page 138: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

138

be deemed appropriate to assemble a representative captive group of Australian Sarus Cranes

to ensure that its unique genetic characteristics can be preserved from ongoing introgression

(see Chapter 4 and below), it would be important to ensure that the founders of the group had

the broadest diversity possible. This would help to avoid problems such as those encountered

with existing captive populations of the White-headed Duck Oxyura leucocephala, which

were found genetically unsuitable for reintroduction; requiring the re-building of a more

diverse captive population, based on birds taken from different areas of their range (Munoz-

Fuentes et al., 2008).

As cranes are large charismatic birds, they have been the subject of a number of ‘flagship’

reintroduction projects, such as the ‘Great Crane Project’ in England

(www.thegreatcraneproject.org.uk); creating significant media focus and fostering cross-

community interest in nature conservation. The potentially close relationship between the

Australian Sarus Crane and its Philippine relative suggests that should a flagship Sarus Crane

reintroduction to its former range in Luzon be promulgated (J. C. Gonzalez, pers. comm.),

Australian birds could (subject to further genetic investigation) provide the founding stock.

Key Finding 2: Introgression between Australian Sarus Cranes and Brolgas has been

confirmed. Although introgression is still relatively low, it is occurring at a significantly

higher rate than previously realised. Migration of both species between the Atherton

Tablelands and Gulf Plains has been confirmed, using genetic evidence from shed feathers.

As discussed in relation to Key Finding 1 above, the Australian Sarus Crane is significantly

different from the two other extant Sarus Crane subspecies. Although the current rate of

Brolga/Australian Sarus Crane introgression (2.58%) does not yet appear to be putting the

genetic integrity of either Brolgas or Australian Sarus Cranes at risk, as discussed in Chapter

Page 139: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

139

4, Lamichhaney et al. (2016) have demonstrated that introgression at only a single locus may

have important evolutionary consequences. Whether the larger body size of hybrids identified

by Archibald (1981) confers a selective advantage over the parental species cannot yet be

determined but ongoing Brolga/Sarus Crane introgression may differentially favour alleles

which promote adaptation to environmental changes brought about by such stresses as

climate change or agricultural intensification. This suggests that there is sufficient

uncertainty about introgression stimulated by anthropogenic influences to lend weight to the

potential value of establishing a viable captive ‘insurance group’ of Australian Sarus Cranes;

as well as (possibly) a group representative of the northern Brolga population.

The use of shed feathers in this research has confirmed that movement/migration occurs

between the Atherton Tablelands and Gulf Plains and suggests an opportunity for ‘citizen

science’. Subject to the appropriate permitting and training, shed feathers from various

locations in crane breeding and flocking habitat could be collected by members of birding

NGOs and Indigenous rangers, using simple easy-to-follow instructions and minimal

equipment (forceps, plastic gloves and bags, and pre-paid postal bags). The preparation and

genetic analyses of these feathers could, if funded sufficiently, provide extremely valuable

data for strategic decisions on crane conservation.

Key Finding 3: Brolgas and Sarus Cranes are synanthropic on the Atherton Tablelands,

having taken full advantage of combinable crops such as maize and artificial roost sites such

as irrigation storages. This may explain the rapid rise in Australian Sarus Crane numbers

since its formal discovery in the 1960s; as well as the propensity for mixing between Brolgas

and Sarus Cranes, and their consequent introgression.

Page 140: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

140

The sometimes intimate dependence of cranes on agriculture has been described in Chapters

1, 4, 5 and 6 of this thesis. The growth in the area and farmed intensity of combinable crops

(e.g. maize and peanuts) on the Atherton Tablelands appears to have coincided with an

increase in both Brolga and Australian Sarus Crane numbers; with records of crop usage by

Brolgas going back to at least the 1930s. This mirrors the growth in crane numbers in

response to agricultural expansion and intensification in other countries (see Chapter 1).

In relation to the propensity for introgression, a finding of this research is that at sites with

higher Brolga or Australian Sarus Crane counts, the species with lesser numbers in a mixed

flock tended to have proportionally more first year juveniles (see Chapter 4). This puzzling

differential distribution of juveniles of one species amongst flocks of the other, which could

be related to differing species-specific flocking propensity amongst individuals, potentially

increases opportunities for inter-species interactions. As juvenile experience can override

imprinting (Gallagher, 1976), this could significantly affect the number of hybridisation

opportunities created when juveniles are at the critical age for pair-bonding.

Key Finding 4: Brolgas and Sarus Cranes are differentially distributed across the Atherton

Tablelands, with Sarus Cranes favouring intrinsically more fertile volcanic soils and Brolgas

concentrated on poorer metamorphic and granitic soils, closer to roosts. Both favour

cultivated land and post-harvest combinable crops such as maize and peanuts.

The apparent differential distribution of Brolgas and Australian Sarus Cranes on the Atherton

Tablelands, with the latter concentrated on areas of intrinsically higher fertility, may be a

consequence of competition. Although similar in size, the largest Australian Sarus Cranes are

larger than the largest Brolgas (Marchant and Higgins, 1993), so if resources are limiting

there could be active exclusion, especially as Sarus Cranes, being larger, may have higher

Page 141: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

141

nutritional requirements. This could also explain not only the selection of the better soils by

Sarus Cranes but also differences in roost site proximity. While, for neither species, is the

distance travelled from the roost site unusual compared to other cranes (Alonso et al., 2004),

Brolgas may be able to obtain their apparently broader dietary requirements (Sundar et al.,

2018) more readily within a shorter distance of roost sites than Sarus Cranes, which could be

willing to travel further to feed. More speculatively, nutritional limitations may also help

explain the restricted distribution of Sarus Cranes in Australia, where most soils have

intrinsically low fertility (Flannery, 1994), the Atherton Tablelands being a rare exception to

this.

An issue arising from the choice of more fertile agricultural land by Australian Sarus Cranes

is their potential greater vulnerability to agronomic trends. Interviews with Atherton

Tablelands farmers have confirmed a strong intention to shift towards higher value tree crops

(which do not provide foraging opportunities for cranes) on more fertile, higher-yielding

land, likely bringing a decrease in foraging opportunities earlier to Australian Sarus Cranes

than to Brolgas.

Key Finding 5: Global changes in agronomy, harvesting technology and cropping are

combining to reduce food availability for cranes on agricultural land. There is no reason to

expect Australia will be different and interviews with farmers on the Atherton Tablelands

have confirmed this.

Interviews with farmers on the Atherton Tablelands have confirmed that globally increasing

efficiencies in agronomy and harvesting of combinable crops (both resulting in less post-

harvest waste) are being rapidly adopted. Not only will this affect the availability of both

crane species to forage on the Atherton Tablelands, it is also relevant in relation to the

Page 142: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

142

development of new agricultural areas in Cape York Peninsula, as well as the Gulf Plans and

other locations where these new technologies may be deployed as part of cropping

development. Whether this will lead to overall decreases in crane numbers as less dry season

food resource is available or increased raiding of establishing crops (or both), is not yet clear

in Australia but experience from elsewhere (see Chapter 1) suggests that both are plausible

outcomes, for which farmers and conservation managers should already be actively planning.

7.3 Future Scenarios

It is a time of rapid land use change in Australia’s tropical north. To explore what this might

mean for Brolgas and Australian Sarus Cranes, a set of future scenarios and associated

potential land management practices are postulated below. These have been formulated

based on the most frequent issues aired by farmers, conservationists, policy-makers and the

wider community during this research.

Scenario 1: Waste grain reduces from combinable crops, due to improvements in agronomy

and harvesting efficiency – cranes begin to focus on newly-planted crops; negative farmer

attitudes intensify; cranes are persecuted; numbers decline. This could be tested by: (i)

monitoring of crane distribution and numbers on the Atherton Tablelands through late dry

season counts; and (ii) farmer attitude surveys.

Scenario 2: As a result of commercial and policy shifts in irrigation water use towards

higher value agriculture dominated by tree and other high value perennial crops on the

Atherton Tablelands – cranes concentrate on remaining properties; damage is caused to

newly-planted combinable crops, as large numbers of hungry cranes look for food; cranes are

persecuted and their numbers decline. This could be tested by: (i) monitoring of crane

Page 143: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

143

distribution and numbers on the Atherton Tablelands through late dry season counts; and (ii)

farmer attitude surveys.

Scenario 3: Combinable crops expand into new areas such as CYP and the Gulf Plains -

cranes move to newly-cultivated areas; crane populations increase but with less waste grain

the potential for crop damage and persecution increases. This could be tested by: (i)

monitoring of crane distribution and numbers in late dry season counts; and (ii) shed feather

sampling at new sites and the Atherton Tablelands (confirming trajectories of crane

movements).

Scenario 4: Crane-based tourism – cranes are deliberately fed; farmers are compensated by

tourism operators/local government; crane numbers rise: potentially more hybridisation

occurs due to closer sympatry of juveniles. This could be tested by: (i) monitoring of

distribution and numbers in late dry season counts; and (ii) genetic sampling of shed feathers.

Scenario 5: Changes in cattle grazing regimes alter breeding area suitability – technological

and husbandry changes alter the way in which cattle use crane breeding habitat in the Gulf

Plains and CYP. This could be tested by monitoring of crane distribution and numbers in late

breeding season counts.

Scenario 6: De-stocking of significant areas of crane breeding habitat occurs following

purchase of cattle properties by nature conservation NGOs (this is already happening

extensively in northern CYP) and the widespread uptake of plant-based meat products - cattle

are removed at a large scale; leading to changes in land management practices and the

structure of vegetation; with as yet unknown outcomes for cranes. This could be tested by:

monitoring of crane distribution and numbers in late breeding season counts.

