<<

Copyright and use of this thesis This thesis must be used in accordance with the provisions of the Copyright Act 1968. Reproduction of material protected by copyright may be an infringement of copyright and copyright owners may be entitled to take legal action against persons who infringe their copyright. Section 51 (2) of the Copyright Act permits an authorized officer of a university library or archives to provide a copy (by communication or otherwise) of an unpublished thesis kept in the library or archives, to a person who satisfies the authorized officer that he or she requires the reproduction for the purposes of research or study. The Copyright Act grants the creator of a work a number of moral rights, specifically the right of attribution, the right against false attribution and the right of integrity. You may infringe the author’s moral rights if you: - fail to acknowledge the author of this thesis if you quote sections from the work - attribute this thesis to another author - subject this thesis to derogatory treatment which may prejudice the author’s reputation For further information contact the University’s Director of Copyright Services sydney.edu.au/copyright

The role of ecological interactions: how intrinsic and extrinsic factors shape the spatio-temporal dynamics of populations

Aaron C. Greenville

A thesis submitted in fulfilment of the requirements for the degree of Doctor of Philosophy Faculty of Science The University of Sydney, February 2015

Declaration of originality

I hereby declare that the work contained in this thesis is my own and contains the

results of an original investigation, except where otherwise referenced or acknowledged. This

work was carried out while I was enrolled as a student for the degree of Doctor of Philosophy

in the School of Biological Sciences, The University of Sydney. This thesis has not been

previously submitted for examination at this, or any other university.

Aaron Greenville

February 2015

Contents

List of figures ...... v List of tables ...... ix Glossary:...... xiii Acknowledgments ...... xv Thesis Abstract ...... xix Preface ...... xxv Chapter 1: General introduction...... 1 Introduction ...... 2 Climate change and extreme events ...... 5 Importance of including multiple population drivers ...... 6 Spatial and temporal population dynamics ...... 7 Resource pulse environments ...... 8 Thesis conceptual model ...... 9 Study region: the Simpson , central Australia ...... 12 Thesis aims ...... 16 Overview of thesis ...... 16 References: ...... 19 Chapter 2: Extreme climatic events drive mammal irruptions: regression analysis of 100-year trends in desert rainfall and temperature ...... 29 Abstract ...... 30 Introduction ...... 31 Materials and methods ...... 34 Study region ...... 34 Temperature ...... 35 Rainfall ...... 36 Small mammals ...... 38 Results ...... 40 Temperature ...... 40 Rainfall ...... 44 Extreme rainfall return intervals ...... 48 Rainfall variability ...... 48 Small mammals ...... 50 Discussion ...... 52 Temperature ...... 53 Rainfall ...... 54

i

Case study: small mammals ...... 56 Implications of climate change ...... 57 Acknowledgements ...... 58 References: ...... 59 Chapter 3: Extreme rainfall events predict irruptions of rat plagues in central Australia ...... 65 Abstract ...... 66 Introduction ...... 67 Methods ...... 69 Study region ...... 69 Prediction 1: Large rainfall events drive long-haired rat irruptions ...... 72 Prediction 2: Long-haired rats irrupt from multiple refugia ...... 72 Prediction 3: Colonisation of new sites will be rapid ...... 74 Prediction 4: Young males arrive first at new sites ...... 75 Results ...... 76 Prediction 1: Large rainfall events drive long-haired rat irruptions ...... 76 Prediction 2: Long-haired rats irrupt from multiple refugia ...... 79 Prediction 3: colonisation of new sites will be rapid ...... 82 Prediction 4: Young males arrive first at new sites ...... 83 Discussion ...... 87 Prediction 1: large rainfall events drive long-haired rat irruptions ...... 87 Prediction 2: long-haired rats irrupt from multiple refugia ...... 88 Prediction 3: colonisation of new sites will be rapid ...... 89 Prediction 4: young males arrive first at new sites...... 89 Acknowledgements ...... 90 References: ...... 92 Chapter 4: Spatial and temporal dynamics in mammal populations: the role of intrinsic and extrinsic factors ...... 95 Abstract ...... 96 Introduction ...... 97 Material and Methods...... 100 Study region ...... 100 Small mammals ...... 100 Wildfire ...... 102 Rainfall ...... 102 Spinifex cover and seed ...... 103 Multivariate autoregressive state-space models ...... 103 Measures of synchrony ...... 107

ii

Results ...... 107 Rodents ...... 107 Dasyurids ...... 113 Discussion ...... 116 Acknowledgements ...... 121 References: ...... 122 Chapter 5: Spatial and temporal synchrony in population dynamics in variable environments ...... 127 Abstract ...... 128 Introduction ...... 129 Material and Methods...... 132 Study region ...... 132 ...... 132 Wildfire ...... 134 Rainfall ...... 134 Spinifex cover ...... 135 Multivariate autoregressive state-space models ...... 135 Measures of synchrony ...... 138 Results ...... 139 ...... 139 Dragons ...... 145 Discussion ...... 147 Acknowledgements ...... 150 References: ...... 151 Chapter 6: Bottom up and top down processes interact to modify intraguild interactions in resource-pulse environments ...... 155 Abstract ...... 156 Introduction ...... 157 Conceptual models ...... 160 Case study ...... 162 Materials and methods ...... 163 Study region ...... 163 Remote cameras ...... 164 Data analyses ...... 165 Results ...... 166 Discussion ...... 170 Acknowledgements ...... 177

iii

References: ...... 178 Chapter 7: Predicting species interactions under a changing climate ...... 183 Abstract ...... 184 Introduction ...... 185 Methods and materials ...... 189 Study region ...... 189 Small mammals and reptiles ...... 189 Predator monitoring ...... 190 Spinifex cover and seed ...... 190 Rainfall and wildfire ...... 191 Structural equation modelling ...... 191 Model justification ...... 193 Changes in rainfall and wildfire from climate change ...... 195 Results ...... 196 Structural equation modelling ...... 196 Climate change ...... 198 Discussion ...... 199 Climate change ...... 200 Conclusion ...... 202 Acknowledgements ...... 203 References: ...... 204 Chapter 8: General Discussion and Conclusions ...... 211 Overview ...... 212 Achievements ...... 213 Limitations and future research ...... 217 Concluding remarks ...... 219 Acknowledgements ...... 220 References: ...... 222 Appendices ...... 227 Appendix 1: ...... 228 Appendix 2: ...... 229 Appendix 3: ...... 244 Appendix 4: ...... 246 Appendix 5: ...... 251 Appendix 6: ...... 257

iv

List of figures

List of figures

Figure 1.1: Conceptual trophic interaction model that incorporates space as a potential driver of populations. It allows for bottom-up and top-down processes to be incorporated, as well as indirect interactions. The strengths of interactions and bottom-up or top-down effects are not assumed to be constant and may vary temporally. Adapted from Dickman et al. (2014). Artwork by A. Foster...... 11

Figure 1.2: Annual rainfall at nine sites surveyed for biotic and abiotic interactions in the , central Australia. The timing of extreme rainfall events is similar across sites, but there is otherwise large spatial variation in rainfall events. Data were recorded at site-based automatic weather stations (Environdata, Warwick, ), installed in 1995...... 15

Figure 1.3: Extent of spinifex (hummock) grasslands in Australia, as mapped by Department of the Environment and Water Resources (2007) and location of the Simpson Desert. The study region (insert) encompasses four properties: Tobermorey, Cravens Peak, Carlo and Ethabuka. Detail of sampling sites on these properties is shown in Fig. A2.1...... 15

Figure 2.1: Location of study region; the Simpson Desert, Australia, in relation to the Bureau of Meteorology (BOM) weather stations. is the location of the live mammal trapping used in this study...... 35

Figure 2.2: Mean minimum annual temperatures (° C) from a) Boulia, b) , c) , and d) weather stations, Simpson Desert, central Australia. Broken grey lines show 95th, 90th, 50th, 10th and 5thquantiles...... 41

Figure 2.3: Mean maximum annual temperatures (° C) from a) Birdsville, b) Oodnadatta, c), Alice Springs and d) Boulia weather stations, Simpson Desert, central Australia. Broken grey lines show 95th, 90th, 50th, 10th and 5thquantiles...... 42

Figure 2.4: Annual rainfall (mm) recorded each year for a) Bedourie, b) Glenormiston, c) Undoolya, d) Oodnadatta, e) Birdsville, f) Marion Downs, and g) Alice Springs weather stations, Simpson Desert, central Australia. Broken grey lines show 95th, 90th, 50th, 10th and 5thquantile. 90 and 95thquantiles represent boom years, 50th is the median and 10th and 5th represent bust years...... 46

Figure 2.5: The response of a) rodent and b) dasyurid marsupial captures to rainfall in the preceding year. Rainfall was recorded from on-site automatic weather station at Ethabuka (Environdata, Warwick, Queensland). Missing values were filled in from the nearest weather station at Sandringham, 63 km away. Lines depict predicted values from piecewise regression. Rodent captures show a response to rainfall only after a threshold of 418 mm (~ 0.95 quantile) is reached. Labels on (a) indicate year of irruption. No relationship between rainfall and dasyurids was found. Small mammal captures were standardised per 100 trap nights (TN) and averaged per year (n = 22 years)...... 51

Figure 3.1: Location of study sites across , Carlo Station and Ethabuka Reserve, Simpson Desert. Insert shows locations of Bureau of Meteorology weather stations used for predicting rat plagues...... 71

v

List of figures

Figure 3.2: Predicted probability of a long-haired rat, Rattus villosissimus, irruption from annual rainfall the year before. Rat plagues were collated from records within the last 100 years for Alice Springs, Glenormiston, Boulia and Ethabuka Reserve. Predicted results from a quasi-binomial generalised linear model. Grey triangle represent 2010 rainfall event. Dashed lines represent standard errors...... 77

Figure 3.3: Captures of long-haired rats, Rattus villosissimus, for each grid at Main Camp in May 1991, Simpson Desert, Queensland. Higher captures were found on grids in the south at the beginning of the 1991 irruption...... 80

Figure 3.4: Captures of long-haired rat (per grid), Rattus villosissimus, from a) distance and b) direction (angle) to nearest drainage line, Simpson Desert, Queensland. Points in (a) represents observed captures and lines represent predicted captures from GLM. Points on circle in (b) represent trapping grids and length of arrows are proportional to the numbers of captures for each grid...... 81

Figure 3.5: Predicted captures per grid from negative binomial generalised linear models for captures of long-haired rats, Rattus villosissimus, across each trip in a) 1991 irruption and b) 2011 irruption, Simpson Desert, Queensland. Lines indicate standard errors...... 83

Figure 3.6: Mean captures (per grid) of each category of reproductive condition for a) female and b) male long-haired rats, Rattus villosissimus, for each site at the beginning of the 2011 irruption (May 2011), Simpson Desert, Queensland. Each site sampled two grids, except Main Camp which had four. Lines indicate standard errors...... 85

Figure 3.7: Mean mass (g) of male and female long-haired rats, Rattus villosissimus, captured during the 2011 irruption, Simpson Desert, Queensland. Sites pooled for each trip. Numbers above bars represent total captures of males and females. Lines indicate standard errors...... 86

Figure 4.1: Predicted population size (line) and captures from each of the nine sub- populations (dots; log+1 captures/100 trap nights) from MARSS models for (a) Pseudomys hermannsburgensis, (b) Notomys alexis, (c) Dasycercus blythi single population and (d) Dasycercus blythi populations that experienced a wildfire (red line) and populations that did not experience a wildfire (blue line), Simpson Desert, central Australia. Time-series data were collected from nine sites (sub-populations) monitored 2–6 times per year for 17–22 years (1990–2012 for one site and 1995–2012 for eight sites; 130 sampling trips). Shaded areas indicate 95% confidence intervals...... 112

Figure 4.2: Predicted population size (line) and captures (dots; log+1 captures/100 trap nights) from a MARSS model for Sminthopsis youngsoni depicting asynchronous spatial population dynamics, Simpson Desert, central Australia. Time-series data were collected from nine sites (sub-populations) monitored 2–6 times per year for 17–22 years (1990–2012 for one site and 1995–2012 for eight sites; 130 sampling trips). Shaded areas indicate 95% confidence intervals...... 114

Figure 4.3: Predicted population size (line) and captures (dots; log+1 captures/100 trap nights) from a MARSS model for Ningaui ridei depicting asynchronous spatial population dynamics, Simpson Desert, central Australia. Time-series data were collected from nine sites (sub-populations) monitored 2–6 times per year for 17–22 years (1990–2012 for one site and

vi

List of figures

1995–2012 for eight sites; 130 sampling trips). Shaded areas indicate 95% confidence intervals...... 115