Page 144: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

144

Scenario 7: Significant increase in mining adjacent to key breeding season habitat – leading

to greater disturbance adjacent to significant areas of crane breeding habitat in CYP and

elsewhere. This could be tested by: monitoring of crane distribution and numbers in late

breeding season counts.

Scenario 8: The Philippine government decides to reintroduce Sarus Cranes as a flagship

agri-conservation project and seeks suitable founding stock and expertise. This could be

tested/investigated by: (i) further genetic analyses of Philippine Sarus Crane museum

specimens; (ii) surveys of suitability of sites for potential reintroduction to Luzon; and (iii)

surveys of local Luzon communities to test their preparedness to accept a reintroduction

project.

Although these scenarios are speculative, 1, 2, 3 and 4 are potentially critical in the short

term, with Scenario 4 potentially offering a solution for 1, 2 and 3; retaining cranes within the

Atherton Tablelands KBA and providing financial benefits to facilitate ongoing crane

management. For scenario 4 to have the best chance of working, early, comprehensive and

durable engagement between the Cairns-based regional tourism industry, landholders, local

and State government is essential, assisted by community-based initiates such as the Atherton

Tablelands Crane Week (www.craneweek.org) and Birdlife Australia’s annual Crane Count.

Scenarios 5 and 6 are generated by different but related objectives, bringing about significant

ecological change within crane habitat. Some observers (the author and G. Archibald pers.

obs.) believe that there may be synergistic relationships between cattle grazing and crane

foraging, roosting and breeding habitat, making this important to investigate. The

development of Scenario 8 would require a formal approach from a Philippine agency and

facilitation by an organisation such as the ICF.

Page 145: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

145

The debate around crane conservation in agricultural landscapes is becoming more intense.

The ICF have seen this and responded by bringing available knowledge together in its

landmark report ‘Cranes and Agriculture: A Global Guide for Sharing the Landscape’ (Austin

et al., 2018). This does a comprehensive job of bringing together research from almost all

crane states. However, due to the dearth of Australian research here it necessarily only makes

relatively small mention of our continent. A key role for this thesis is therefore to stimulate

research that can fill the gaps in relation to the interface between agriculture, Brolgas and

Australian Sarus Cranes.

Page 146: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

146

CHAPTER 8 - CONCLUSIONS AND RECOMMENDATIONS

8.1 Conclusions

Based on a multidisciplinary perspective to crane conservation recommended by the ICF, this

thesis has reached four conclusions in areas of key importance to crane conservation:

1. Confirmation of the Australian Sarus Crane as a distinct subspecies. The

Australian Sarus Crane has unique genetic traits and is distinct from the two clinal

Asian Sarus Crane subspecies; underpinning its retention as a separate conservation

management unit. Although based on only one individual and subject to further

investigation, a closer genetic relationship is likely to exist between Australian and

Philippine Sarus Cranes; re-confirming the distinct status of the extinct Philippine

subspecies.

2. Confirmation of introgression between Brolgas and Australian Sarus Cranes:

This research has confirmed introgression between Brolgas and Australian Sarus

Cranes, and although currently of relatively low incidence (2.58%), research in other

taxa suggests that this may be evolutionarily significant. Migration between flocking

areas on the Atherton Tablelands and breeding areas in the Gulf Plains has also been

confirmed.

3. Definition of the relationship between agriculture and cranes on the Atherton

Tablelands: This research has established the differential distribution of cranes on the

Atherton Tablelands and that both harvested maize and cultivated land are critical to

crane usage of the Atherton Tablelands. It has also established that Australian Sarus

Cranes make differential use of intrinsically more fertile land than Brolgas, probably

as a result of competitive exclusion; and that the proximity of roosts to foraging areas

is of greater significance for Brolgas, possibly related to a broader dietary base.

Page 147: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

147

4. Determination of the attitudes of farmers towards cranes: Interviews with farmers

have confirmed visual observations that maize, cultivated land and other combinable

crops are of significant importance to cranes. The interviews also confirmed that

farmers intend to radically reduce waste grain at harvest through improved efficiency,

as well as move to higher value perennial crops which provide no foraging

opportunities for cranes. Both changes could potentially lead to higher levels of

damage to remaining areas of early stage combinable crops on the Atherton

Tablelands, similarly in newly-developed cropping areas. Farmer interest in crane-

related tourism has also been confirmed.

8.2 Recommendations

Cranes are charismatic, long-lived and readily observable. They are used to epitomise prized

societal values in cultures ranging from the Australian Aboriginal ‘Buralga’ to Indian Hindu

‘Ramayana’ myths, both of which continue to help protect cranes in contemporary

landscapes. Cranes also provide globally-recognised commercial emblems, such as those

displayed on the tailplanes of Japanese and German airlines, on Chinese credit cards and

Australian wines. This combination of charisma, cultural significance and commercial

recognition makes them ideal ambassadors for conservation in agricultural and pastoral

landscapes.

The cranes and landscape of the Atherton Tablelands could provide a fulcrum around which

to reconnect farmers and communities with nature. Cranes readily lend themselves to

outreach programmes designed to foster greater awareness, understanding, and appreciation

of the landscapes that farmers share with nature, as has been done in England with the Great

Crane Project. Crane-based tourism and community-based events could further reinforce

Page 148: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

148

these relationships, providing opportunities to demonstrate the value of nature to the broader

community.

Based on the research findings and issues explored in this thesis, to ensure that cranes can to

continue to prosper in Australia’s rapidly-changing agricultural and pastoral landscapes I

make the following recommendations; capable of implementation either as individual

measures or as components of multi-measure integrated programmes:

Recommendation 1: Framing and adoption of policies and regulations by all levels of

government, designed to sustain biodiversity within agricultural and pastoral landscapes;

including formal protection of wetlands and other crane habitat.

Recommendation2: Fostering of longitudinal working relationships between farmers,

conservation practitioners, agricultural experts, government and other key stakeholders to

explore effective mitigation measures for farming/crane conflicts.

Recommendation 3: Development of a farmer-friendly crane management handbook (in an

appropriate range of media), setting-out practical measures to reduce conflicts between cranes

and farmers, and build a long-term relationship; including damage avoidance and protection

measures, habitat creation techniques and supplementary feeding opportunities.

Recommendation 4: Fostering of research which identifies practices that can enhance both

agricultural sustainability and biodiversity, whilst meeting demands for increased agricultural

output and efficiencies. For example, trialling crane-repellent seed coatings.

Recommendation 5: Fostering and development of partnerships between tourism operators,

farmers and local government (with the assistance of State government) to identify

opportunities for crane-based tourism.

Page 149: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

149

Recommendation 6: Investigation of the relationship between cattle grazing and cranes,

especially in wet season breeding and dry season foraging habitat, as well as in the vicinity of

roost sites.

Recommendation 7: Investigation of the feasibility of establishing a genetically-balanced

captive group of Australian Sarus Cranes at a well-resourced institution, with appropriate

avian management skills; safeguarding birds free from admixture of Brolga genes (NB – It

may also be appropriate to investigate the establishment of a representative captive group

from the northern population of the Brolga).

I also recommend that further work be undertaken on the genetic relationships between the

Philippine Sarus Crane and other subspecies. Should this confirm that the Philippine and

Australian subspecies are closely-related, and subject to agreement by Philippine agencies, a

flagship reintroduction project similar to the UK’s Great Crane Project may be feasible and

would benefit from the outputs of all of the above seven recommendations above.

8.3 A Final Word

The Brolga is Queensland’s official bird emblem and supports the right hand side of its coat

of arms, a sheaf of grain occupying the bottom left hand quadrant. With agricultural

intensification inevitable, we must ensure cranes are not put in jeopardy of disappearing from

either Queensland’s coat of arms or its rural landscapes by insensitive land use change.

Fig. 8.1 Queensland Coat of Arms (with author endorsements).

Page 150: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

150

ACKNOWLEDGEMENT OF ASSISTANCE

Supervisory team

• Professor Stephen Garnett, Research Institute for the Environment and Livelihoods (RIEL),

CDU (Principal Supervisor).

• Dr George Archibald, International Crane Foundation.

• Dr Ian Leiper, RIEL, CDU (from 2017).

• Professor Michael Lawes, formerly at RIEL, CDU (until 2017).

Financial support

• Charles Darwin University (CDU) Postgraduate Research Scholarship (2013-16).

• North Queensland Wildlife Trust.

• Norman Wettenhall Foundation.

• CDU travel and accommodation support grants.

• CDU supplementary research funding grants.

• BirdLife Australia.

Technical assistance

• Genetic analyses: Dr Martin Haase of the University of Greifswald.

• Statistical design and analyses: Dr Don Franklin of RIEL, CDU and Ecological

Communications.

• Spatial analysis: Dr Ian Leiper of RIEL, CDU.

• Veterinary clearances: Dr Annabelle Olsson of Boongarry Veterinary Clinic.

• Additional genetic samples: Nuchjaree Purchkoon and Dr Boripat Siriaroonrat of Korat

Zoo, Thailand; Robert van Zalinge of RIEL, CDU; Professor Michael Wink of the University

of Heidelberg, Germany; Tin Nwe Latt of Mahidol University, Thailand; Dr Chris Milensky

Page 151: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

151

of the Division of Birds at the Smithsonian Institution, Washington; and Emily Imhoff of

Cincinnati Museum.

Editorial support

• Manuscript editing: Supervisory team.

Fieldwork

• Volunteers: Harry Nevard, Bronwen Nevard, Dr Ruairidh Nevard, Baroness Vera von

Schwertzell zu Willingshausen, Baron Dominic von Schwertzell zu Willingshausen, Stella

Freund, Jürgen Freund, Dr John Grant, Trevor Adil and Jess Harris.

• Technical assistance: Dr David Westcott, Adam McKeown and Dr Inka Veltheim.

Laboratory work

• Technical support: Silke Fregin, University of Greifswald.

Page 152: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

152

APPENDIX – PERMITS, APPROVALS AND AUTHORITIES

Permit number: WISP13984714 – QP&WS Scientific Purposes Permit.

Permit number: 545 – ABBBS Banding Project Permit.

Permit number: ACLTE47K – ACBPS Export Permit

Approval number: A13019 – CDU Ethics Approval for Animal Based Research.

Approval number: H15003 – CDU Ethics Approval for Human Based Research.

Authority number: 000002/FD-2011 – CITES approval.

Authority number: PWS2014-AU-001240 – CITES approval.

Authority number: PWS2015-AU-000119 – CITES approval.

Authority number: KH1108 – CITES approval.

Page 153: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

153

BIBLIOGRAPHY

Akiyama, T., Nishida, C., Momose, K., Onuma, M., Takami, K. and Masuda, R. (2017).

Gene duplication and concerted evolution of mitochondrial DNA in crane species. Mol

Phylogen Evol, 106: 158-163.

Alexandratos, N. Bruinsma, J. (2012). World agriculture towards 2030/2050: the 2012

andrevision. ESA Working Paper No. 12-03. Food and Agriculture Organization, Rome.

Allendorf, F.W., Leary, R.F., Spruell, P. and Wenburg, J.K. (2001) The problems with

hybrids: setting conservation guidelines. Trends in Ecology and Evolution, 16: 613–622.

Alonso, J.C., Bautista, L.M. and Alonso, J.A. (2004). Family‐based territoriality vs flocking

in wintering common cranes Grus grus. Journal of Avian Biology, 35(5): 434-444.

Amano, T., Ushiyama, K. and Higuchi, H. (2008). Methods of predicting risks of wheat

damage by white-fronted geese Journal of Wildlife Management, 72: 1845-1852.

Anderson, M.J., Gorley, R.N. and Clarke, K.R. (2008). PERMANOVA+ for PRIMER: Guide

to Software and Statistical Methods. PRIMER-E: Plymouth, UK. 214 pp.

Andryuschenko, Y. A. (2011). The Demoiselle Crane on Agricultural Lands in the Ukraine.

In: (Eds E. Ilyashenko, S. Winter) Cranes of Eurasia (biology, distribution, migrations,

management). Issue 4. Moscow: Crane Working Group of Eurasia. pp. 476–483.

Anteau, M.J., Sherey, M.H. and Bishop. A.A. (2011). Location and agricultural practices

influence spring use of harvested cornfields by cranes and geese in Nebraska. Journal of

Wildlife Management, 75: 1004–1011.

Page 154: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

154

Archibald, G.W. (1975). Introducing the ‘Sarolga’. Unpublished note; International Crane

Foundation, Wisconsin (ICF library ref: 08-028).

Archibald, G.W. (1976). The Unison Call of Cranes as a Useful Taxonomic Tool. PhD

Thesis, Faculty of Cornell University Graduate School. December 1976.

Archibald, G.W. (1976) Crane taxonomy as revealed by the unison call. In Proceedings of the

International Crane Workshop (ed J.C. Lewis), pp. 225-251. Oklahoma State University.

Archibald, G.W. (1981) Introducing the Sarolga. In Crane Research around the World.

Proceedings of the International Crane Symposium at Sapporo, Japan in 1980 and Papers

from the World Working Group on Cranes. (eds J.C. Lewis, H. Masatomi), pp. 213-215.

International Council for Bird Preservation.

Archibald, G.W. and Swengel, S.R. (1987). Comparative ecology and behaviour of eastern

sarus cranes and brolgas in Australia. In ‘Proceedings of the 1985 Crane Workshop.

International Crane Foundation’. (Ed. J. C. Lewis.) pp. 107–116. (Platte River Whooping

Crane Habitat Maintenance Trust, and US Fish and Wildlife Service: Grand Island, NB.)

Archibald, G.W., Sundar, K. S. G. and Barzen, J. (2003). A review of the three subspecies of

Sarus Cranes Grus antigone. Journal of the Ecological Society (India), 16:5-15.

Arnol, J. D., White, D. M. and Hastings, I. (1984). Management of the brolga (Grus

rubicundus) in Victoria. Department of Conservation, Forests and Lands, Fisheries and

Wildlife Service.

Arnold, M.L. (1997) Natural hybridization and evolution. Oxford University Press.

Page 155: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

155

Aryal, A., Shrestha, T.K., Sen, D.S., Upreti, B. and Gautam, N. (2009). Conservation regime

and local population ecology of Sarus Crane (Grus antigone antigone) in west-central region

of Nepal. Journal of Wetlands Ecology, 3:1–11.

Attwood, S.W. and Johnston, D.A. (2001). Nucleotide sequence differences reveal genetic

variation in Neotricula aperta (Gastropoda: Pomatiopsidae), the snail host of schistosomiasis

in the Lower Mekong Basin. Biological Journal of the Linnean Society, 73:23-41.

Austin, J.E. (2012). Conflicts between Sandhill Cranes and farmers in the Western United

states: evolving issues and solutions. In: Harris J, (ed.). Proceedings of the Cranes,

Agriculture, and Climate Change Workshop, 28 May-3 June 2010, 132–140 International

Crane Foundation.

Austin, J.E., Morrison, K.L. and Harris, J.T., eds. (2018). Cranes and Agriculture: A Global

Guide for Sharing the Landscape: International Crane Foundation. Baraboo, Wisconsin,

USA.

Australian Broadcasting Corporation (2013). www.abc.net.au/news/2013-08-14/queensland-

protected...brolgas-poison/4887360

Avery, D.T. (1997). Saving the planet with pesticides, biotechnology and European farm

reform, Global Food Issues, Vol 1. Hudson Institute, Indianapolis.

Baker, A.N., Smith, A.N.H. and Pichler, F.B. (2002). Geographical variation in Hector’s

dolphin: recognition of new subspecies of Cephalorhynchus hectori. ,Journal of the Royal

Society of New Zealand, 32:713–727.

Page 156: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

156

Balmford, B., Green, R.E., Onial, M, Phalan, B. and Balmford, A. (2019). How imperfect can

land-sparing be before land-sharing is more favourable for wild species? Journal of Applied,

Ecology, 59:73-84.

Balmford, A., Amano, T., Bartlett, H., Chadwick, D., Collins, A., Edwards, D., Field, R.,

Garnsworthy, P., Green, R., Smith, P., Waters, H., Whitmore, A., Broom, D., Chara, J.,

Finch, T., Garnett, E., Gathorne-Hardy, A., Hernandez-Medrano, J., Herrero, M., Hua,

F., Latawiec, A., Misselbrook, T., Phalan, B., Simmons, B., Takahashi, T., Vause, J., zu

Ermgassen, E. and Eisner, R. (2018). The environmental costs and benefits of high-yield

farming. Nature Sustainability, 1: 477–485.

Balmford, A., Beresford, J., Green, J., Naidoo, R., Walpole, M. and Manica, A. (2009). A

global perspective on trends in nature-based tourism. PLoS Biology, 7(6): e1000144.

Ban, N.C., Mills, M., Tam, J., Hicks, C.C., Klain, S., Stoeckl, N., Bottrill, M.C., Levine, J.,

Pressey, R.L., Satterfield, T. and Chan, K.M. (2013). A social-ecological approach to

conservation planning: embedding social considerations. Frontiers in Ecology and the

Environment, 11:194–202.

Barrett, G., Silcocks, A., Barry, S. and Poulter, R. (2003) The new atlas of Australian birds.

Royal Australasian Ornithological Union, Melbourne.

Barton, N.H. and Slatkin, M. (1986) A quasi-equilibrium theory of the distribution of rare

alleles in a subdivided population. Heredity, 56:409-415.

Barzen, J. and Ballinger, K. (2017) Sandhill and Whooping Cranes. Wildlife Damage

Management Technical Series. US Department of Agriculture, Animal and Plant Inspection

Service.

Page 157: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

157

Barzen, J., Lacy, A. and Harris, J. (2012). Developing solutions to Sandhill Crane damage to

seedling corn in the upper Midwest, USA. In: Harris J, editor. Proceedings of the Cranes,

Agriculture, and Climate Change Workshop at Muraviovka Park, 28 May–3 June, 2010.

International Crane Foundation, Baraboo, Wisconsin, USA.

Benton, T (2016). The many faces of food security. International Affairs, 6:1505-15

Bignal, E.M. and McCracken, D.I. (2000). The nature conservation value of European

traditional farming systems. Environmental Review, 8:149–171.

Bird, M.I., Taylor, D. and Hunt, C. (2005). Palaeoenvironments of insular Southeast Asia

during the Last Glacial Period: a savanna corridor in Sundaland? Quaternary Science

Reviews, 24(20-21):2228-2242.

BirdLife International (2018a) State of the world’s birds: taking the pulse of the planet.

Cambridge, UK: BirdLife International.

BirdLife International (2018b). Important Bird Areas factsheet: Atherton Tablelands.