Figure 5.1: Predicted population size (line) and captures (dots; log+1 captures/100 trap nights) estimated from MARSS models for (a) Ctenotus pantherinus (red line: burnt, blue line unburnt), (b) Ctenotus dux (red line burnt, blue line: unburnt) and (c) Ctenotus ariadnae, Simpson Desert, central Australia. Time-series data were collected from nine sites (sub- populations) monitored 2–6 times per year for 17–22 years (1990–2012 for one site and 1995–2012 for eight sites; pooled per year). Shaded areas indicate 95% confidence intervals...... 144

Figure 5.2: Predicted population size (line) and captures (dots; log+1 captures/100 trap nights) estimated from MARSS models for (a) Lerista labialis synchronous population and (b) L. labialis populations that had experienced a wildfire (red line: burnt, blue line unburnt), Simpson Desert, central Australia. Time-series data were collected from nine sites (sub- populations) monitored 2–6 times per year for 17–22 years (1990–2012 for one site and 1995–2012 for eight sites; pooled per year). Shaded areas indicate 95% confidence intervals...... 145

Figure 5.3: Predicted population size (line) and captures (dots; log+1 captures/100 trap nights) estimated from MARSS models for (a) Ctenophorus nuchalis and (b) Ctenophorus isolepis populations at either oases (blue line) or open desert sites (black line), Simpson Desert, central Australia. Time-series data were collected from nine sites (sub-populations) monitored 2–6 times per year for 17–22 years (1990–2012 for one site and 1995–2012 for eight sites; pooled per year). Shaded areas indicate 95% confidence intervals...... 146

Figure 6.1: Conceptual model depicting the population dynamics of prey in resource-pulse environments. ‘Booms’, or population irruptions, arise from large resource pulses that are stimulated by events such as flooding rainfall. Declines represent subsequent, sometimes brief, periods of negative increase, while ‘busts’ are often long periods when resources and prey numbers are relatively low...... 159

Figure 6.2: Conceptual models of dominant and subordinate predator interactions in resource-pulse environments. Three processes are possible; a) top-down only, b) bottom-up only and c) both top-down and bottom-up. Linear relationships are shown for simplicity, but non-linear relationships may be possible. Scale represents abundance...... 161

Figure 6.3: Interactions between prey population phase (bust, boom and decline) and the effects of dominant predators on subordinate predators: a) dingo-red fox, b) dingo-feral cat and c) red fox-feral cat, represented by numbers of photographic images from two years of continuous remote camera trapping in the Simpson Desert, central Australia. Log numbers are shown on the y-axis. Solid and dashed lines are predicted values from GLMs...... 168

Figure 6.4: Daily activity patterns of the dingo, red fox, feral cat and rodents (prey) during a) bust, b) boom and c) decline phases of the prey population, Simpson Desert, central Australia. Numbers represent 24-hour clock times, arrows represent mean activity times and grey boxes frequency (square-root) of observations recorded each hour from two years of continuous remote camera trapping...... 170

Figure 6.5: Complex dominant-subordinate predator interactions in a) the boom phase and b) the bust or decline phases of their prey. Line weights represent relative interaction strengths.

vii

List of figures

Dashed lines represent indirect interactions. This model predicts that, in times when the prey population is most vulnerable to predation from the mesopredators (bust and decline phases), the top predator has a net positive effect on prey. Artwork by A. Foster...... 176

Figure 7.1: Overview of the modelling process used to integrate multiple datasets into a trophic interactions model, informed by the results of this and past studies. Once significant pathways (biotic and abiotic interactions) are found using structural equation modelling (SEM), the model is refined and used to predict species’ population changes due to changes in rainfall and wildfire, in line with climate change scenarios...... 188

Figure 7.2: (a) Conceptual model describing the proposed interactions between drivers and trophic levels in the Simpson Desert, Australia. (b) Results from structural equation model. Values above drawings are percentage deviance explained. Standardised path coefficients are shown for each arrow. Arrow thickness proportional to effect size and represent significant path coefficients (P<0.05). Artwork by A. Foster...... 197

Figure 7.3: Projected results (±SE) for changes in (a) spinifex cover (b) spinifex seed index, and (c) rodent captures with all mammalian predators present, introduced predators (cats and foxes) removed and all mammalian predators removed, under a changing rainfall and wildfire scenario...... 198

Figure 8.1: Summary of thesis findings on biotic and abiotic interactions in an arid resource- pulse environment. This trophic interaction model highlights how interactions are context- specific and can change over space and time. It incorporates both bottom-up and top-down processes, but the strength of these processes can switch. Grey arrows are possible future interactions to explore. Artwork by A. Foster...... 214

viii

List of tables

List of tables

Table 2.1: Quantile regression results for minimum temperature change over time (years), Simpson Desert, central Australia. In each row a significant P-value indicated by * implies that there is a non-zero slope for that quantile. For each weather station (Boulia, Alice Springs, Birdsville, Oodnadatta) the slope estimate for each of the extreme quantiles (0.5, 0.1, 0.9, and 0.95) was compared to the median quantile (0.5). Overlaps in the confidence intervals (CI) indicated similar rates of change for all comparisons...... 43

Table 2.2: Quantile regression results for maximum temperature change over time (years), Simpson Desert, central Australia. In each row a significant P-value indicated by * implies that there is a non-zero slope for that quantile. For each weather station (Boulia, Alice Springs, Birdsville, Oodnadatta) the slope estimate for each of the extreme quantiles (0.5, 0.1, 0.9, and 0.95) was compared to the median quantile (0.5). Overlaps in the confidence intervals (CI) indicated similar rates of change for all comparisons...... 44

Table 2.3: Quantile regression results for annual rainfall changes over time (years) for each weather station, Simpson Desert, central Australia. The estimates in bold indicate that the slope does not overlap the confidence interval (CI) for the median slope (0.5 quantile). Significant P-value indicated by *...... 47

Table 2.4: Quantile regression results for intra-annual rainfall coefficient of variation for weather stations in the Simpson Desert, central Australia. Significant P-value indicated by *...... 49

Table 2.5: Piecewise regression results for the response of small mammal captures to annual rainfall from the preceding year, Ethabuka Reserve, Simpson Desert, central Australia. Small mammal captures were standardised per 100 trap nights (TN) and averaged per year (n = 22 years). Significant P-value indicated by *...... 51

Table 2.6: Piecewise regression results for the response of small mammal captures to annual minimum and maximum temperatures, Ethabuka Reserve, Simpson Desert, central Australia. Small mammal captures were standardised per 100 trap nights (TN) and averaged per year (n = 22 years). Significant P-value indicated by *...... 52

Table 3.1: Year and location of each long-haired rat, Rattus villosissimus, irruption from historical records. ESA AZRC = Ecological Society of Australia Arid Zone Research Chapter...... 78

Table 3.2: Analysis of deviance results for the effect of distance to nearest drainage line on live-trapping captures of the long-haired rat, Rattus villosissimus, Simpson Desert, central Australia in 2011. A negative binomial generalized linear model was used, with log-link function. Deviance follows the chi-square distribution. * indicates significance P<0.05...... 80

Table 3.3: Analysis of variance results for the effect of direction (angle) to nearest drainage line on live-trapping captures of the long-haired rat, Rattus villosissimus, Simpson Desert, central Australia. * indicates significance P<0.05...... 82

Table 3.4: ANOVA results for the influence of age class (adult or sub-adult) and site on captures of the long-haired rat, Rattus villosissimus, Simpson Desert, central Australia. *indicates significance P<0.05...... 84

ix

List of tables

Table 3.5: Analysis of deviance results for the influence of sex (male or female) and site on captures of the long-haired rat, Rattus villosissimus, Simpson Desert, central Australia...... 84

Table 3.6: Analysis of deviance results for reproductive condition of female and male long- haired rats, Rattus villosissimus. Reproductive condition non-parous, parous and pregnant/lactating female captures and non-scrotal, semi-scrotal and scrotal male captures in May 2011. *indicates significance P<0.05...... 85

Table 3.7: ANOVA results investigating if body mass of the long-haired rat, Rattus villosissimus, differs across trips and between sex, Simpson Desert, central Australia. *indicates significance P<0.05...... 86

Table 4.1: AICc values for MARSS models describing five possible spatial sub-population structures, with and without density dependence, for five species of small mammals in the Simpson Desert, central Australia. Data are based on 17–22 years of live-trapping from nine sites. Smallest AICc, highlighted in bold, indicates the best-fitting model...... 109

Table 4.2: Model estimates (95% confidence intervals) for best fitting MARSS population models (lowest AICc) for captures of small mammals (100 trap nights, log +1 transformed), Simpson Desert, central Australia. Density dependence occurs if the diagonal element of B (Bii) is less than one. There is only one B element, as density dependence was assumed to be equal for each sub-population...... 110

Table 4.3: The results of covariate MARSS models for captures of small mammals (100 trap nights, log +1 transformed), Simpson Desert, central Australia. Covariates were entered into models as their own process, and the coefficients from the off-diagonals of the B matrix measure the strength of their interaction with small mammal population growth rates. Covariates were z-scored so that direct comparisons can be made. Density dependence occurs if the diagonal element Bii is less than one. There is only one B element, as density dependence was assumed to be equal for each sub-population. Each covariate was considered significant if the 95% confidence intervals (CI) did not cross zero...... 111

Table 5.1: AICc values for MARSS models describing five possible spatial sub-population structures, with and without density dependence, for six species of reptiles in the Simpson Desert, central Australia. Data are based on 17–22 years of live-trapping from nine sites. Smallest AICc, highlighted in bold, indicates the best-fitting model...... 141

Table 5.2: Model estimates (95% confidence intervals) for best fitting MARSS population models (lowest AICc) for captures of reptiles (100 trap nights, log +1 transformed), Simpson Desert, central Australia. Density dependence occurs if Bii < 1...... 142

Table 5.3: The results of the covariate MARSS models for captures of reptiles (100 trap nights, log +1 transformed), Simpson Desert, central Australia. Covariates were entered into models as their own process, and the coefficients from the off-diagonals of the B matrix measure the strength of their interaction with reptile population growth rates. The diagonals of the B matrix indicate the strength of density dependence, which occurs when Bii < 1. Covariates were z-scored so that direct comparisons can be made. Each covariate was considered significant if the 95% confidence intervals (CI) did not cross zero (in bold). .... 143

Table 6.1: GLM results for number of detections of in photographs from remote camera trapping of dingoes, red fox and feral cat in the Simpson Desert, southwestern

x

List of tables

Queensland. Cameras were active for two years and photographs were pooled for each phase each year. Phase is the population phase of prey (bust, boom or decline)...... 167

Table 7.1: Individual models for each node in the structural equation model. Trip was entered as a random factor to account for the repeated measures. Mean rainfall event size2 = Mean rainfall event size two trips prior...... 193

Table 8.1: Summary of the key findings and achievements of this thesis, and the research area they advance...... 213

xi

xii

Glossary

Glossary:

Abiotic interactions: interactions that are external to those within or between species; e.g. climate. Asynchronous populations: sub-populations that have different temporal trajectories. Biotic interactions: interactions within or between species populations. Bottom up: effects on species populations from biota in lower trophic levels (usually primary producers), or increases in available resources. Density dependence: factors that limit the size of species populations that are dependent on the number of individuals in the population. Density independence: factors that limit the size of species populations that are external to the number of individuals in the population. Extreme temperature or rainfall events: Weather events that lie within <10th quantile (extreme cold or dry) or >90th quantile (extreme hot or wet). Extrinsic factors: population drivers that are external to the interactions within populations. Intrinsic factors: interactions within populations that drive their dynamics. Mesopredators: predators whose populations are limited by other, usually larger predators. Moran effect or theorem: states that sub-populations with a common density dependent structure can be synchronized by a spatially correlated density independent factor, such as climate. Multivariate autoregressive state-space modelling (MARSS): a time-series modelling framework that allows one to investigate different sub-population structures and covariates that may be important in driving them, while also incorporating both process and observation variability. Observation variability: represents temporal variability in population size due to sampling error (e.g. temporal changes in detectability or error resulting in only a sub-sample of the population being counted). Piecewise regressions: regressions that use two or more regression lines, joined at a break point, which can be used to identify thresholds. Population boom and bust cycles: rapid irruptions of populations that can reach a many-fold increase in the starting population abundance, followed by a collapse back to the low- level starting population. Population dynamics: the study of the fluctuations in the abundance of individuals and the causes for them in species populations. Process variability: represents temporal variability in population size due to environmental and demographic stochasticity.

xiii

Glossary

Refugia: sites that provide resources that allow species populations to survive in times of resource shortage. Resource pulses: rapid short-term increase in a limiting population resource. Structural equation modelling (SEM): Probabilistic models containing or specifying multiple causal pathways. SEMs are characterized by attempting to satisfy the criteria for drawing causal inferences and permitting endogenous variables to be functions of other endogenous variables, thereby potentially containing indirect effects (see Grace et al. 2012). Synchronous populations: sub-populations that share the same trajectories. Top predators: Large predators in an ecosystem whose populations are self-limiting (see Wallach et al. 2015). Top-down: effects on species populations from higher trophic levels. Trophic interaction model: a conceptual framework developed for this thesis to explore the interactions between species and their environment.