Downloaded from: http://www.birdlife.org [accessed 15 February 2018]

BirdLife International. (2018c). Antigone antigone. The IUCN Red List of Threatened

Species 2016: e.T22692064A93335364. http://dx.doi.org/10.2305/IUCN.UK.2016-

3.RLTS.T22692064A93335364.en. [accessed 27 September 2018].

Blackwell. B. F., Helon, D. A. and Dolbeer, R. A. (2002). Repelling sandhill cranes from

corn: whole-kernel experiments with captive birds. Crop Protection, 20:65-68.

Blanford, W. (1895) Grus (antigone) sharpei. Bulletin of the British Ornithologists’ Club, 5:7.

Page 158: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

158

Blyth, E. and Tegetmeier, W. G. (1881) The Natural History of the Cranes. Horace Cox,

London.

Bomford, M. and Sinclair, R., (2002). Australian research on bird pests: impact, management

and future directions. Emu, 102(1):29-45.

Borad, C.K. and Mukherjee, A. (2002). Cranes in agricultural ecosystem. China Crane News,

6 (supplement):24–25.

Bouffard, S.H., Cornely, J.E. and Goroshko. O.A. (2005). Crop depredation by cranes at

Daursky State Biosphere Reserve, Siberia. Proceedings of the North American Crane

Workshop, 9:145–150. International Crane Foundation.

Brownsmith, C. B. (1978). Sarus Cranes dominant over Brolgas in captivity. Emu, 78:98.

Bull, J.W., Lloyd, S.P. and Strange, N. (2017). Implementation gap between the theory and

practice of biodiversity offset multipliers. Conservation Letters, 10:656–669.

Burke, J.M. and Arnold M.L. (2001) Genetics and the fitness of hybrids. Annual Review of

Genetics, 35:31-52.

Calinski, R.B. and Harabasz, J. (1974) A dendrite method for cluster analysis.

Communications in Statistics, 3:1-27.

Chatto, R. (2006). The distribution and status of waterbirds around the coast and coastal

wetlands of the Northern Territory. Technical report no. 76, Parks and Wildlife Commission

of the Northern Territory, Northern Territory, Australia.

Page 159: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

159

Clarke, R.H., Gordon, I.R. and Clarke, M.F., (2001). Intraspecific phenotypic variability in

the black-eared miner (Manorina melanotis); human-facilitated introgression and the

consequences for an endangered taxon. Biological Conservation, 99(2):145-155.

Clement, M., Snell, Q., Walke, P., Posada, D. and Crandall, K. (2002). TCS: estimating gene

genealogies. Proc. 16th International Parallel Distribution Process Symposium, 2:184.

Collins, P. (1995). The birds of Broome. Broome Bird Observatory, Broome, Western

Australia.

Darriba, D., Taboada, G.L., Doallo, R. and Posada, D. (2012). jModelTest 2: more models,

new heuristics and parallel computing. Nature Methods, 9(8):772.

Dayrat, B. (2005). Towards Integrative Taxonomy. Biol. J. Linnean Soc, 83(3): 407-415

Deane, P.M. (1979). The First Industrial Revolution. Cambridge University Press, New York

Dell Inc. (1984-2016). Statistica 13.2. Dell Inc.: Round Rock, Texas.

del Hoyo, J. and Collar, N. J. (2014). HBW and BirdLife International Illustrated Checklist of

Birds of the World (Volume 1 Non-Passerines). Lynx Edicions, in association with BirdLife

International. Barcelona.

Dessauer, H.C., Gee, G.F. and Rogers, J.S. (1992). Allozyme evidence for crane systematics

and polymorphisms within populations of Sandhill, Sarus, Siberian, and Whooping cranes.

Molecular Phylogenetics and Evolution, 1:279-288.

Department of Agriculture and Fisheries (2015). Tablelands Agricultural Profile. Department

of Agriculture and Fisheries, Mareeba, Queensland.

Page 160: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

160

Department of the Environment and Energy (2012) Interim Biogeographic Regionalisation

for Australia, Version 7 and Sub-regions. Commonwealth of Australia, Canberra.

Department of Environment and Energy (2018). EPBC Act List of Threatened Fauna.

http://www.environment.gov.au/cgi-bin/sprat/public/publicthreatenedlist.pl#birds_extinct

[Accessed 9/27/2018].

Department of Transport and Main Roads (2016) State controlled roads – Queensland. State

of Queensland (Department of Transport and Main Roads).

de Quieroz, K. (2007) Species concepts and species delimitation. Systematic Biology 56:879–

886.

Dessauer, H.C., Gee, G.F. and Rogers, J.S. (1992). Allozyme evidence for crane systematics

and polymorphisms within populations of Sandhill, Sarus, Siberian, and Whooping cranes.

Molecular Phylogenetics and Evolution, 1(4): 279-288.

Devitt, T.J., Baird, S.J.E. and Moritz, C. (2011). Asymmetric reproductive isolation between

terminal forms of the salamander ring species Ensatina eschscholtzii revealed by fine-scale

genetic analysis of a hybrid zone. BMC Evolution Biology, 11: 245.

Donnellan, S.C., Armstrong, J., Pickett, M., Milne, T., Baulderstone, J., Hollfelder, T. and

Bertozzi, T. (2009). Systematic and conservation implications of mitochondrial DNA

diversity in emu-wrens, Stipiturus (Aves: Maluridae). Emu, 109(2): 143-152.

Dowling, T.E. and Secor, C.L. (1997) The role of hybridization and introgression in the

diversification of animals. Annual Review of Ecology and Systematics, 28:593-619.

Page 161: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

161

Earl D.A. and von Holdt B.M. (2012). STRUCTURE HARVESTER: a website and program

for visualizing STRUCTURE output and implementing the Evanno method. Conservation

Genetics Resources, 4:359-361.

Eisner, R., Waters, H.S., Green, R., Edwards, D., Field, R., Fenuik, C. and Balmford, A.

(2017). Does higher-yielding agriculture mean more environmental harm? Aspects of

Applied Biology, 136:63–70.

ESRI 2016. ArcGIS Desktop: Release 10.4.1. Redlands, CA: Environmental Systems

Research Institute.

Evanno, G., Regnaut, S. and Goudet, J. (2005). Detecting the number of clusters of

individuals using the software STRUCTURE: a simulation study. Molecular Ecology,

14:2611-2620

Fahrig, L., Baudry, J., Brotons, L., Burel, F.G., Crist, T.O., Fuller, R.J., Sirami, C.,

Siriwardena, G.M. and Martin, J.L., (2011). Functional landscape heterogeneity and animal

biodiversity in agricultural landscapes. Ecology Letters, 14 (2):101-112.

Falush, D., Stephens, M. and Pritchard, J.K. (2007). Inference of population structure using

multilocus genotype data: dominant markers and null alleles. Molecular Ecology Notes,

7:574-578.

Fishbein, M. and Ajzen, I. (2011). ‘Predicting and changing behaviour: the reasoned action

approach’. (Taylor and Francis: New York.)

Fitzpatrick, B.M. (2012) Estimating ancestry and heterozygosity of hybrids using molecular

markers. BMC Evolutionary Biology, 12:131.

Page 162: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

162

Flannery, T.F. (1994). The Future Eaters: An ecological history of the Australasian lands and

people. Reed, Melbourne.

Frankham, R., Ballou, J.D., Dudash, M.R., Eldridge, M.D., Fenster, C.B., Lacy, R.C.,

Mendelson III, J.R., Porton, I.J., Ralls, K. and Ryder, O.A. (2012). Implications of different

species concepts for conserving biodiversity. Biological Conservation, 153:25-31.

Franklin, D.C. (2008). Report 9: The waterbirds of Australian tropical rivers and wetlands.

In: Lukacs, G.P. and Finlayson, C.M. (Eds.) A Compendium of Ecological Information on

Australia’s Northern Tropical Rivers. Sub-project 1 of Australia’s Tropical Rivers - an

Integrated Data Assessment and Analyses (DET18). A report to Land and Water Australia.

National Centre for Tropical Wetland Research, Townsville, Queensland.

Frost, G. (2015). Rare coastal sage scrub habitat provides a home for threatened gnatcatcher

and many other species. http://ca.audubon.org/news/rare-coastal-sage-scrub-habitat-provides-

home-threatened-gnatcatcher-and-many-other-species.

Gallagher, J. (1976). Sexual imprinting: effects of various regimens of social experience on

mate preference in Japanese Quail (Coturnix coturnix japonica). Behaviour, 57:91-115.

Galluzzi, G., Van Duijvendijk, C., Collette, L., Azzu, N. and Hodgkin, T. (2011).

Biodiversity for food and agriculture. Contributing to food security and sustainability in a

changing world. PAR platform. Food and Agriculture Organisation, Rome.

Garnett, S. T., Woinarski, J. C. Z., Crowley, G. M. and Kutt, A. S. (2010). Biodiversity

conservation in Australian tropical rangelands. In ‘Can rangelands be wildlands?: wildlife

and livestock in semi-arid ecosystems’. (Eds J. du Toit, R. Kock, and J. Deutsch.) 191–234.

Blackwell Publishing, Oxford.

Page 163: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

163

Garnett, S.T and Christidis, L, (2007). Implications of changing species definitions for

conservation purposes. Bird Conservation International, 17:187-195.

Garnett, S.T. and Christidis, L. (2017). Taxonomy anarchy hampers conservation. Nature,

546:25-27.

Garnett, S.T. and Crowley, G.M. (2000) The Action Plan for Australian Birds 2000.

Environment Australia, Canberra.

Garnett, S.T. and Crowley, G.M. (2008). The history of threatened birds in Australia and its

offshore islands. Pp. 353-401 in Davis, W.E., Recher, H.F. and Boles, W.E. (eds.)

Contributions to the History of Australasian Ornithology. Nuttall Ornithological Club,

Harvard, Massachusetts.

Garnett, S.T., Franklin, D.C., Ehmke, G., VanDerWal, J.J., Hodgson, L., Pavey, C., Reside,

A.E., Welbergen, J.A., Butchart, S.H.M., Perkins, G.C. and Williams, S.E. (2013) Climate

change adaptation strategies for Australian birds. National Climate Change Adaptation

Research Facility, Gold Coast

Garnett, S.T., Kleinschmidt, M., Jackson, M.V., Zander, K.K. and Murphy, S.A. (2016).

Social landscape of the Night Parrot in western Queensland, Australia. Pacific Conservation

Biology, 22:360-366.

Gavrilets, S. (2003). Models of speciation: what have we learned in 40 years? Evolution,

57:2197-2215.

Gill HB. The first record of the Sarus Crane in Australia. Emu. 1967;69: 49-52.

Page 164: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

164

Gippoliti S, Cotterill FP, Zinner D, Groves CP. Impacts of taxonomic inertia for the

conservation of African ungulate diversity: an overview. Biological Reviews. 2018;93(1):

115-130.

Goudet J. FSTAT (version 1.2): a computer program to calculate F-statistics. Journal of

Heredity. 1995;86: 485-486.

Getis, A. and Ord, J.K. (1992). The analysis of spatial association by use of distance

statistics. Geographical Analysis, 24(3):189-206.

Gill, H.B. (1967). The first record of the Sarus Crane in Australia. Emu, 69: 49-52.

Gippoliti, S., Cotterill, F.P., Zinner, D. and Groves, C.P. (2018). Impacts of taxonomic inertia

for the conservation of African ungulate diversity: an overview. Biological Reviews, 93(1):

115-130.

Goudet, J. (1995). FSTAT (version 1.2): a computer program to calculate F-statistics. Journal

of Heredity, 86:485-486.

Grant, J.D.A. (2005). Recruitment rate of Sarus Cranes (Grus antigone) in northern

Queensland. Emu, 105:311-315.

Grant, P.R. and Grant, B.R. (2008). How and Why Species Multiply – The Radiation of

Darwin’s Finches. Princeton University Press, Princeton and Oxford.

Grant, P.R. and Grant, B.R. (2016). Introgressive hybridization and natural selection in

Darwin’s finches. Biological Journal of the Linnean Society, 117:812-822.

Page 165: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

165

Gray, A.P. (1958). Bird Hybrids. Commonwealth Agricultural Bureau, Farnham Royal,

Bucks, England.

Greiner, R. (2015). Motivations and attitudes influence farmers' willingness to participate in

biodiversity conservation contracts. Agricultural Systems, 137:154–165.

Guerrero, A.M., McKenna, R., Woinarski, J.C.Z., Pannell, D.J, Wilson, K.A. and Garnett,

S.T. (2017). Threatened species recovery planning in Australia: Learning from two case

studies. National Environmental Science Programme, Threatened Species Recovery Hub.

Haase, M. and Ilyashenko, V. (2012). A Glimpse on Mitochondrial Differentiation among

four Currently Recognized Subspecies of the Common Crane Grus grus. Ardeola, 59(1):131-

135.

Hachisuka, M. (1941). Further contributions to the ornithology of the Philippine Island. Tori,

11:61–89.

Hall, T.A. (1999). BioEdit: a user-friendly biological sequence alignment editor and analysis

program for Windows 95/ 98/NT. Nucleic Acids Symposium Series, 41:95–98.

Hanspach, J.A., Collier, D.J., French, N., Dorresteijn, I., Schultner, J. and Fischer, J. (2017).

From trade‐offs to synergies in food security and biodiversity conservation. Frontiers in

Ecology and the Environment. doi.org/10.1002/fee.1632.

Harding, C. (2001) Use of remote sensing and geographic information systems to predict

suitable breeding habitat for the Brolga Grus rubicundus in south-west Victoria. Diss. Centre

for Environmental Management, University of Ballarat, Victoria, Australia.

Page 166: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

166

Harris, J. and Mirande, C. (2013). A global overview of cranes: status, threats and

conservation priorities. Chinese Birds 4:189–209.

Harris, J. (2010). Introduction: Cranes, Agriculture and Climate Change pp1-14. in “Cranes,

Agriculture, and Climate Change. Proceedings of a workshop organized by the International

Crane Foundation and Muraviovka Park for Sustainable Land Use”. Ed. J.Harris.

International Crane Foundation, Baraboo.

Hasegawa, O., Ishibashi, Y. and Abe, S. (2000). Isolation and characterisation of

microsatellite loci in the red-crowned crane Grus japonensis. Molecular Ecology, 9:1677-

1678.

Heinrich, J.W. and Craven, S.R (1992). The economic-impact of Canada Geese at the

Horicon Marsh, Wisconsin. Wildlife Society Bulletin, 20:364-371.

Hedrick, P.W. (2011). Genetics of Populations. Jones and Bartlett Publishers, Sudbury, MA,

USA.

Herring, M.W. (2005). Threatened species and farming. Brolga: management of breeding

wetlands in northern Victoria. Arthur Rylah Institute for Ecological Research, Heidelberg,

Australia.

Herring, M.W. (2007) Brolga breeding habitat: managing wetlands on your farm. Murray

Catchment Management Authority, Australia.

Horwich, R.H. (1996) Imprinting, attachment, and behavioral development in cranes. In

Cranes: Their biology, husbandry and conservation (eds Ellis, D.H., Gee, G.F. and Mirande,

C.M.), pp. 117-22. Washington, D.C.: Department of the Interior, National Biological

Service/International Crane Foundation.

Page 167: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

167

Hou, X., Xu, P., Lin, Z., d'Urban-Jackson, J., Dixon, A., Bold, B., Xu, J., Zhan, X. and Hou,

A. (2018). An integrated tool for microsatellite isolation and validation from the reference

genome and their application in the study of breeding turnover in an endangered avian

population. Integr Zool. 2018 Jan 9. doi: 10.1111/1749-4877.12305

Hughes, M. R. and Blackman, J. G. (1973). Cation content of salt gland secretion and tears in

the Brolga Grus rubicundus (Perry) (Aves: Gruidae). Australian Journal of Zoology, 21:515-

518.

Ilyashenko, E. and King, S. (2018). Crane responses to changes in agriculture. In Austin, J.E.,

Morrison K. and Harris, J.T., eds. Cranes and Agriculture: A Global Guide for Sharing the

Landscape. Baraboo, International Crane Foundation, Wisconsin, USA.

Immelmann, K. (1972). Sexual and other long-term aspects of imprinting in birds and other

species. Advances in the Study of Behaviour, 4:147-174.

Jesmer, B.R., Merkle, J.A., Goheen, J.R., Aikens, E.O., Beck, J.L., Courtemanch, A.B.,

Hurley, M.N., McWhirter, D.E., Miyasaki, H.M., Monteith, K.L. and Kauffman, M.J. (2018).

Is ungulate migration culturally transmitted? Evidence of social learning from translocated

animals. Science, 361:1023-1025.

Johnsgard, P. A. (1983) Cranes of the World. Indiana University Press, Bloomington

Jones, K.L., Barzen, J.A. and Ashley, M.V. (2005) Geographical partitioning of

microsatellite variation in the Sarus Crane. Animal Conservation, 8:1-8.

Joseph, L., Bishop, K.D., Wilson, A.C., Edwards, S.V, Iova B, and Campbell, C.D. (2019).

Review of evolutionary research on birds of the New Guinean savannas and closely

Page 168: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

168

associated habitats of riparian rainforests, mangroves and grasslands. Emu. DOI:

10.1080/01584197.2019.1615844

Kaczensky, P., Guillaume, C., Huber, D., Andrén, H. and Linnell, J. (2012). Status,

management and distribution of large carnivores – bear, lynx, wolf and wolverine in Europe.

European Commission

Kalinowski, S.T., Taper, M.L. and Marshall, T.C. (2007). Revising how the computer

program CERVUS accommodates genotyping error increases success in paternity

assignment. Molecular Ecology, 16:1099-1106.

Kamp, J., Urazaliev, R., Balmford, A., Donald, P.F., Green, R.E., Lamb, A.J. and Phalan, B.

(2015). Agricultural development and the conservation of avian biodiversity on the Eurasian

steppes: a comparison of land-sparing and land-sharing approaches. Journal of Applied

Ecology, 52:1578–1587.

Katoh, K., Rozewick, J. and Yamada, K.D. (2017). MAFFT online service: multiple

sequence alignment, interactive sequence choice and visualization. Briefings in

Bioinformatics: https://doi.org/10.1093/bib/bbx108.

Keyghobadi, N. (2007). The genetic implications of habitat fragmentation for animals.

Canadian Journal of Zoology, 85:1049-1064.

Kingsford, R.T., Porter, J.L. and Halse, S.A. (2012). National waterbird assessment.

Waterlines report, National Water Commission, Canberra.

Page 169: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

169

Koch, E.L., Neiber, M., Walther, F. and Hausdorf, B. (2017). High gene flow despite

opposite chirality in hybrid zones between enantiomorphic door-snails. Molecular Ecology,

26:3998-4012.