References: Grace, J. B., D. R. Schoolmaster, G. R. Guntenspergen, A. M. Little, B. R. Mitchell, K. M. Miller, and E. W. Schweiger. 2012. Guidelines for a graph-theoretic implementation of structural equation modeling. Ecosphere 3: art73. Wallach, A. D., I. Izhaki, J. D. Toms, W. J. Ripple, and U. Shanas. 2015. What is an apex predator? Oikos In press.

xiv

Acknowledgments

Acknowledgments

I would first like to thank my two supervisors, Chris Dickman and Glenda Wardle,

who patiently waited for 10 years for me to come around to the idea of undertaking a PhD. I

first got seduced by the fauna of central Australia when Chris took me on to do a third year

research project. This cascaded into honours where I learnt first-hand about extreme rainfall events. My first trip, 15 years ago, ended in a helicopter airlift after rain made the roads impassable for months. Booms and busts were on display and witnessing how the flora and fauna responded was truly remarkable. I was hooked. Chris’ humble, yet cutting-edge approach to research, academia and passion for the Simpson Desert was a truly life-shaping experience.

I still remember my first meeting with Glenda when first employed as a Research

Assistant on a newly funded ARC grant that Glenda and Chris had successfully obtained.

Glenda took the time to sit me down and explain the grant and its experimental design. The novel project and design sparked more interest in experimental design and statistics and

Glenda was always there to help and encourage me on this journey. It was a journey that has led to this thesis. Glenda always encourages her students to do their best, and she encouraged me to grow my skills and interests in population biology and modelling. Always with an emphasis on good science and grounded in field data.

Special thanks to my wife, Alison Foster, for countless hours of love and support. For runs to the home brew shop for much needed “PhD supplies”, picking up my slack at home, dealing with tradies for the renovations, when I couldn’t. Thanks for all the proof reading and fantastic illustrations that accompany my work. I couldn’t have done it without your love and understanding.

xv

Acknowledgments

Special thanks to the Desert Ecology Research Group, Dickman and Wardle labs for all the support, coffee breaks, great discussions of ideas and issues. Special thanks to Bobby

Tamayo, Dave Nelson, Chin-Liang Beh, Nathan Emery, George Madani, Alan Kwok, Emma

Spencer, Tom Newsome, Eveline Rijksen, Jenna Bytheway, Max Tischler, Tony Popic,

Channing Hughes and Nicole Hills. Thanks to Vuong Nguyen for help with population modelling and discussions on population dynamics. Also to Mathew Crowther for some great and colourful “couch” breaks, discussions on many aspects of work, statistics and some great team efforts running workshops on R, statistics and GIS. Thanks also to Al Glen for some great discussions on predator ecology and the wine of the Marlborough region. A big thanks to Alan Kwok for proof reading the entire thesis!

For all things from pubs, photography, venting, frisbee, eating, coffee, tech support, bush regeneration at our home and adventures around town and bush, I would like to thank

Sam Travers, Yvonne Davila, Mike Taylor, Trevor Wilson, Marion Winkler, Nick Colman,

Miguel Bedoya-Perez, Nathan Emery, Alan Kwok and Bobby Tamayo. Thanks for keeping me sane.

Thanks to my family, both the Greenvilles and Fosters, for all your support over the years, and for my parents for giving me the gift of an education. They believed in me and always encouraged me to pursue my dreams, even if the job prospects were thin.

This research would not have been possible without the help of 900 plus volunteers over the last 25 years and also without funding, which was provided by the Australian

Research Council, an Australian Postgraduate Award, Paddy Pallin Grant, Royal Zoological

Society of NSW and the Australian Government’s Terrestrial Ecosystems Research Network

(TERN).

xvi

“The dogger and prospector follow the explorer; the survey party follows both and makes record of their findings; and hard upon their heels has been the stockman with his cattle, horses, donkeys, and camels, his sheep and goats and dogs; and the great hosts of the uninvited also – the rabbits, the foxes and the feral cats.

The results of all this are hailed by the statistician and economist as progress, and a net increase in the wealth of the country, but if the devastation which is worked to the flora and fauna could be assessed in terms of the value which future generations will put upon them, it might be found that our wool-clips, and beef and timber trades have been dearly bought”

— H.H. Finlayson, The Red Centre, 1935

xvii

xviii

Abstract

Thesis Abstract

A fundamental question in ecology is the relative role of the environment in shaping the distribution and abundance of species. Interactions among species also shape population dynamics and thus determine the composition of assemblages across space and time.

Resolving the boundaries, and interplay, between biotic (species–species) compared to abiotic (species–environment) influences, remains of theoretical and practical importance to ecologists and managers of wildlife populations. However, identifying important biotic and abiotic interactions is often difficult, particularly when multiple trophic levels are involved, but doing so can provide key evidence of altered environments. A major challenge faced by ecologists today is predicting responses of species’ populations in changing environments. To provide more accurate predictions about the fate of biological diversity and how losses might be managed, simultaneous analyses of multiple drivers and their interactions are needed. And such predictions are needed increasingly, as disturbances, such as those due to climate

change, become more frequent and intense. Nowhere is this research agenda more imperative

than in the world’s already-stressed arid regions.

Arid environments are characterised by low and unpredictable rainfall (< 250 mm a

year), which often may occur in large falls of short duration. These discrete events trigger

resource pulses in arid regions that contrast with the more-reliable and seasonal rainfalls

typical of many temperate regions. In addition to its temporal irregularity, rainfall is highly

unpredictable across space; hence the availability of resources in arid environments is

spatially and temporally variable. High resource variability provides a uniquely suitable

environmental template to investigate the role of biotic and abiotic interactions in regulating

ecological systems.

xix

Abstract

I tackled this fundamental question about the factors that influence populations in

several stages using the biota of arid Australia. I first document patterns in the main environmental drivers, rain and fire, and then analysed long time-series data on population dynamics of mammals and reptiles in relation to these drivers. Once the foundational patterns were established, I quantified trophic interactions among species to develop a new conceptual model for this system. I concluded that both biotic and abiotic interactions are important for driving populations in resource-pulse environments, and that interactions are context-specific and can change over space and time.

The first aim of my thesis was to quantify changes in the climate of central Australia over the last century. Thus, I investigated changes in median and extreme temperature and rainfall events at 11 weather stations around the Simpson Desert at two spatial scales: at local, individual weather stations, and at the Simpson Desert regional scale (Chapter 2). I found that median and extreme annual minimum and maximum temperatures have increased at both spatial scales over the past century, consistent with increases in global temperature over this period. Rainfall changes were inconsistent across the Simpson Desert; data from individual weather stations showed increases in annual rainfall, increased frequency of large rainfall events, or more prolonged droughts, depending on their location. Using long-term live-trapping records (22 years) of desert small mammals as a case study, I demonstrated that irruptions of rodents were driven by extreme rainfalls (>95th quantile) whereas small dasyurid marsupials were not responsive. I also found that extreme rainfall events successfully predict irruptions of the usually-rare long-haired rat (Rattus villosissimus), and provided new insight into how the rats invade from multiple sites that are used as refuges during drought (Chapter 3). As wildfires predictably follow extreme rainfalls, an increase in the magnitude and frequency of extreme rainfall events is likely to drive changes in

xx

Abstract

populations of these species through direct and indirect changes in predation pressure and

wildfires. The prediction concerning predation is tested in aim four, below.

The second aim of my thesis was to quantify the most significant biotic and abiotic

interactions driving the temporal and spatial population dynamics of desert small mammals

and reptiles. I used multivariate state-space models to investigate the spatial population

structures of the study species using 17–22 years of intensive live-trapping data from nine spatially distinct sites in central Australia. Sub-populations of rodents, reptiles and a carnivorous marsupial, the mulgara, were synchronous or had two structures if driven by large-scale processes such as rainfall and wildfire, but insectivorous dasyurid marsupials were asynchronous and structured by local events (Chapters 4 and 5). Density dependence was detected for the first time in all species, but was weakest in dasyurids and reptiles. These findings suggest that local environmental stochasticity is more important than intrinsic factors in driving dasyurid population dynamics, whereas populations of rodents, reptiles and the mulgara are driven by both extrinsic and intrinsic factors that operate at the landscape scale.

The third aim of my thesis was to first develop a conceptual framework that predicts how top-down and bottom-up interactions may regulate sympatric predator populations in environments that experience resource pulses (Chapter 6). I tested this model using two years of remote-camera trapping data to uncover spatial and temporal interactions between a top predator, the dingo Canis dingo, and the smaller, or mesopredatory, European red fox Vulpes vulpes and feral cat Felis catus, during population booms, declines and busts (troughs) in numbers of their prey (rodents). I found that dingoes predictably suppress the mesopredators and that the effects are strongest during declines and busts in prey numbers. Given that

resource pulses are usually driven by large yet infrequent rains, I concluded that top predators

like the dingo provide net benefits to prey populations by suppressing mesopredators during

prolonged bust periods when prey populations are low and potentially vulnerable.

xxi

Abstract

The fourth aim of my thesis was to incorporate all the key biotic and abiotic

interactions identified above into one modelling framework. The trophic interaction model

was intended to provide more-accurate predictions of population responses to climate change

by using simultaneous analyses of multiple drivers (wildfire and rainfall) and their interactive

effects, as well as interactions among species across trophic groups (Chapter 7). I used

structural equation modelling to integrate remote camera trapping and live-trapping of

vertebrates with long term (17–22 years) vegetation data from the Simpson Desert. I then used these models to predict how changes in rainfall and wildfire events, in-line with future climate scenarios, will percolate up the trophic levels and interact with top-down effects from mammalian carnivores over the next 100 years. Contrary to expectations, predicted increases in vegetation cover from increased rainfall will be offset by increases in wildfire frequency, leading to mean decreases in spinifex cover and, in turn, spinifex seed production over the next 100 years. However, I found no change in predicted rodent numbers with decreased spinifex seed under the climate change scenario. In line with expectations, however, when mammalian predators were removed from the model there was an increase in predicted captures of rodents that was greatest when only introduced predators (foxes and feral cats) were removed, suggesting strong top-down limitation of the system by introduced predators.

The conceptual framework and empirical results from my thesis have been useful for predicting direct and indirect interactions among existing species in Australia’s arid regions, but this framework does have limitations. Firstly, it does not account for novel interactions.

For example, invasive species such as European rabbit, Oryctolagus cuniculus, and house mouse, Mus musculus, are largely absent in the study region, but increases in the magnitude and frequency of large (>90th quantile) rainfall events may allow these species to establish.

The influence of these species on vegetation and predator populations should be the focus of

future research. Secondly, weak interactions may not be well represented in the modelling

xxii

Abstract

framework, but they can have important effects on the overall stability of food webs. Lastly, little is known about other important components of the ecosystem, such as nutrient cycling, invertebrates and interactions between reptiles, dasyurids and eutherian predators. These areas should be foci for further research.

xxiii

xxiv

Preface

This thesis is set out as a series of papers, an option available at The University of

Sydney. The papers have been published, accepted or submitted to appropriate peer-reviewed scientific journals in the form presented here or similar. As a consequence, there is some redundancy among chapters and journal-specific formatting.

Chapters that have been a published or accepted to peer-reviewed journals are indicated at the beginning of each chapter. My co-authors on these papers are Professors

Chris R. Dickman, Glenda M. Wardle and others where appropriate. My contribution to these papers was in each case very substantial, and included conceptualisation of ideas, engagement in fieldwork, completion of data manipulations and analyses, and writing and editing drafts of all the chapters. I am first author on each of the published and submitted papers. Professors Chris Dickman and. Glenda Wardle provided supervision of the project, including discussions on initial concepts, design and analyses of data, and improved earlier versions of the chapters. In addition, some chapters have been published as popular science articles for newsletters, blogs and magazines. A copy of these is included in the appendix.

As supervisors of this PhD thesis, we concur with the statements above about Mr

Greenville’s contributions.

Chris Dickman Glenda Wardle

xxv

Chapter 1: General Introduction

Chapter 1: General introduction

The sun sets over the in the Simpson Desert. Photograph by Aaron Greenville.

1

Chapter 1: General Introduction

Introduction

Ecological interactions are important in driving the dynamics of species populations and patterns of community assembly, but the relative importance of biotic (species–species)

compared to abiotic (species–environment) interactions has been much debated. Following

discussions in the early part of the twentieth century (e.g. Ross 1911, Elton 1927), debate

about key factors that influence population dynamics intensified during the 1950s and 1960s.