Kopelman, N.M., Mayzel, J., Jakobsson, M., Rosenberg, N.A. and Mayrose, I. (2015).

CLUMPAK: a program for identifying clustering modes and packaging population structure

inferences across K. Molecular Ecology Resources,15:1179-1191.

Krajewski, C., Sipiorski, J.T. and Anderson, F.E. (2010). Complete Mitochondrial Genome

Sequences and the Phylogeny of Cranes (Gruiformes: Gruidae). The Auk, 127:440-452.

Krapu, G.L., Brandt D.A. and Cox, R.R. (2004). Less waste corn, more land in soybeans and

the switch to genetically modified crops: trends with important implications to wildlife

management. Wildlife Society Bulletin, 32:127–136.

Lambeck, K., Yokoyama, Y. and Purcell, T. (2002). Into and out of the Last Glacial

Maximum: sea-level change during Oxygen Isotope Stages 3 and 2. Quaternary Science

Reviews, 21(1-3):343-360.

Lamichhaney, S., Berglund, J., Wang, C., Almén, M.S., Webster, M.T., Grant, R.B., Grant,

P.R. and Andersson, L. (2016) A beak size locus in Darwin’s Finches facilitated character

displacement during a drought. Science, 352:62-84.

Lane, S. J., Azuma, A. and Higuchi H (1998). Wildfowl damage to agriculture in Japan.

Agric. Ecosyst. Environ., 70: 69-77.

Lavery, H.J. and Blackman J.G. (1969). The cranes of Australia. Queensland Agricultural

Journal, 95:156-162.

Page 170: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

170

Leigh, J.W. and Bryant, D. (2015). PopART: Full-feature software for haplotype network

construction. Methods in Ecology & Evolution, 6(9):1110–1116.

Le Roy, E. (2010). What has been done in the Champagne-Ardenne to prevent crane damages

on farmlands since 2004. In Nowald, G., Alexander, W., Fanke, J., Weinhardt, E. and

Donner, N. (Eds.), Proceedings of the VIIth European crane conference breeding, resting,

migration and biology, Stralsund, Germany (2010), pp. 53-56

Leito, A. (2014). Current situation of the Eurasian Crane in France and recent evolutions.

Scientific abstracts of VIII European Crane Conference (10–14 November 2014, Gallocanta,

Spain.

Lofdahl, K. (1979). Sympatry and Speciation in Australian Cranes. Unpublished research

proposal; International Crane Foundation, Wisconsin (ICF library Ref 04-053).

Lorenzen, B. and Madsen, J. (1986). Feeding by geese in the Filsø farmland, Denmark, and

the effect of grazing on yield structure of spring barley. Holarctic Ecology, 9:305-311.

Lundgren, S. (2013). Current status of the Common Crane in Sweden. Breeding, resting, and

colour banding. In: Nowald G, Weber A, Fanke J, Weinhardt E, and Donner N, editors.

Proceedings of the 7th European Crane Workshop. 16–18. Kranichschutz Deutschland. Groß

Mohrdorf.

Mahon, N., Crute, I and di Bonito, M. (2018). Towards a broad based and holistic framework

for Sustainable Intensification. Land Use Policy, 77:576-597.

Mallet, J., Ehrlich, P., Gill, F., McCormack, J. and Raven, P. (2018). In Comments on

Petition of Pacific Legal Foundation, et al., for Rule-Making Under the Administrative

Procedure Act (Which Aimed to Promulgate New Regulatory Definitions of “Species” and

Page 171: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

171

“Subspecies” Under the Endangered Species Act). http://nrs.harvard.edu/urn-

3:HUL.InstRepos:37568375.

Manfredo, M.J. (1989). Human dimensions of wildlife management. Wildlife Society

Bulletin, 17: 447-449.

Marchant, S. and Higgins, P.J. (eds). (1993). Handbook of Australian, New Zealand and

Antarctic Birds. Vol. 2. Raptors to Lapwings. Oxford University Press, Melbourne.

Martien, K.K., Leslie, M.S., Taylor, B.L., Morin, P.A., Archer, F.I., Hancock‐Hanser, B.L.,

Rosel, P.E., Vollmer, N.L., Viricel, A. and Cipriano, F. (2017). Analytical approaches to

subspecies delimitation with genetic data. Marine Mammal Science, 33:27-55.

Mattheissen, P. (2002). Birds of Heaven. The Harvill Press, London.

Matthiessen, P. (2014). From the Archives: The Cranes of Hokkaido, by Peter Matthiessen,

Audubon (7 April 2014) 165.

McCann, K.I., Thero, L.J and Morrison, K. (2007). Conservation priorities for the Blue Crane

(Anthropoides paradiseus) in South Africa - the effects of habitat changes on distribution and

numbers. Ostrich: Journal of African Ornithology, 78:205–211.

McCormack, J.E. and Maley, J.M. (2015). Interpreting negative results with taxonomic and

conservation implications: Another look at the distinctness of coastal California

Gnatcatchers. The Auk, 132(2):380-388.

Meares, K., Dawson, D.A., Horsburgh, G.J., Glenn, T.C., Jones, K.L., Braun, M.J., Perrin,

M.R. and Taylor, T.D. (2009). Microsatellite loci characterized in three African crane species

(Gruidae, Aves). Molecular Ecology Research, 9:308-311.

Page 172: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

172

Meine, C. D. and Archibald, G. W. (eds). (1996). The Cranes: Status survey and conservation

action plan. IUCN, Gland, Switzerland and Cambridge, UK.

Meirmans P.G. (2012). AMOVA-based clustering of population genetic data. Journal of

Heredity, 103:744-750.

Meirmans, P.G. and Van Tienderen, P.H. (2004). GENOTYPE and GENODIVE: two

programs for the analysis of genetic diversity of asexual organisms, Molecular Ecology

Notes, 4:792-794.

Meirmans, P.G. (2013). AMOVA-based clustering of population genetic data. Journal of

Heredity,103:744-750.

Menkhorst, P., Rogers, D., Clarke, R., Davies, J., Marsack, P. and Franklin, K. (2017). The

Australian Bird Guide. CSIRO Publishing, Clayton South, Australia.

Mewes, W. (2010). Population development, range of distribution and population density of

Common Cranes Grus grus in Germany and its federal states. Vogelwelt, 131:75–92.

Mewes W. (2012). Why are breeding populations of Eurasian Cranes increasing and

spreading in Germany? Abstracts of the International Workshop Management of Common

Cranes at the Hula Valley, Israel: Past, Present and Future (16–18 December 2012).

Miller, A. (2016). The development of microsatellite loci through next generation

sequencing, and a preliminary assessment of population genetic structure for the iconic

Australian crane, Brolga (Antigone rubicunda). Nature Glenelg Trust, Warrnambool,

Victoria.

Page 173: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

173

Miller, A., Veltheim. I., Nevard, T.D., Gan, H.M. and Haase, M. (2019). Microsatellite loci

and the complete mitochondrial DNA sequence characterized through next generation

sequencing and de novo genome assembly, and a preliminary assessment of population

genetic structure for the Australian crane, Antigone rubicunda. Avian Biology Research,

doi/10.1177/1758155919832142.

Miller, M.P. and Haig, S.M. (2010). Identifying shared genetic structure patterns among

Pacific Northwest forest taxa: insights from use of visualization tools and computer

simulations. PLoS One, 5, e13683.

Miller, S.A., Dykes, D.D. and Polesky, H.F. (1988). A simple salting out procedure for

extracting DNA from human nucleated cells. Nucleic Acids Research, 16:1215.

Mooney, S. D., Sniderman, J. M., Kershaw, P., Haberle, S. G. and Roe, J. (2017). Quaternary

Vegetation in Australia. In D. A. Keith (Ed.), Australian Vegetation (3rd ed., pp. 63-88).

Cambridge, United Kingdom: Cambridge University Press.

Mudrik, E.A., Kashentseva, T.A., Redchuk, P.S. and Politov, D.V. (2015). Microsatellite

variability data confirm low genetic differentiation of western and eastern subspecies of

common crane Grus grus L. (Gruidae, Aves). Molecular Biology, 49(2):260–266.

Munoz-Fuentes, V., Green, A.J. and Sorenson, M.D. (2008). Comparing the genetics of wild

and captive populations of White-headed Ducks Oxyura leucocephala: consequences for

recovery programmes. Ibis, 150:807-815.

Myers, A. (2001). Factors influencing the nesting success of Brolgas, Grus rubicundus, in

Western Victoria. Honours thesis. School of Ecology and Environment, Deakin University,

Victoria, Australia.

Page 174: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

174

Nei, M. (1987). Molecular evolutionary genetics. Columbia University Press, New York.

Nevard, T.D., Haase, M., Archibald, G., M., Leiper, I., Van Zalinge, R., Purchkoon, N.,

Siriaroonrat, B., Latt, T.N., Wink, M. and Garnett, S.T. (in press). Subspecies in the Sarus

Crane Antigone antigone revisited; with particular reference to the Australian population.

[Submitted for publication to PLOS ONE 26.07.2019].

Nevard, T.D., Leiper, I., Archibald, G. and Garnett, S.T. (2018). Farming and cranes on the

Atherton Tablelands, Australia. Pacific Conservation Biology, doi:10.1071/PC18055.

Nevard, T.D., Franklin, D.C., Leiper, I., Archibald, G., Garnett, S.T. (2019a). Agriculture,

Brolgas and Australian Sarus Cranes on the Atherton Tablelands, Australia. Pacific

Conservation Biology, doi:10.1071/PC18081.

Nevard, T.D., Haase, M., Archibald, G., Leiper, I. and Garnett, S.T. (2019b). The Sarolga:

conservation implications of genetic and visual evidence for hybridization between the

Brolga Antigone rubicunda and the Australian Sarus Crane Antigone antigone gillae. Oryx,

doi:10.1017/S003060531800073X.

Newell, P. and Taylor, O. (2018). Contested landscapes: the global political economy of

climate-smart agriculture. Journal of Peasant Studies, 45:108–129.

Nilsson, L., Bunnefeld, N., Persson, J. and Månsson J. (2016). Large grazing birds and

agriculture — predicting field use of common cranes and implications for crop damage

prevention. Agriculture, Ecosystems and Environment, 219:163–170.

Nix, H.A. and Switzer, M.A. (1991). Rainforest Animals. Atlas of Vertebrates Endemic to

Australia’s Tropics. Australian Parks and Wildlife Service. Canberra.

Page 175: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

175

Nowald, G. (2010). Cranes and People: Agriculture and Tourism. pp.60-65. in Cranes,

Agriculture, and Climate Change. Proceedings of a workshop organized by the International

Crane Foundation and Muraviovka Park for Sustainable Land Use. Ed. Harris J. International

Crane Foundation, Baraboo.

Nowald, G. (2012). Cranes and people: agriculture and tourism. In: Harris, J., editor.

Proceedings of the Cranes, Agriculture, and Climate Change Workshop at Muraviovka Park.

Russia, 28 May-3 June 2010. 60–64. International Crane Foundation.

Nowald, G. and Mewes, W. (2010). Cranes and people. In: Koga, K., Hu, D. and Momose,

K., editors. Cranes and people: prologue to a new approach for conservation of the Red-

crowned Crane. 13–14, Kushiro, Japan.

Nunn, N. and Qian, N. (2010). The Columbian exchange: A history of disease, food, and

ideas. Journal of Economic Perspectives 24:163–188.

Patten, M.A. and Remsen J.V. Jr (2017). Complementary roles of phenotype and genotype in

subspecies delimitation. Journal of Heredity, 108(4): 462-464.

Patten, M.A. and Unitt, P. (2002). Diagnosability versus mean differences of Sage Sparrow

subspecies. The Auk, 119:26–35.

Payseur, B.A. and Rieseberg, L.H. (2016) A genomic perspective on hybridization and

speciation. Molecular Ecology, 25:2337-2360.

Pearse, A.T., Krapu, G.L., Brandt, D.A. and Kinzel, P.J. (2010). Changes in agriculture and

abundance of snow geese affect carrying capacity of Sandhill cranes in Nebraska. Journal of

Wildlife Management 74:479–488.

Page 176: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

176

Pennisi, E. (2016). Shaking up the Tree of Life. Science, 354:817-821.

Petit, R.J., el Mousadik, A. and Pons, O. (1998). Identifying populations for conservation on

the basis of genetic markers. Conservation Biology,12:844-855.

Phalan, B., Bertzky, M., Stuart, H. M., Donald, P.F., Scharlemann, J.P.W., Stattersfield, A.J.

and Balmford A. (2013). Crop expansion and conservation priorities in tropical countries.

PLOS ONE, 8(1).

Phalan, B., Green, R.E., Dicks, L.V., Dotta, G., Feniuk, C., Lamb, A., Strassburg, B.B.N.,

Williams, D.R., zu Ermgassen, E.K.H.J. and Balmford, A. (2016). How can higher-yield

farming help to spare nature? Science, 351:450–451.

Prange, H. (2010). Migration and resting of the Common Crane Grus grus and changes in

four decades. Vogelwelt, 2:155–168.

Prange, H. (2012). Reasons for changes in crane migration patterns along the West-European

Flyway. In: Harris, J, editor. Proceedings of the Cranes, Agriculture and Climate Change

Workshop, 28 May – 3 June 2010, 35–48. International Crane Foundation.

Prange, H. (2015). Distribution and migration of the Eurasian Crane on the West-European

route. In: Ilyashenko, E.I. and Winter, S.V. (eds). Cranes of Eurasia (biology, distribution,

captive breeding). 5:287-312. Moscow: Crane Working Group.

Pritchard, J.K., Stephens, M. and Donnelly, P. (2000). Inference of population structure using

multilocus genotype data. Genetics,155:945–959.

Queensland Government (2016). Land use mapping – 1999 to 2015 – Wet Tropics and

Northern Gulf NRM regions.

Page 177: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

177

Ramankutty, N., Foley, J.A. and Olejniczak, N.J. (2002). People on the land: changes in

population and global croplands during the 20th century. Ambio, 31:251−257.

Ramasamy, R.K., Ramasamy, S., Bindroo, B.B. and Naik, V.G. (2014) STRUCTURE PLOT:

a program for drawing elegant STRUCTURE bar plots in user friendly interface.

SpringerPlus, 3:431.

Ray, D.K., Mueller, N.D., West, P.C. and Foley, J.A. (2013). Yield Trends Are Insufficient to

Double Global Crop Production by 2050.PLOS ONE, 19:8(6)

Raymond, M. and Rousset, F. (1995). GENEPOP (version 1.2): population genetics software

for exact tests and ecumenicism. Journal of Heredity, 86:248-249.

Reardon, M. (2007). Brolga Country: Travels in Wild Australia. Allen and Unwin. Crows

Nest, Australia.

Redchuk, P., Fesenko, G.V. and Slyusar, N.V. (2015). Migration Routes of the Eurasian

Crane in Ukraine. In: Ilyashenko EI, Winter SV, editors. Cranes of Eurasia (biology,

distribution, captive breeding) 5. Proceedings of International Scientific Conference “Cranes

of Eurasia: biology, conservation, management,” p 313–334. Trans-Baikalia, Russia, 1‒4

September 2015. Moscow, Russia: Crane Working Group of Eurasia.

Rhymer, J.M. and Simberloff, D. (1996). Extinction by hybridization and introgression.

Annual Review of Ecology and Systematics, 27:83–109.

Rieseberg, L.H., Archer, M.A. and Wayne, R.K. (1999). Transgressive segregation,

adaptation and speciation. Heredity, 83:363-372.

Roberts, A. (2017). Tamed. Ten Species that Changed Our World. Windmill Books, London.

Page 178: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

178

Roberts, D.G., Baker, J. and Perrin, C. (2011). Population genetic structure of the endangered

Eastern Bristlebird, Dasyornis brachypterus; implications for conservation. Conservation

Genetics, 12(4)1075-1085.

Robinson, R.A., Wilson, J.D. and Crick, H.Q. (2001). The importance of arable habitat for

farmland birds in grassland landscapes. Journal of Applied Ecology, 38(5):1059-1069.

Ronquist, F., Teslenko, M., van der Mark, P., Ayres, D.L/, Darling, A., Höhna, S., Larget, B.,

Liu, L., Suchard, M.A. and Huelsenbeck, J.P. (2012). MrBayes 3.2: efficient Bayesian

phylogenetic inference and model choice across a large model space. Systema Biologica,

61:539-542.

Rousset, F. (2008). Genepop'007: a complete reimplementation of the Genepop software for

Windows and Linux. Molecular Ecology Resources, 8:103-106.

Runge, C.A., Gallo‐Cajiao, E., Carey, M.J., Garnett, S.T., Fuller, R.A. and McCormack, P.C.,

(2017). Coordinating domestic legislation and international agreements to conserve migratory

species: A case study from Australia. Conservation Letters, 10:765-772.

Runge, C.A., Martin, T.G., Possingham, H.P., Willis, S.G. and Fuller, R.A., (2014).

Conserving mobile species. Frontiers in Ecology and the Environment, 12:395-402.

Safner, T., Miller, M.P., McRae, B., Fortin, M.J. and Manel, S. (2011). Comparison of

Bayesian clustering and edge detection methods for inferring boundaries in landscape

genetics. International Journal of Molecular Science, 12:865-889.

Page 179: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

179

Salvi, A. (2015). The Eurasian Crane in France: evolutions in the last four decades In:

Ilyashenko, E.I. and Winter, S.V. (eds). Cranes of Eurasia (biology, distribution, captive

breeding). 5:191–205. Moscow: Crane Working Group.

Sankhom, R., Warrit, N., Pimpakan, T. and Wiwegweaw. (2018). A Genetic Diversity of

Captive Eastern Sarus Crane in Thailand. Inferred from Mitochondrial Control Region

Sequence and Microsatellite DNA markers. Tropical Natural History.18(2):84-96.

Sankhom, R., Warrit, N. and Wiwegweaw, A. (2018). Screening and application of

microsatellite markers for genetic diversity analysis of captive eastern sarus crane Grus

antigone sharpie (Blanford 1895) in Thailand. Zoo Biology, 37(5):310-319.

Sandvik, J. (2010). Results of the colour ringing of cranes in Norway. Abstracts of the VIIth

European Crane Conference, Oct. 14-17, Stralsund, Germany.

Schodde, R. (1988). New sub-species of Australian birds. Canberra Bird Notes, 13:1

Schodde, R. and Mason, I. J. (1999). Directory of Australian Birds: Passerines. CSIRO

Publishing, Melbourne.

Scott, H.A. and Scott R.M. (1996). A conservation program for the Blue Crane in the

Overberg, Southern Cape, South Africa. In: Beilfuss, R.D., Tarboton, R.W. and Gichuki,

N.N. (eds). Proceedings of the 1993 African Crane and Wetland Training Workshop, 8-15

August 1993. p395–402. International Crane Foundation.