Nicholson (1954) argued that biotic factors, such as density dependence and species

interactions were more important than abiotic factors such as climate. He and other

proponents (e.g. MacArthur 1958, Lack 1966, 1967) argued that climate merely influenced

the initial abundance of populations and that internal population processes or species

interactions, such as competition, were responsible for the currently observed patterns of

abundance (Nicholson 1954). In contrast, Andrewartha and Birch (1954) stated that the

distributions and abundances of animals were influenced most strongly by climate or other

abiotic factors, and argued further that these usually density independent factors acted to

continually reset populations to low numbers. Their view was encapsulated in the title of their

seminal text “The distribution and abundance of animals” (Andrewartha and Birch 1954).

Based mostly on empirical evidence, Andrewartha and Birch (1954) suggested that

the environment of a species has four components: weather, food, species interactions and

habitat, and an organism’s chance of survival and—as a result—population size depended on

one or more of these components (Andrewartha and Birch 1954, Andrewartha and Birch

1960). They introduced the concept of relative shortage and described mechanisms for

population regulation that did not require density dependence (White 1978, Meats 2010). Up

until the publication of Andrewartha and Birch’s seminal text, the field of population biology

was dominated by thinking from theoretical models (Andrewartha and Birch 1954, Meats

2

Chapter 1: General Introduction

2010), often based on stable and regular systems. However these models did not accurately predict population changes in the field or in highly variable environments (e.g. Wiens 1977,

Caughley 1987, Krebs 1995, Krebs 2002a, b) and the study of population dynamics in

variable systems still has much to offer today.

Nonetheless, both sides of the debate had some common ground and recognised that

interactions could change in nature and intensity across time and space (Andrewartha and

Birch 1954, Nicholson 1954, MacArthur 1958, 1972). With the resurgence of niche theory in

the 1960s, following the seminal works of Hutchinson (1957, 1959), it became appreciated

increasingly that both biotic and abiotic factors were often important for driving species’

populations (James et al. 1984, Dunson and Travis 1991, Turchin 2003). Of course, the joint

action of biotic and abiotic factors complicates interpretation of population drivers (Harrison

and Cornell 2008), and how both biotic, and abiotic interactions drive populations and

species’ assemblages are still among the most fundamental questions that are addressed in

ecology (Baskett and Salomon 2010).

Complex systems with multiple species are difficult to study. Interactions can have

positive or negative effects on one or more of the species in an assemblage, resulting in

mutualism and facilitation (Baskett and Salomon 2010, Butterfield et al. 2010) or competition

and predation (Bouskila 1993), and can regulate assemblages through trophic interactions

(Johnson and VanDerWal 2009, Letnic et al. 2011a). Abiotic interactions, such as those

between organisms and the soil (Greenville and Dickman 2009) or organisms and climate

(Dickman et al. 1999a, Dickman et al. 1999b) are equally central in regulating populations

and assemblages, and are increasingly important to understand in a world that is experiencing

a rapidly changing climate.

3

Chapter 1: General Introduction

Interactions can be either direct or indirect. A direct interaction is one where a biotic

or abiotic factor affects one or more species (i.e. there is only one step between driver and

species), whereas an indirect interaction is one where the effects of a biotic or abiotic factor

on another species arise only in the presence of another species or abiotic factor (i.e. there are

more than two steps between driver and species) (Wootton 2002). The presence of these different kinds of interactions leads to added complexity in ecological systems, and may have strong current, as well as evolutionary effects on species’ populations and assemblage structure (Werner and Peacor 2003, Walsh 2013). For example, the presence of shrubs has a

significant positive effect on European lynx abundance, but only when rabbits are present

(Grace 2008). Similarly, habitat diversity has an effect only on habitat generalist butterflies

via a positive effect on host-plant richness (Menéndez et al. 2007). Such interactions may

also show spatial and temporal variability due to the addition or removal of species (Díaz-

Castelazo et al. 2010), seasonal changes (Johnson et al. 2009), disturbance, or differing

climatic and nutrient patterns across space (Winemiller 1990).

Ecological interactions also are often assumed to be linear and to increase in intensity

with increasing frequencies of contact between individuals, but some research suggests that

linearity may not always occur and that thresholds instead may be important. Within species,

for example, the Allee effect describes situations where population growth rates are slow

until populations reach a density threshold, at which point growth rates accelerate until

traditional (negative) density effects become manifest (Allee 1931, Møller and Legendre

2001). Threshold effects can be important also in interactions between species or between

species and the environment. For example, rainfall may influence a population only when it

reaches a certain volume or falls in a certain season (Hodgkinson and Müller 2005, López et

al. 2008). Dominant predators may suppress sub-ordinate predators only after they reach a

certain density (Letnic et al. 2011a) or populations may aggregate after a certain percentage

4

Chapter 1: General Introduction of the landscape contains their preferred habitat (With and Crist 1995). These thresholds may differ both spatially and temporally, or may be reached only at certain times or locations.

Climate change and extreme events

Global surface temperatures have increased by an average of 0.85 °C since 1880 and are projected to increase by at least 1.5 °C towards the end of the 21st century (IPCC 2014).

Changes to mean rainfall patterns are projected to be region-specific, with increases in mid- latitude areas in the northern hemisphere, but contrasting changes in other latitudes (IPCC

2014). High latitudes and mid-latitude wet regions are projected to experience an increase in mean rainfall and mid-latitude and subtropical dry regions are projected to experience decreases (IPCC 2014). Changes in climate may increase extinction risk, especially for life in regions with flat landscapes, such as Australia, which provide limited opportunity for species to shift their ranges to places with more favourable climates (Hughes 2011, IPCC 2014).

Climate extremes, such as heat waves, droughts, floods, cyclones and wildfires, have revealed significant vulnerability of some ecosystems and many human systems to current climate change, and these events are also likely to increase in frequency and intensity across the globe (IPCC 2014, Maron et al. 2015). These infrequent but often catastrophic events may play key roles in regulating ecosystems, but understanding how event-driven dynamics govern patterns in biological diversity remains an elusive goal (Jentsch et al. 2011, Smith

2011). We can expect fluctuations in climate to have direct effects on populations of organisms, but there are also likely to be many subtle or indirect effects, mediated through bottlenecks in resource availability or shifts in competitive or predatory relationships, that will further influence natural systems in complex ways (Brown et al. 1997, Smith et al. 1998,

Epps et al. 2004, Butt et al. 2015).

5

Chapter 1: General Introduction

Importance of including multiple population drivers

Identifying biotic interactions that are important population or community drivers can

be difficult, particularly when multiple trophic levels are involved, but can also be useful for

detecting evidence of environmental alteration (Popic and Wardle 2012). Under a changing

climate, it has become recognised increasingly that many, if not most, species interactions are

climate-dependent due to changes in the phenology, activity patterns, and reproduction rates

of the interactants (Tylianakis et al. 2008, Gilman et al. 2010). For example, small shifts in

plant phenology may have large effects on mutualisms, such as those underpinning plant-

pollinator networks (Memmott et al. 2007). Extreme weather events, such as droughts, can

disrupt flowering phenology, leading to declines or even local extinctions of pollinators,

which in turn can have cascading effects on frugivores and seed dispersal (Harrison 2000,

2001).

Most studies have investigated non-interactive effects of global climate change across trophic levels, and have also focused on single drivers of species’ population dynamics

(Tylianakis et al. 2008). Other studies have focused on the ecology of individual species, via

habitat or niche modelling and attributes such as dispersal capacities, while ignoring biotic

interactions and evolutionary considerations (Lavergne et al. 2010). In order to provide more

accurate predictions, simultaneous analyses of multiple drivers are needed to investigate the

interactive effects of changes in populations from both biotic and abiotic processes.

The timing in the importance of bottom-up and top-down interactions can change due to abiotic factors (Sinclair and Krebs 2002). For example, Letnic et al. (2011b) found that the abundance and species richness of small mammals increases with increasing primary productivity and that the abundance of small mammals decreased with increasing predator

6

Chapter 1: General Introduction

activity. Predator abundances lagged behind those of small mammals and exerted top-down control only after bottom-up effects weakened, illustrating that species interactions can be

(and probably often are) context-dependent (Letnic et al. 2011b, Chamberlain et al. 2014).

Interactions between abiotic factors are also important for predicting species- interactions and for modifying the habitat templet (sensu Southwood 1977) upon which the interactions take place. For example, global phases of the El Nino-Southern Oscillation cycle drive differing patterns of wildfires across the Americas and Oceania (Le Page et al. 2008,

Pastro et al. 2014). Wildfires, in turn, simplify habitat structure and provide more-open

hunting grounds for predators, increasing per capita predation on prey (McGregor et al.

2014). Thus simultaneous analysis of the effects of rainfall and wildfire is important when

predicting the dynamics of species’ populations, as well as additional biotic factors that might

be operative.

Spatial and temporal population dynamics

Understanding the temporal and spatial dynamics of species’ populations is important

both to uncover the interactions that influence demography and to identify populations that

are vulnerable to extinction (Ward et al. 2010). The causes of temporal variation in

population size have been much studied (Dickman et al. 1999b, 2001, Shenbrot et al. 2010),

but in widely distributed species these drivers may vary across space so that distinct sub-

population structures occur regionally or across the geographical range (Hinrichsen and

Holmes 2009, Ward et al. 2010). Spatial sub-structuring of populations may be dictated

extrinsically by the local environment or by factors intrinsic to local populations; sub- structuring may persist even in meta-populations where dispersal occurs (Ranta et al. 2006).

7

Chapter 1: General Introduction

In contrast, if a driver is common and widespread, all sub-populations may be synchronized

over large regional areas; this is the Moran effect (Moran 1953, Bjørnstad et al. 1999).

How biotic and abiotic factors interact to influence species’ population dynamics is

important for understanding spatial synchrony. For example, biotic factors, such as density

dependence, commonly influence population growth rate at local scales (Bateman et al.

2011). Local environmental conditions, such as access to drought refugia (Dickman et al.

2011), can drive population dynamics such that sub-populations are independent from each

other. However, multiple sub-populations can also interact with each other if they are close

enough and share similar environmental conditions. For example, populations that fall within

the same rainfall zone may experience similar levels of productivity (Woodman et al. 2006),

or populations that experience the same disturbance event, such as wildfire (Letnic et al.

2004, Evans et al. 2010), may have similar population trajectories. Lastly, if a driver such as

a widespread climatic event, predation or dispersal is experienced across a species’ range, all

sub-populations may exhibit the same dynamics (Moran 1953, Bjørnstad et al. 1999, Ranta et

al. 2003).

Resource pulse environments

Arid regions are characterised by low and unpredictable rain (<250 mm a year) that

often falls in discrete events (Noy-Meir 1973). These resource pulses typify arid regions and

often occur only over short time intervals (Nano and Pavey 2013). In addition to this

temporal variation, rainfall in arid regions usually is highly unpredictable across space (Noy-

Meir 1973, Westoby 1979). This can result in a temporal and spatial interaction with the

availability of resources in the environment (Noy-Meir 1973). How species interact with this variable environment and with each other is a fundamental question for arid zone ecology.

8

Chapter 1: General Introduction

Researchers often model the population dynamics of individual species with rainfall as one of

the parameters in their models. For example, Lima et al. (2008) found that summer rainfall

had a significant effect on the population sizes of two kangaroo rats, but

there was also a negative feedback on one of the species through competition. Dickman et al.

(1999a) found that rainfall positively influenced populations of two dragon species in the

Simpson Desert, but increases in the dominant plant cover resulting from rainfall had a

negative effect on the abundance of one of the dragon species over time. How such species

change over space, as well as time, is largely unknown. This is possibly due to the rarity of

long-term datasets that have been replicated both spatially and temporally.

Environments that experience resource pulses from rapid increases in primary

productivity, such as from extreme rainfall events (flood rains), show dramatic ‘booms’ and

‘busts’ in populations of primary consumer organisms (Caughley 1987, Yang et al. 2008,

Dickman et al. 2010, Yang et al. 2010, Wardle et al. 2013). Even though resource pulses may be infrequent, many consumers efficiently exploit them; for example, populations of some mammals increase 10–60 fold within six months via elevated reproduction and immigration

(Dickman et al. 1999b, 2010). This produces different phases in their population cycles, with long busts, or periods of low numbers, and short-lived booms followed by rapid declines (see

Chapter 6, Fig. 6.1).