Sensis, (2015). Cairns District and Atherton Tablelands Yellow Pages 2014/15.

Shanni, I., Labinger, Z. and Alon. D. (2012). A review of the crane-agriculture conflict, Hula

Valley, Israel. In: Harris J, editor. Proceedings of the Cranes, Agriculture, and Climate

Page 180: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

180

Change Workshop at Muraviovka Park. Russia, 28 May-3 June 2010. pp101–105.

International Crane Foundation.

Sharpe, R.B. (1894). Catalogue of Birds in the British Museum, vol. 23. British Museum of

Natural History, London.

Sherfy, M.H., Anteau, M.J. and Bishop, A.A. (2011). Agricultural practices and residual corn

during spring crane and waterfowl migration in Nebraska. Journal of Wildlife Management,

75:995–1003.

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

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

Soltis, P. and Soltis, D.E. (2009). The role of hybridization in plant speciation. Annual

Review of Plant Biology, 60:561-588.

Southwood, T.R.E. and Henderson, A. (2000). Ecological Methods (3rd

Edition). Blackwell

Science Ltd, Oxford

Steven, R., Morrison, C., Arthur, J.M., and Castley, J.G. (2015). Avitourism and Australian

Important Bird and Biodiversity Areas. PLOS ONE 10:e0144445.

Stone, Z.L., Tasker, E. and Maron, M. (2018). Grassy patch size and structure are important

for northern Eastern Bristlebird persistence in a dynamic ecosystem. Emu, 118:269-280.

Sugden, L. G., Clark, R. G., Woodsworth, E. J and Greenwood H. (1988). Use of cereal fields

by foraging Sandhill Cranes in Saskatchewan. Journal of Appied Ecology, 25:111-124.

Page 181: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

181

Sundar, K.S.G., Grant, J.D.A., Veltheim, I., Kittur, S., Brandis. K., McCarthy, M.A. and

Scambler, E.C. (2018). Sympatric cranes in northern Australia: abundance, breeding success,

habitat preference and diet. Emu, doi: 10.1080/01584197.2018.1537673.

Szabo, J.K., Khwaja, N., Garnett, S.T. and Butchart, S.H.M. (2012). Global patterns and

drivers of avian extinctions at the species and subspecies level. PLOS ONE 7:e47080.

Van Velden, J.L., Smith, T. and Ryan, P.G. (2016). Cranes and Crops: Investigating farmer

tolerances toward crop damage by threatened blue cranes (Anthropoides paradiseus) in the

Western Cape, South Africa. Environmental Management, 58(6):972-983.

Templeton, A.R, Shaw, K., Routman, E. and Davis, S.K. (1990). The genetic consequences

of habitat fragmentation. Annals of the Missouri Botanical Garden, 77:13-27.

Timmer, C.P., Falcon, W.P. and Pearson, S.R. (1983). Food policy Analysis. Johns Hopkins

University Press, Baltimore.

Tobias, J. A., Seddon, N., Spottiswoode, N., Pilgrim, J.D., Fishpool, L.D.C. and Collar, N.J.

(2010). Quantitative criteria for species delimitation. Ibis, 152:724–746.

Todesco, M., Pascual, M.A., Owens, G.L., Ostevik, K.L., Moyers, B.T., Hübner, S., Heredia,

S.M., Hahn, M.A., Caseys, C., Bock, D.G. and Rieseberg, L.H. (2016). Hybridization and

extinction. Evolutionary Applications, 9:892-908.

Toews, D.P.L and Brelsford, A. (2012). The biogeography of mitochondrial and nuclear

discordance in animals. Molecular Ecology, 16:3907-3930.

Page 182: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

182

Tofft, J. (2013). Current status of the Common Crane (Grus grus) in Denmark. In: Nowald G,

Weber A, Fanke J, Weinhardt E and Donner N, eds. Proceedings of the 7th European Crane

Workshop. pp19–21. Kranichschutz Deutschland.

Townsville Daily Bulletin (1935). Atherton Tableland Notes, page 5, Friday 20 September

1935. National Library of Australia http://nla.gov.au/nla.news-article62410202

Tracey, J, Bomford, M, Hart, Q, Saunders, G. and Sinclair, R (2007). Managing Bird Damage

to Fruit and Other Horticultural Crops.

http://www.dpi.nsw.gov.au/agriculture/horticulture/pestsdiseases-hort/multiple/managing-

bird-damage

Triggs, S.J. and Daugherty, C.H. (1996). Conservation and genetics of New Zealand

parakeets. Bird Conservation International, 6:89-101.

Tscharntke, T., Clough, Y., Wanger, T.C., Jackson, L., Motzke, I., Perfecto, I., Vandermeer,

J. and Whitbread, A. (2012). Global food security, biodiversity conservation and the future of

agricultural intensification. Biological Conservation, 151:53-9.

US Fish & Wildlife Service. Service Determines Coastal California Gnatcatcher is a

Subspecies and Remains Threatened under Endangered Species Act.

https://www.fws.gov/news/ShowNews.cfm?ref=service-determines-coastal-california-

gnatcatcher-is-a-subspecies-and-&_ID=35782. 2016 [Cited 14 May 2019].

Van der Geer, A., Lyras, G., de Vos, J. and Dermitzakis, M (2010). Evolution of Island

Mammals, Adaption and Extinction of Placental Mammals on Islands. Wiley-Blackwell,

Oxford.

Page 183: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

183

Van Oosterhout, C., Hutchinson, W.F., Wills, D.P.M. and Shipley, P. (2004). MICRO-

CHECKER: software for identifying and correcting genotyping errors in microsatellite data.

Molecular Ecology Notes, 4:535-538.

Von Treuenfels, C-A. (2006). The Magic of Cranes. Abrams, New York.

Voris, H.K. (2000). Maps of Pleistocene sea levels in Southeast Asia: Shorelines, river

systems and time durations. Journal of Biogeography, 27:1153-1167.

Vijay, N., Bossu, C.M., Poelstra, J.W., Weissensteiner, M.H., Suh, A., Kryukov, A.P. and

Wolf, J.B.W. (2016). Evolution of heterogeneous genome differences across multiple contact

zones in a crow species complex. Nature Communications, 7:13195.

Waits, L.P., Luikart, G. and Taberlet, P. (2001). Estimating the probability of identity among

genotypes in natural populations: cautions and guidelines. Molecular Ecology, 10, 249-256.

Walkinshaw, L. H. (1973). Cranes of the World. Winchester Press, New York.

Wallin, H., Kvamme, T. and Bergsten, J. (2017). To be or not to be a subspecies: description

of Saperda populnea lapponica ssp. (Coleoptera, Cerambycidae) developing in Downy

Willow (Salix lapponum L.). ZooKeys, 69:103-148

Watanabe, T (2006). Comparative breeding ecology of Lesser Sandhill Cranes (Grus

canadensis canadensis) and Siberian Cranes (G. leucogeranus) in Eastern Siberia. A

dissertation by Tsuyoshi Watanabe submitted to the Office of Graduate Studies of Texas

A&M University in partial fulfilment of the requirements for the degree of Doctor of

Philosophy. December 2006.

Page 184: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

184

White, D. M. (1987). The status and distribution of the Brolga in Victoria, Australia. In:

Proceedings of the 1983 International Crane Workshop. Baraboo, Wisconsin, USA:

International Crane Foundation. Based on the proceedings of the workshop held in Bharatpur,

India, February 1983. (eds G. Archibald and R. F. Pasquier) pp. 115-31.

White, M. E. (1997). After The Greening. Kangaroo Press, Kenthurst.

Whitfield, S., Challinor, A.J. and Rees, R.M. (2018). Frontiers in climate smart food systems:

outlining the research space. Frontiers in Sustainable Food Systems, 2:2.

Whitlock, M.C., McCauley, D.E. (1999). Indirect measures of gene flow and migration: FST ≠

1/(4 Nm + 1). Heredity, 82: 117-125.

Wilson, E.O. and Brown, W.L. Jr (1953). The subspecies concept and its taxonomic

application. Systema Zoologica, 2:97–111.

Winker, K. (2009). Reuniting phenotype and genotype in biodiversity research. BioScience,

59:657-665.

Wood, T.C. and Krajewski, C. (1996). Mitochondrial sequence variation among the

subspecies of Sarus Crane (Grus antigone). The Auk, 113:655-633.

Wright, H.L., Lake, I.R. and Dolman, P.M. (2012). Agriculture - a key element for

conservation in the developing world. Conservation Letters, 5:11–19.

Zachos, F.E. (2018). Species Concepts in Biology: Historical Development, Theoretical

Foundations and Practical Relevance. Switzerland, Springer Nature.

Zink, R.M., Groth, J.G., Vázquez-Miranda, H. and Barrowclough, G.F. (2013).

Phylogeography of the California Gnatcatcher (Polioptila californica) using multilocus DNA

Page 185: Interactions between Brolgas and Sarus Cranes in Australia · Research Institute for the Environment and Livelihoods Interactions between Brolgas and Sarus Cranes in Australia Timothy

185

sequences and ecological niche modelling: Implications for conservation. The Auk,

130(3):449-458.

Zong, Y., Huang, G., Li, X.Y. and Sun, Y.Y. (2016). Late Quaternary tectonics, sea-level

change and lithostratigraphy along the northern coast of the South China Sea. Geological

Society of London, Special Publications, 429(1):123-136.