Thesis conceptual model

In a recent synthesis by Morton et al. (2011) the key ecological processes that inform our understanding of arid Australia were outlined, and four key abiotic factors were highlighted: infertile soils that are well-sorted, unpredictable rainfall, rare and extreme rainfall events, and wildfire. Infertile soils and unpredictable rainfall lead to high

9

Chapter 1: General Introduction

unpredictability in the arrival of resource pulses and low digestibility of plant matter (Morton et al. 2011). Thus, consumers are dominated by detritivores and granivores, with life-history traits that can take advantage of irregular pulses in plant production (Stafford Smith and

Morton 1990, Morton et al. 2011). Nonetheless, some higher-level consumers such as reptiles have become extremely diverse in arid environments, especially in central Australia, due to their lower energy requirements. These ectotherms can take advantage of consumers, such as termites that may be available only transiently (Pianka 1985, Morton and James 1988). In addition to the irruptive population dynamics of some vertebrates, such as birds and mammals, these consumers often exhibit great dispersal ability (relative to their size) and can move towards areas that have received recent rain (Dickman et al. 1995, Letnic 2002,

Haythornthwaite and Dickman 2006b). Thus space is also an important component in

understanding arid systems (Fig. 1.1), and a critical element to consider when interpreting

population and community dynamics and the factors that shape them.

Scarcity of water generally limits primary production in arid environments and

rainfall usually falls in discrete events (Noy-Meir 1973). However, it is proposed that extreme rainfall events, which can be spatially variable, are needed to trigger resource-pulses in vegetation growth and seed output. In turn, subsequent increases in consumer populations will follow (Letnic and Dickman 2006).

Paradoxically, extreme rainfall events increase the chance of wildfires occurring by stimulating growth in vegetation cover which, after the cover has dried, increases fuel loads

(Letnic and Dickman 2006, Greenville et al. 2009, Nano et al. 2012). In turn, wildfires can reduce shelter for cover-dependent species or expose consumers to increased predation pressure (Sutherland and Dickman 1999, Torre and Díaz 2004). For example, micro- carnivores may be associated with vegetation cover and use the resources provided within

10

Chapter 1: General Introduction that cover (Haythornthwaite and Dickman 2006a), but also be limited by populations of larger sympatric micro-carnivores (Fig. 1.1, Dickman 2003).

Figure 1.1: Conceptual trophic interaction model that incorporates space as a potential driver of populations. It allows for bottom-up and top-down processes to be incorporated, as well as indirect interactions. The strengths of interactions and bottom-up or top-down effects are not assumed to be constant and may vary temporally. Adapted from Dickman et al. (2014). Artwork by A. Foster.

Irruptions of consumers, although brief, may drive populations of meso-predators

(mid-sized predators that are subordinate to larger top-predators) or allow them to invade from mesic areas that fringe many desert regions (Letnic and Dickman 2006, 2010). Dietary analysis suggests that mesopredators can exert considerable predation pressure on some consumer species, and that top-down effects may become strongly apparent at these times

11

Chapter 1: General Introduction

(Pavey et al. 2008, Letnic and Dickman 2010, Cupples et al. 2011, Spencer et al. 2014). In contrast, top predators may limit populations of mesopredators and convey an indirect benefit to small prey by reducing total predation pressure (Letnic and Koch 2010, Letnic et al. 2011a,

Ripple et al. 2013, Ripple et al. 2014).

This thesis develops a trophic interaction model that incorporates biotic and abiotic interactions as potential drivers of species’ population dynamics at local and landscape levels

(Fig. 1.1). It allows for bottom-up and top-down processes to be incorporated, as well as indirect interactions. The strengths of interactions and bottom-up or top-down effects are not

assumed to be constant and may vary temporally. The conceptual model builds on the work

of Morton et al. (2011) above and adapts the model used in Dickman et al. (2014) by

focusing on two key abiotic drivers, wildfire and rainfall. Biotic interactions between species

and trophic groups are limited to vegetation and vertebrates (mammals and reptiles) to

provide a test case of the model, although I acknowledge that further interactions will likely

involve other vertebrates, invertebrates and organisms such as soil microbiota. I then use the

model to explore how global climate change may affect rainfall and wildfire, and how

changes in these abiotic factors may influence biotic interactions between trophic levels (see

thesis aims below).

Study region: the Simpson Desert, central Australia

Australia is characterised by a variable climate and poor soils (Orians and Milewski

2007), which are most evident in the continent’s arid interior (Morton et al. 2011). Flooding

rains can fall at any time or place and lead to a boom in resources, but booms are short-term

and then are followed by long periods of dry conditions, which can result in a bust in

resources (see Fig. 1.2 for local rainfall records and Chapter 2 for long-term rainfall records).

12

Chapter 1: General Introduction

Australian are highly variable compared to deserts on other continents (Stafford Smith

and Morton 1990, Morton et al. 2011) and hence are ideal places to investigate spatial and

temporal variation in ecological interactions. Previous research has concentrated on direct

interaction models, such as the ecological refuge model (Morton 1990, Dickman et al. 2011),

the predator refuge model, state and transition models and the fire-mosaic model (Letnic and

Dickman 2010). No single model has adequately predicted arid zone assemblages

(reviewed in Letnic and Dickman 2010). Most likely, combinations of environmental, life

history and human influences interact, both directly and indirectly to regulate arid zone

assemblages (Stafford Smith and Morton 1990, Morton et al. 2011).

For the past 25 years, the Desert Ecology Research Group at The University of

Sydney has been collecting data on reptiles, small mammals, predators, climate and

vegetation from the Simpson Desert, Australia and joined the Long-Term Ecological

Research Network (LTERN) in 2012 (Dickman et al. 2014). Our study region is located within the greater hummock or spinifex—grasslands, dominated by species of Triodia.

Hummock grasslands cover approximately 25% of Australia (Fig. 1.3) and represent the most widely distributed biome on the continent (Allan and Southgate 2002). The Desert Ecology

Research Group operates over an 8000 km2 study region, running 17 sites (9–12 were

suitable for this thesis) comprising more than 100 trapping grids (36 pitfall traps over 1 ha)

(see Chapters 2 and 3 for map of sites). At each of these grids, reptile and small mammal data are recorded using pitfall live-trapping techniques (see Dickman et al. 1999a, 1999b) and

vegetation surveys are also conducted using plot-based cover methods. Remote camera traps

are used to record predator (cat, fox and dingo) activity. Climate data are recorded from

automatic weather stations set up at each site (Fig. 1.2). In addition, long-term (>100 years) rainfall records have been maintained by the Bureau of Meteorology for sites within 200 km of the study region, which can be used to investigate long-term trends (see Chapter 2). High

13

Chapter 1: General Introduction native animal diversity occurs throughout the study region, with some of the most diverse dasyurid marsupial and reptile communities in the world. Invasive species, such as the

European rabbit, Oryctolagus cuniculus, and house mouse, Mus musculus, are rare, appearing only occasionally after flood rain (Pianka 1985, Downey and Dickman 1993, Dickman 2003,

Letnic and Dickman 2006). The high diversity and extensive long-term, spatially replicated datasets have allowed me in this thesis to explore the temporal and spatial patterns of interactions between biotic and abiotic factors that drive populations of the study biota.

14

Chapter 1: General Introduction

1000 Carlo 900 Field River North Field River South 800 Kunnamuka Swamp East 700 Main Camp 600 Shitty Site South Site 500 Tobermorey East 400 Tobermorey West 300 200 (mm) rainfall Annual 100 0 1990 1995 2000 2005 2010

Figure 1.2: Annual rainfall at nine sites surveyed for biotic and abiotic interactions in the Simpson Desert, central Australia. The timing of extreme rainfall events is similar across sites, but there is otherwise large spatial variation in rainfall events. Data were recorded at site-based automatic weather stations (Environdata, Warwick, Queensland), installed in 1995.

Figure 1.3: Extent of spinifex (hummock) grasslands in Australia, as mapped by Department of the Environment and Water Resources (2007) and location of the Simpson Desert. The study region (insert) encompasses four properties: Tobermorey, Cravens Peak, Carlo and Ethabuka. Detail of sampling sites on these properties is shown in Fig. A2.1.

15

Chapter 1: General Introduction

Thesis aims

The specific aims of this thesis and chapters that they are addressed in are:

1. To characterise temporal trends in extreme rainfall and temperature events across

space. In addition, to investigate whether extreme weather events are important for

driving species populations (Chapters 2 and 3);

2. To investigate the population dynamics and spatial sub-population structure of small

mammals and reptiles species. In addition, to identify important intrinsic and extrinsic

factors that regulate species’ populations (Chapters 4 and 5);

3. To identify spatial and temporal interactions between sympatric mammalian predators

(Chapter 6);

4. To test the conceptual model (Fig. 1.1) by incorporating the important species

interactions identified above and, in-line with climate-change predictions, investigate

the effects of increases in wildfire frequency and magnitude of rainfall events on

species interactions between multiple trophic levels (Chapter 7).

Overview of thesis

This chapter (Chapter 1) provides a general introduction into the relevant disciplines used in this thesis and the conceptual model used to investigate spatio-temporal biotic and abiotic interactions that are thought to occur in resource-pulse environments. This chapter then established the general aims of thesis. I will now outline the structure of the thesis and how each aim is addressed.

16

Chapter 1: General Introduction

The first aim of my thesis was to quantify century-long changes in the climate of central Australia (Chapter 2). Thus, I investigated changes in the median and extreme temperature and rainfall events at 11 weather stations around the Simpson Desert at two spatial scales: local (individual weather stations) and regional scale (Simpson Desert). I then

showed that native rodent populations irrupt after an extreme rainfall threshold (>95th

quantile) is reached and also that in contrast, dasyurids showed no response. Chapter 2 has

been published in Ecology and Evolution (see Greenville et al. 2012). I then provide another

case study and investigate the population dynamics of a native plague rodent, the long-haired

rat (Rattus villosissimus), and show that this species irrupts after extreme rainfall events and

can rapidly colonize new sites (Chapter 3). This chapter has been published in a special issue

of Austral Ecology: Greening of arid Australia: new insights from extreme years (see

Greenville et al. 2013). As wildfires predictably follow extreme rainfall events (Letnic and

Dickman 2006, Greenville et al. 2009), an increase in the magnitude and frequency of

extreme rainfall events is likely to drive changes in populations of the long-haired rat and

other native rodent species through direct and indirect changes in predation pressure and

wildfires. The prediction concerning predation is tested in aim four, below (Chapter 7).

The second aim of my thesis was to quantify the most significant biotic and abiotic

interactions driving the temporal and spatial population dynamics of desert small mammals

and reptiles. Firstly, the spatial sub-population structure is investigated using the novel

multivariate autoregressive state-space modelling (MARSS) technique and then biotic and

abiotic covariates are added to the model. Using Moran’s theorem I predict that species with

synchronous spatial dynamics are driven by factors that operate at the landscape scale, such

as resource-pulses from large-scale rainfall events or wildfires, whereas species with

asynchronous sub-populations will be influenced largely by factors operating at local scales

(Chapters 4 and 5).

17

Chapter 1: General Introduction

The third aim of my thesis was to first develop a conceptual framework that predicts

how top-down and bottom-up interactions may regulate sympatric predator populations in

environments that experience resource pulses (Chapter 6). Then I tested this model using two years of remote-camera trapping data to uncover spatial and temporal interactions between a top predator, the dingo Canis dingo, and the smaller, mesopredatory, European red fox

Vulpes vulpes and feral cat Felis catus, during population booms, declines and busts in numbers of their prey (rodents). These finding were published in Oecologia (see Greenville et al. 2014).

The fourth aim of my thesis was to incorporate all the key biotic and abiotic interactions identified above into one modelling framework and test the conceptual model provided in this first chapter (Fig. 1.1). I used structural equation modelling to find the significant interactions (pathways) in the conceptual model. Then I predicted population responses to climate change by using simultaneous analyses of multiple drivers (wildfire and rainfall), and investigated how the effects of these drivers would percolate up the trophic levels and interact with top-down effects from mammalian carnivores over the next 100 years

(Chapter 7).

Finally, in Chapter 8 I provide a summary discussion and general conclusions gained from my work and place my results in the context of the current literature. I then suggest areas of future research.

18

Chapter 1: General Introduction

References:

Allan, G. and R. Southgate. 2002. Fire regimes in the spinifex landscapes of Australia, Pages 145-176 in R. A. Bradstock, J. E. Williams, and A. M. Gill, editors. Flammable Australia: the fire regimes and biodiversity of a continent. Cambridge University Press, Cambridge. Allee, W. C. 1931. Animal Aggregations. University of Chicago Press, Chicago. Andrewartha, H. and L. Birch. 1960. Some recent contributions to the study of the distribution and abundance of insects. Annual Review of Entomology 5: 219-242. Andrewartha, H. G. and L. C. Birch. 1954. The distribution and abundance of animals. University of Chicago Press, Chicago, USA. Baskett, M. L. and A. K. Salomon. 2010. Recruitment facilitation can drive alternative states on temperate reefs. Ecology 91: 1763-1773. Bateman, A. W., A. Ozgul, T. Coulson, and T. H. Clutton-Brock. 2011. Density dependence in group dynamics of a highly social mongoose, Suricata suricatta. Journal of Animal Ecology: 628–639. Bjørnstad, O. N., R. A. Ims, and X. Lambin. 1999. Spatial population dynamics: analyzing patterns and processes of population synchrony. Trends in Ecology & Evolution 14: 427-432. Bouskila, A. 1993. Interactions between predation risk and competition: a field study of kangaroo rats and snakes. Ecology. 76: 165-178. Brown, J. H., T. J. Valone, and C. G. Curtin. 1997. Reorganization of an arid ecosystem in response to recent climate change. Proceedings of the National Academy of Sciences 94: 9729-9733. Butt, N., L. Seabrook, M. Maron, B. S. Law, T. P. Dawson, J. Syktus, and C. A. McAlpine. 2015. Cascading effects of climate extremes on vertebrate fauna through changes to low-latitude tree flowering and fruiting phenology. Global Change Biology: n/a-n/a. Butterfield, B. J., J. L. Betancourt, R. M. Turner, and J. M. Briggs. 2010. Facilitation drives 65 years of vegetation change in the . Ecology 91: 1132-1139. Caughley, G. 1987. Ecological relationships, Pages 159-187 in G. Caughley, N. Shepherd, and J. Short, editors. Kangaroos: their ecology and management in the sheep rangelands of Australia. Cambridge University Press, New York. Chamberlain, S. A., J. L. Bronstein, and J. A. Rudgers. 2014. How context dependent are species interactions? Ecology Letters: 881–890.

19

Chapter 1: General Introduction

Cupples, J. B., M. S. Crowther, G. Story, and M. Letnic. 2011. Dietary overlap and prey selectivity among sympatric carnivores: could dingoes suppress foxes through competition for prey? Journal of Mammalogy 92: 590-600. Department of the Environment and Water Resources. 2007. Australia’s native vegetation: a summary of Australia’s major vegetation groups. Canberra. Díaz-Castelazo, C., P. R. Guimarães, P. Jordano, J. N. Thompson, R. J. Marquis, and V. Rico-Gray. 2010. Changes of a mutualistic network over time: reanalysis over a 10- year period. Ecology 91: 793-801. Dickman , C., G. M. Wardle, J. Foulkes, and N. de Preu. 2014. Desert complex environments, Pages 379-438 in D. Lindenmayer, E. Burns, N. Y. Thurgate, and A. Lowe, editors. Biodiversity and environmental change: monitoring, challenges and direction. CSIRO Publishing, Vic. Dickman, C. R. 2003. Species interactions: indirect effects, Pages 158-170 in P. Attiwill and B. Wilson, editors. Ecology: an Australian perspective. Oxford University Press, Melbourne, Vic. Dickman, C. R., A. C. Greenville, C.-L. Beh, B. Tamayo, and G. M. Wardle. 2010. Social organization and movements of desert rodents during population “booms” and “busts” in central Australia. Journal of Mammalogy 91: 798-810. Dickman, C. R., A. C. Greenville, B. Tamayo, and G. M. Wardle. 2011. Spatial dynamics of small mammals in central Australian desert habitats: the role of drought refugia. Journal of Mammalogy 92: 1193-1209. Dickman, C. R., A. S. Haythornthwaite, G. H. McNaught, P. S. Mahon, B. Tamayo, and M. Letnic. 2001. Population dynamics of three species of dasyurid marsupials in arid central Australia: a 10-year study. Wildlife Research 28: 493-506. Dickman, C. R., M. Letnic, and P. S. Mahon. 1999a. Population dynamics of two species of dragon lizards in arid Australia: the effects of rainfall. Oecologia 119: 357-366. Dickman, C. R., P. S. Mahon, P. Masters, and D. F. Gibson. 1999b. Long-term dynamics of rodent populations in arid Australia: the influence of rainfall. Wildlife Research 26: 389-403. Dickman, C. R., M. Predavec, and F. J. Downey. 1995. Long-range movements of small mammals in arid Australia: implications for land management. Journal of Arid Environments 31: 441-452. Downey, F. J. and C. R. Dickman. 1993. Macro- and microhabitat relationships among lizards of sandridge desert in central Australia, Pages 133-138 in D. Lunney and D. 20

Chapter 1: General Introduction

Ayers, editors. Herpetology in Australia: a diverse discipline. Royal Zoological Society of New South Wales, Sydney, Australia. Dunson, W. A. and J. Travis. 1991. The role of abiotic factors in community organization. The American Naturalist 138: 1067-1091. Elton, C. 1927. Animal Ecology. Sidgwick and Jackson, London. Epps, C. W., D. R. McCullough, J. D. Wehausen, V. C. Bleich, and J. L. Rechel. 2004. Effects of climate change on population persistence of desert-dwelling mountain sheep in California. Conservation Biology 18: 102-113. Evans, M. E. K., K. E. Holsinger, and E. S. Menges. 2010. Fire, vital rates, and population viability: a hierarchical Bayesian analysis of the endangered Florida scrub mint. Ecological Monographs 80: 627-649. Gilman, S. E., M. C. Urban, J. Tewksbury, G. W. Gilchrist, and R. D. Holt. 2010. A framework for community interactions under climate change. Trends in Ecology & Evolution 25: 325-331. Grace, J. B. 2008. Structural equation modeling for observational studies. Journal of Wildlife Management 72: 14-22. Greenville, A. C. and C. R. Dickman. 2009. Factors affecting habitat selection in a specialist fossorial . Biological Journal of the Linnean Society 97: 531-544. Greenville, A. C., C. R. Dickman, G. M. Wardle, and M. Letnic. 2009. The fire history of an arid grassland: the influence of antecedent rainfall and ENSO. International Journal of Wildland Fire 18: 631-639. Greenville, A. C., G. M. Wardle, and C. R. Dickman. 2012. Extreme climatic events drive mammal irruptions: regression analysis of 100-year trends in desert rainfall and temperature. Ecology and Evolution 2: 2645-2658. Greenville, A. C., G. M. Wardle, and C. R. Dickman. 2013. Extreme rainfall events predict irruptions of rat plagues in central Australia. Austral Ecology 38: 754-764. Greenville, A. C., G. M. Wardle, B. Tamayo, and C. R. Dickman. 2014. Bottom-up and top- down processes interact to modify intraguild interactions in resource-pulse environments. Oecologia 175: 1349-1358. Harrison, R. D. 2000. Repercussions of El Niño: drought causes extinction and the breakdown of mutualism in Borneo. Proceedings of the Royal Society of London. Series B: Biological Sciences 267: 911-915. Harrison, R. D. 2001. Drought and the consequences of El Niño in Borneo: a case study of figs. Population Ecology 43: 63-75. 21

Chapter 1: General Introduction

Harrison, S. and H. Cornell. 2008. Toward a better understanding of the regional causes of local community richness. Ecology Letters 11: 969-979. Haythornthwaite, A. S. and C. R. Dickman. 2006a. Distribution, abundance, and individual strategies: a multi-scale analysis of dasyurid marsupials in arid central Australia. Ecography 29: 285-300. Haythornthwaite, A. S. and C. R. Dickman. 2006b. Long-distance movements by a small carnivorous marsupial: how Sminthopsis youngsoni (Marsupialia: Dasyuridae) uses habitat in an Australian sandridge desert. Journal of Zoology 270: 543-549. Hinrichsen, R. A. and E. E. Holmes. 2009. Using multivariate state-space models to study spatial structure and dynamics, in R. S. Cantrell, C. Cosner, and S. Ruan, editors. Spatial Ecology. CRC/Chapman Hall., London. Hodgkinson, K. C. and W. J. Müller. 2005. Death model for tussock perennial grasses: a rainfall threshold for survival and evidence for landscape control of death in drought. The Rangeland Journal 27: 105. Hughes, L. 2011. Climate change and Australia: key vulnerable regions. Regional Environmental Change 11: 189-195. Hutchinson, G. E. 1957. Concluding remarks. Cold Spring Harbor Symposia on Quantitative Biology 22: 415-427. Hutchinson, G. E. 1959. Homage to Santa Rosalia or why are there so many kinds of animals? The American Naturalist 93: 145-159. IPCC. 2014. Climate Change 2014: Synthesis Report. Contribution of Working Groups I, II and III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Core Writing Team, R.K. Pachauri and L.A. Meyer (eds.)]. IPCC, Geneva, Switzerland. James, F. C., R. F. Johnston, N. O. Wamer, G. J. Niemi, and W. J. Boecklen. 1984. The grinnellian niche of the wood thrush. The American Naturalist 124: 17-47. Jentsch, A., J. Kreyling, M. Elmer, E. Gellesch, B. Glaser, K. Grant, R. Hein, M. Lara, H. Mirzae, S. E. Nadler, L. Nagy, D. Otieno, K. Pritsch, U. Rascher, M. Schädler, M. Schloter, B. K. Singh, J. Stadler, J. Walter, C. Wellstein, J. Wöllecke, and C. Beierkuhnlein. 2011. Climate extremes initiate ecosystem-regulating functions while maintaining productivity. Journal of Ecology 99: 689-702. Johnson, C. N. and J. VanDerWal. 2009. Evidence that dingoes limit abundance of a mesopredator in eastern Australian forests. Journal of Applied Ecology 46: 641-646.

22

Chapter 1: General Introduction

Johnson, J. C., J. J. Luczkovich, S. P. Borgatti, and T. A. B. Snijders. 2009. Using social network analysis tools in ecology: Markov process transition models applied to the seasonal trophic network dynamics of the Chesapeake Bay. Ecological Modelling 220: 3133-3140. Krebs, C. J. 1995. Two paradigms of population regulation. Wildlife Research 22: 1-10. Krebs, C. J. 2002a. Beyond population regulation and limitation. Wildlife Research 29: 1-10. Krebs, C. J. 2002b. Two complementary paradigms for analysing population dynamics. Philosophical Transactions of the Royal Society of London B: Biological Sciences 357: 1211-1219. Lack, D. 1966. Population studies of birds. Clarendon Press, Oxford,. Lack, D. 1967. The natural regulation of animal numbers. 1st edition. Clarendon Press, Oxford,. Lavergne, S., N. Mouquet, W. Thuiller, and O. Ronce. 2010. Biodiversity and climate change: integrating evolutionary and ecological responses of species and communities. Annual Review of Ecology, Evolution, and Systematics 41: 321-350. Le Page, Y., J. M. C. Pereira, R. Trigo, C. da Camara, D. Oom, and B. Mota. 2008. Global fire activity patterns (1996-2006) and climatic influence: an analysis using the World Fire Atlas. Atmospheric Chemistry Physics 8: 1911-1924. Letnic , M. 2002. Long distance movements and the use of fire mosaics by small mammals in the Simpson Desert, central Australia. Australian Mammalogy 23: 125-134. Letnic, M. and C. R. Dickman. 2006. Boom means bust: interactions between the El Niño/Southern Oscillation (ENSO), rainfall and the processes threatening mammal species in arid Australia. Biodiversity and Conservation 15: 3847-3880. Letnic, M. and C. R. Dickman. 2010. Resource pulses and mammalian dynamics: conceptual models for hummock grasslands and other Australian desert habitats. Biological Reviews 85: 501-521. Letnic, M., C. R. Dickman, M. K. Tischler, B. Tamayo, and C. L. Beh. 2004. The responses of small mammals and lizards to post-fire succession and rainfall in arid Australia. Journal of Arid Environments 59: 85-114. Letnic, M., A. Greenville, E. Denny, C. R. Dickman, M. Tischler, C. Gordon, and F. Koch. 2011a. Does a top predator suppress the abundance of an invasive mesopredator at a continental scale? Global Ecology and Biogeography 20: 343-353.

23

Chapter 1: General Introduction

Letnic, M. and F. Koch. 2010. Are dingoes a trophic regulator in arid Australia? A comparison of mammal communities on either side of the dingo fence. Austral Ecology 35: 167-175. Letnic, M., P. Story, G. Story, J. Field, O. Brown, and C. R. Dickman. 2011b. Resource pulses, switching trophic control, and the dynamics of small mammal assemblages in arid Australia. Journal of Mammalogy 92: 1210-1222. Lima, M., S. K. M. Ernest, J. H. Brown, A. Belgrano, and N. C. Stenseth. 2008. Chihuahuan Desert kangaroo rats: nonlinear effects of population dynamics, competition and rainfall. Ecology 89: 2594-2603. López, B. C., M. Holmgren, S. Sabaté, and C. A. Gracia. 2008. Estimating annual rainfall threshold for establishment of tree species in water-limited ecosystems using tree-ring data. Journal of Arid Environments 72: 602-611. MacArthur, R. H. 1958. Population ecology of some warblers of northeastern coniferous forests. Ecology 39: 599-619. MacArthur, R. H. 1972. Geographical ecology; patterns in the distribution of species. Harper & Row, New York. Maron, M., C. A. McAlpine, J. E. M. Watson, S. Maxwell, and P. Barnard. 2015. Climate- induced resource bottlenecks exacerbate species vulnerability: a review. Diversity and Distributions 21: 731-743. McGregor, H. W., S. Legge, M. E. Jones, and C. N. Johnson. 2014. Landscape management of fire and grazing regimes alters the fine-scale habitat utilisation by feral cats. PLoS ONE 9: e109097. Meats, A. 2010. Back to Birch, accessed 1/02/2015, http://sydney.edu.au/science/biology/alumni/back_to_birch/alan-meats-story.shtml Memmott, J., P. G. Craze, N. M. Waser, and M. V. Price. 2007. Global warming and the disruption of plant–pollinator interactions. Ecology Letters 10: 710-717. Menéndez, R., A. González-Megías, Y. Collingham, R. Fox, D. B. Roy, R. Ohlemüller, and C. D. Thomas. 2007. Direct and indirect effects of climate and habitat factors on butterfly diversity. Ecology 88: 605-611. Møller, A. P. and S. Legendre. 2001. Allee effect, sexual selection and demographic stochasticity. Oikos 92: 27-34. Moran, P. 1953. The statistical analysis of the Canadian lynx cycle. Australian Journal of Zoology 1: 291-298.

24

Chapter 1: General Introduction

Morton, S. R. 1990. The impact of European settlement on the vertebrate animals of arid Australia: a conceptual model. Proceedings of the Ecological Society of Australia 16: 201-213. Morton, S. R. and C. D. James. 1988. The diversity and abundance of lizards in arid Australia: A new hypothesis. American Naturalist 132: 237-256. Morton, S. R., D. M. Stafford Smith, C. R. Dickman, D. L. Dunkerley, M. H. Friedel, R. R. J. McAllister, J. R. W. Reid, D. A. Roshier, M. A. Smith, F. J. Walsh, G. M. Wardle, I. W. Watson, and M. Westoby. 2011. A fresh framework for the ecology of arid Australia. Journal of Arid Environments 75: 313-329. Nano, C. E., P. Clarke, and C. R. Pavey. 2012. Fire regimes in arid hummock grasslands and Acacia shrublands, Pages 195-214 in R. A. Bradstock, A. M. Gill, and R. J. Williams, editors. Flammable Australia: fire regimes, biodiversity and ecosystems in a changing world. CSIRO Publishing, Vic. Nano, C. E. M. and C. R. Pavey. 2013. Refining the ‘pulse-reserve’ model for arid central Australia: Seasonal rainfall, soil moisture and plant productivity in sand ridge and stony plain habitats of the Simpson Desert. Austral Ecology 38: 741–753. Nicholson, A. 1954. An outline of the dynamics of animal populations. Australian Journal of Zoology 2: 9-65. Noy-Meir, I. 1973. Desert ecosystems: environment and producers. Annual Review of Ecology and Systematics 4: 25-51. Orians, G. H. and A. V. Milewski. 2007. Ecology of Australia: the effects of nutrient-poor soils and intense fires. Biological Reviews 82: 393-423. Pastro, L. A., C. R. Dickman, and M. Letnic. 2014. Fire type and hemisphere determine the effects of fire on the alpha and beta diversity of vertebrates: a global meta-analysis. Global Ecology and Biogeography 23: 1146-1156. Pavey, C. R., S. R. Eldridge, and M. Heywood. 2008. Population dynamics and prey selection of native and introduced predators during a rodent outbreak in arid Australia. Journal of Mammalogy 89: 674-683. Pianka, E. 1985. Some intercontinental comparisons of desert lizards. National Geographic Research 1: 490-504. Popic, T. J. and G. M. Wardle. 2012. Extremes: understanding flower – visitor interactions in a changing climate., Pages 99-106 in D. Lunney and P. Hutchings, editors. Wildlife and climate change: towards robust conservation strategies for Australian fauna. Royal Zoological Society of New South Wales, Mosman, NSW, Australia. 25

Chapter 1: General Introduction

Ranta, E., P. Byholm, V. Kaitala, P. Saurola, and H. Lindén. 2003. Spatial dynamics in breeding performance of a predator: the connection to prey availability. Oikos 102: 391-396. Ranta, E., V. Kaitala, and P. Lundberg. 2006. Ecology of populations. Cambridge University Press, Cambridge. Ripple, W. J., J. A. Estes, R. L. Beschta, C. C. Wilmers, E. G. Ritchie, M. Hebblewhite, J. Berger, B. Elmhagen, M. Letnic, M. P. Nelson, O. J. Schmitz, D. W. Smith, A. D. Wallach, and A. J. Wirsing. 2014. Status and ecological effects of the world's largest carnivores. Science 343: 1241484. Ripple, W. J., A. J. Wirsing, C. C. Wilmers, and M. Letnic. 2013. Widespread mesopredator effects after wolf extirpation. Biological Conservation 160: 70-79. Ross, R. 1911. The prevention of Malaria. 2nd edition. John Murray, Waterloo, London. Shenbrot, G., B. Krasnov, and S. Burdelov. 2010. Long-term study of population dynamics and habitat selection of rodents in the Desert. Journal of Mammalogy 91: 776- 786. Sinclair, A. R. and C. J. Krebs. 2002. Complex numerical responses to top-down and bottom- up processes in vertebrate populations. Philos Trans R Soc Lond B Biol Sci 357: 1221-1231. Smith, F. A., H. Browning, and U. L. Shepherd. 1998. The influence of climate change on the body mass of woodrats Neotoma in an arid region of New Mexico, USA. Ecography 21: 140-148. Smith, M. D. 2011. An ecological perspective on extreme climatic events: a synthetic definition and framework to guide future research. Journal of Ecology 99: 656-663. Southwood, T. R. E. 1977. Habitat, the Templet for Ecological Strategies? Journal of Animal Ecology 46: 337-365. Spencer, E. E., M. S. Crowther, and C. R. Dickman. 2014. Diet and prey selectivity of three species of sympatric mammalian predators in central Australia. Journal of Mammalogy 95: 1278-1288. Stafford Smith, D. M. and S. R. Morton. 1990. A framework for the ecology of arid Australia. Journal of Arid Environments 18: 255-278. Sutherland, E. F. and C. R. Dickman. 1999. Mechanisms of recovery after fire by rodents in the Australian environment: a review. Wildlife Research 26: 405-419. Torre, I. and M. Díaz. 2004. Small mammal abundance in Mediterranean post-fire habitats: a role for predators? Acta Oecologica 25: 137-142. 26

Chapter 1: General Introduction

Turchin, P. 2003. Complex population dynamics: a theoretical/empirical synthesis. Princeton University Press. Tylianakis, J. M., R. K. Didham, J. Bascompte, and D. A. Wardle. 2008. Global change and species interactions in terrestrial ecosystems. Ecology Letters 11: 1351-1363. Walsh, M. R. 2013. The evolutionary consequences of indirect effects. Trends in Ecology & Evolution 28: 23-29. Ward, E. J., H. Chirakkal, M. González-Suárez, D. Aurioles-Gamboa, E. E. Holmes, and L. Gerber. 2010. Inferring spatial structure from time-series data: using multivariate state-space models to detect metapopulation structure of California sea lions in the Gulf of California, Mexico. Journal of Applied Ecology 47: 47-56. Wardle, G. M., C. R. Pavey, and C. R. Dickman. 2013. Greening of arid Australia: New insights from extreme years. Austral Ecology: 731-740. Werner, E. E. and S. D. Peacor. 2003. A review of trait-mediated indirect interactions in ecological communities. Ecology 84: 1083-1100. Westoby, M. 1979. Elements of a theory of vegetation dynamics in arid rangelands. Israel Journal of Botany 28: 169-194. White, T. C. R. 1978. The importance of a relative shortage of food in animal ecology. Oecologia 33: 71-86. Wiens, J. A. 1977. On competition and variable environments: populations may experience "ecological crunches" in variable climates, nullifying the assumptions of competition theory and limiting the usefulness of short-term studies of population patterns. American Scientist 65: 590-597. Winemiller, K. O. 1990. Spatial and temporal variation in tropical fish trophic networks. Ecological Monographs 60: 331-367. With, K. A. and T. O. Crist. 1995. Critical thresholds in species' responses to landscape structure. Ecology 76: 2446-2459. Woodman, J. D., J. E. Ash, and D. M. Rowell. 2006. Population structure in a saproxylic funnelweb spider (Hexathelidae: Hadronyche) along a forested rainfall gradient. Journal of Zoology 268: 325-333. Wootton, J. T. 2002. Indirect effects in complex ecosystems: recent progress and future challenges. Journal of Sea Research 48: 157-172. Yang, L. H., J. L. Bastow, K. O. Spence, and A. N. Wright. 2008. What can we learn from resource pulses. Ecology 89: 621-634.

27

Chapter 1: General Introduction

Yang, L. H., K. F. Edwards, J. E. Byrnes, J. L. Bastow, A. N. Wright, and K. O. Spence. 2010. A meta-analysis of resource pulse–consumer interactions. Ecological Monographs 80: 125-151.

28

Chapter 2: Extreme climatic events drive mammal irruptions

Chapter 2: Extreme climatic events drive mammal irruptions: regression analysis of 100-year trends in desert

rainfall and temperature

A storm cell brings rain to the Simpson Desert and also impedes our access to the study sites. Photograph by Aaron Greenville.

A version of this chapter is published as Greenville A.C., Wardle G.M., Dickman C.R. 2012. Extreme climatic events drive mammal irruptions: regression analysis of 100-year trends in desert rainfall and temperature. Ecology and Evolution 2: 2645–2658. My contribution was very substantial, and included conceptualisation of ideas from existing grants, engagement in fieldwork, completion of data manipulations and analyses, and writing and editing drafts of all the chapters, in consultation with my co-authors.

29

Chapter 2: Extreme climatic events drive mammal irruptions

Abstract

Extreme climatic events, such as flooding rains, extended decadal droughts and heat

waves have been identified increasingly as important regulators of natural populations.

Climate models predict that global warming will drive changes in rainfall and increase the

frequency and severity of extreme events. Consequently, to anticipate how organisms will

respond we need to document how changes in extremes of temperature and rainfall compare

to trends in the mean values of these variables and over what spatial scales the patterns are

consistent. Using the longest historical weather records available for central Australia—100

years—and quantile regression methods, we investigate if extreme climate events have

changed at similar rates to median events, if annual rainfall has increased in variability, and if

the frequency of large rainfall events has increased over this period. Specifically, we

compared local (individual weather stations) and regional (Simpson Desert) spatial scales,

and quantified trends in median (50th quantile) and extreme weather values (5th, 10th, 90th and

95th quantiles). We found that median and extreme annual minimum and maximum

temperatures have increased at both spatial scales over the past century. Rainfall changes

have been inconsistent across the Simpson Desert; individual weather stations showed

increases in annual rainfall, increased frequency of large rainfall events or more prolonged

droughts, depending on the location. In contrast to our prediction, we found no evidence that

annual rainfall had become more variable over time. Using long-term live-trapping records

(22 years) of desert small mammals as a case study, we demonstrate that irruptive events are driven by extreme rainfalls (>95th quantile) and that increases in the magnitude and frequency

of extreme rainfall events are likely to drive changes in the populations of these species

through direct and indirect changes in predation pressure and wildfires.

30

Chapter 2: Extreme climatic events drive mammal irruptions

Introduction

With the global climate changing (IPCC 2007), it is becoming increasingly important

to investigate how changes will manifest at local and regional scales that are relevant to

maintaining biodiversity (Chamaillé-Jammes et al. 2007, Peters et al. 2011). Beyond changes in average temperature and rainfall, extreme climate events such as flooding rain, droughts or heat waves will alter in frequency and severity (CSIRO and Australian Bureau of

Meteorology 2012). These infrequent but often catastrophic events may play key roles in regulating ecosystems, but understanding how event-driven dynamics govern diversity remains an elusive goal (Jentsch et al. 2011, Smith 2011a). We can expect fluctuations in climate to have direct effects on populations of organisms, but there are also likely to be many subtle or indirect effects, mediated through changes in resource availability or shifts in competitive or predatory relationships, that will further influence natural systems in complex ways (Brown et al. 1997, Smith et al. 1998, Epps et al. 2004).

Arid desert environments provide unique opportunities to explore the biotic effects of climatic extremes. Many deserts worldwide are characterized by extreme highs and lows in temperature, but the deserts in Australia are also distinctive for having highly variable and unpredictable rainfalls (Van Etten 2009, Morton et al. 2011) that produce “boom” and “bust” dynamics in constituent biota (Letnic and Dickman 2010). Organisms have varied means of surviving these conditions, such as below-ground resting stages and life history strategies that allow quick responses to extreme climate permutations (Noy-Meir 1973, Ward 2009, Morton et al. 2011). Temperatures in desert environments reflect and in some regions exaggerate the warming trend that has seen global temperatures rise 0.74° C since 1906 (IPCC 2007) and fluctuations also are likely to be amplified in future (Ward 2009, CSIRO and Australian

31

Chapter 2: Extreme climatic events drive mammal irruptions

Bureau of Meteorology 2012). Increases in temperature in arid environments, and the

resulting influence of this on wind patterns, can lead to changes in the rates of evaporation

(Donohue et al. 2010) and to more-subtle effects such as reductions in animal body mass that

have long-term consequences for the persistence of populations (Smith et al. 1998, Vincenzi

et al. 2012). As global temperatures rise, the capacity of the atmosphere to hold moisture

increases and circulation patterns also change (CSIRO and Australian Bureau of Meteorology

2007, IPCC 2007). Hence, changes in global rainfall may not reflect local rainfall changes

and, in arid regions, may potentially result in further increases in rainfall variability. Increases and decreases in rainfall have been reported in arid regions of North America, central

Australia and Africa in recent decades (Allan et al. 1996, Brown et al. 1997, CSIRO and

Australian Bureau of Meteorology 2007, Hoffman and Vogel 2008). For these reasons,

deserts provide ideal templates to investigate how organisms respond to climate extremes

over long periods. With improved understanding of the effects of climate fluctuations in the

recent past, we can then better anticipate how climate change may affect the biological

diversity and function of these hypervariable systems in future.

Already-high variability in rainfall and temperature in arid environments increases the

challenges of detecting changes in long-term trends (Chamaillé-Jammes et al. 2007). A

critical aspect of variability is changes in the maxima and minima for any parameter, as

organisms will have to survive these extremes even if slight changes in mean values are not

problematic. For example, temperatures exceeding 42° C cause death of flying foxes due to

heat stress (Welbergen et al. 2008); hot conditions during drought can even cull populations

of many specialist desert organisms (Fensham and Holman 1999, Low 2011, McKechnie et

al. 2012). In arid environments rainfall stimulates increases in primary productivity (Noy-

Meir 1973, Morton et al. 2011), with extreme rainfall events (>90th quantile) driving

interactions most strongly (Holmgren et al. 2006, Letnic and Dickman 2006, Smith 2011b,

32

Chapter 2: Extreme climatic events drive mammal irruptions

Popic and Wardle 2012). These pulse events are variable in time and space and often short- lived (1–2 years), yet the resources they generate drive population “booms” in consumer organisms such as small mammals through bottom-up processes, provide prey for predators, and increase fuel loads that result subsequently in wildfires (Letnic and Dickman 2006, Pavey et al. 2008, Greenville et al. 2009, Letnic et al. 2011). Extreme weather events are predicted to increase in magnitude and frequency (IPCC 2007, CSIRO and Australian Bureau of

Meteorology 2012) and thus may drive changes in arid systems disproportionately more than in mesic systems.

In this study, we collate long-term (100+ year) climate records from weather stations located on the periphery of the Simpson Desert, a vast arid region in central Australia, and examine trends in the climate over time. We compare trends at local (individual weather stations) and regional (Simpson Desert) scales, and focus in particular on changes in extreme climatic events which we define here as annual temperature or rainfall events in the 5th, 10th,

90th and 95th quantiles. These quantiles thus represent extreme cold or dry (5th and 10th quantile) and extreme hot or wet (90th and 95th quantile) climate events. We also investigate if extreme climate events are changing at similar rates to median (50th quantile) events, and ask how one case study group of organisms—small mammals—has responded to climatic fluctuations over a period of 22 years. Our climate-specific predictions are that:

1. annual temperatures, expressed as median and extreme values of maximum and

minimum annual temperatures;

2. the magnitude of extreme annual rainfall events;

3. intra-annual rainfall variability; and

4. the frequency of extreme annual rainfall events have all increased over time but at

different rates across local and regional scales. Conversely, we predict that:

33

Chapter 2: Extreme climatic events drive mammal irruptions

5. median annual rainfall will show an inconsistent pattern over time across local and

regional scales. From previous work we assume that small mammals require some

threshold in annual rainfall before their populations are able to irrupt, or “boom”. In

addition, because extreme temperatures may affect mammal populations through heat

stress, disruption of torpor or other physiological processes, we predict that:

6. thresholds for this case study group are extreme rainfall or minimum and maximum

temperature events (>90th quantile).

We use our findings to discuss the implications of change in extreme climate events for this arid system.

Materials and methods

Study region

This study was carried out within the 200 and 400 mm rainfall isopleths in central

Australia (Fig. 2.1). The Simpson Desert represents the major region in this rainfall band in

Australia and occupies 170 000 square kilometres (Fig. 2.1); fields comprise 73% of

this area, with smaller areas consisting of clay pans, rocky outcrops and gibber flats

(Shephard 1992). The sand dunes run parallel in a north-south direction aligned with the prevailing southerly wind. The dunes are up to 10 m high and spaced 0.6–1 km apart (Purdie

1984). Vegetation is predominantly grassland dominated by spinifex (Triodia basedowii)

with interdune swales covered variously with small stands of gidgee trees (Acacia

georginae), other woody Acacia shrubs or mallee-form eucalypts; low-lying clay pans fill

with water temporarily after heavy rain.

34

Chapter 2: Extreme climatic events drive mammal irruptions

During summer, daily temperatures usually exceed 40º C and minima in winter often

fall below 5º C (Purdie 1984). Rain falls mostly in summer, but heavy rains can occur locally or regionally throughout the year. Eleven weather stations closest to the desert (Fig. 2.1) have median annual rainfalls ranging from 153.1–216.2 mm collected over periods of 47 to 126 years (Bureau of Meteorology 2012).

Figure 2.1: Location of study region; the Simpson Desert, Australia, in relation to the Bureau of Meteorology (BOM) weather stations. Ethabuka Reserve is the location of the live mammal trapping used in this study.

Temperature

Mean annual minimum and maximum temperature data (calendar year) were obtained from Boulia (1888–2011), Alice Springs (1878–2011), Birdsville (1954–2011), and

Oodnadatta (1940–2011) weather stations (Bureau of Meteorology 2012), central Australia

35

Chapter 2: Extreme climatic events drive mammal irruptions

(Fig. 2.1). These stations were chosen because they have long-term (>50 years) temperature records and occur around the eastern and western edges of the Simpson Desert. Due to the remoteness of the study region and lack of any permanent settlements, no long term weather stations exist within the central region of the Simpson Desert.

To identify potential changes in the magnitude of extreme values of the above temperature variables over time (years), we used the median (50th quantile), two measures of

the upper extreme values (90th and 95th quantiles) and two measures of the lower extreme

values (5th and 10th quantiles) in analyses. Quantile regression was employed to test if the

slopes of these quantiles (5th, 10th, 50th, 90th and 95th) differed from zero. To estimate

standard errors and P-values, we used a bootstrap method with 10 000 replacements as

described in Koenker (1994). Quantile regressions were performed using the quantreg 4.76

package (Koenker 2011) in R 2.14.1 (R Development Core Team 2011).

Where significant non-zero slopes in annual temperatures were detected, confidence

intervals were calculated for each parameter. Changes over time in extreme annual minimum

and maximum temperatures (5th, 10th, 90th and 95th quantiles) were each compared to changes

in the median annual temperature (50th quantile), and significantly different changes at any

time between an outlier quantile and the median were identified by non-overlapping

confidence intervals.

Rainfall

Rainfall records (calendar year) within the 200 and 400 mm isopleths were obtained

from 11 weather stations in the study region (Fig. 2.1): Arltunga (1901–2011), Glenormiston

(1890–2011), Boulia (1886–2011), Marion Downs (1913–2011), Sandringham (1965–2011),

36

Chapter 2: Extreme climatic events drive mammal irruptions

Bedourie (1932–2011), Mt Dare (1950–2011), Oodnadatta (1892–2011), Alice Springs

(1874–2011), Undoolya (1898–2011) and Birdsville (1892–2011) (Bureau of Meteorology

2012). These weather stations have the longest records for the Simpson Desert study region; without weather stations throughout the desert the records from these small towns and pastoral station localities provide the most comprehensive datasets that are available.

As for the temperature data, we used quantile regressions to investigate changes in the magnitude of extreme and median rainfall events over time (years). If changes were detected, we calculated confidence intervals for each parameter to determine if there was a different rate of change (slope) in the magnitude of extreme annual rainfall (5th, 10th, 90th and 95th quantiles) compared to that of the median (50th quantile). Significant differences in rates of

change were deemed to occur if the confidence intervals did not overlap.

To investigate if the frequency of extreme rainfall events increased over time, each extreme event (≥95th quantile) was given a consecutive number and the return time (years)

until the next such event was calculated for 10 weather stations. A generalized linear mixed

model (GLMM), with a log-link function and Poisson distribution was chosen to best model

the return time, with extreme event as the fixed explanatory variable. Weather station was

entered into the model as a random factor, with both intercept and slope permitted to vary.

This allowed for local variation in annual rainfall and for relaxation of the assumption of

equal slopes of return time for each weather station (Gelman and Hill 2007). GLMMs were

performed using the lme4 package (Bates et al. 2011) in R 2.14.1 (R Development Core

Team 2011).

We also calculated the coefficient of variation (CV) for monthly rainfall for each

weather station to investigate temporal change in the variability of intra-annual rainfall. This

was done by calculating the mean monthly rainfall for each year divided by its standard

37

Chapter 2: Extreme climatic events drive mammal irruptions deviation. As above, quantile regression was used to investigate changes in the variability of annual rainfall in the median (50th quantile) and extreme ranges (5th, 10th, 90th and 95th quantiles). This measure was used because there was only one weather station at each site and it was not appropriate to average over all weather stations due to high local variation.

Small mammals

Live-trapping of small mammals was carried out on Ethabuka Reserve, a 2140 km2 property located in the north-eastern part of the Simpson Desert study region (Fig. 2.1). Small mammals were live-trapped using pitfall traps (16 cm diameter, 60 cm deep), each equipped with a 5 m drift fence of aluminium flywire to increase trap efficiency (Friend et al. 1989).

Pitfalls were arranged in a grid formation comprising six lines of six pitfall traps spaced 20 m apart to cover 1 ha. The top line of traps was positioned on a dune crest and the bottom line

100 m distant in the swale so that each grid sampled the topographic range of the dune field.

The location of each grid was randomly chosen, with grids 0.5–2 km apart. Six grids were operated in 1990 and 12 for the remainder of the study (Dickman et al. 1999, 2001, 2011).

Traps were opened for 3–6 nights every 2–4 months from 1990–2011. On each sampling occasion captured animals were identified, weighed, individually marked and sex determined as in Dickman et al. (1999, 2001). Captures were standardized per 100 trap nights (TN: trap nights = traps × nights opened) and averaged for each year.

To investigate whether small mammals responded to extreme rainfall and temperature events, the annual (calendar year) rainfall antecedent to capture and the annual minimum and maximum temperatures for captures that year were calculated from an on-site automatic weather station (Environdata, Warwick, Queensland). Missing values of rainfall were filled in from the nearest weather station at Sandringham, 63 km away (r = 0.81, P <0.001) and

38

Chapter 2: Extreme climatic events drive mammal irruptions missing values of temperature were filled from Boulia, 173 km away (r = 0.56, P = 0.02).

The 1-year lag was used to account for the time required for small mammals to respond, via immigration and breeding, to a rainfall event.

We divided mammal captures from the 22 years of live-trapping (397 476 trap nights) into rodents and dasyurids as these groups respond differently to rainfall and have different physiological strategies in response to food shortages (Dickman et al. 1999, 2001, Geiser

2004, Haythornthwaite and Dickman 2006). Captures of the several species in each group were combined; separate analyses of their dynamics are given in Chapters 3 and 4. Rodents included Pseudomys hermannsburgensis (sandy inland mouse; 4208 captures), Notomys alexis (spinifex hopping mouse; 2731 captures), P. desertor (desert mouse; 858 captures),

Rattus villosissimus (long-haired rat; 386 captures) and Mus musculus (house mouse; 306 captures). Dasyurids included Dasycercus blythi (brush-tailed mulgara; 438 captures),

Ningaui ridei (wongai ningaui; 501 captures), Sminthopsis youngsoni (lesser hairy-footed dunnart; 1301 captures), S. hirtipes (hairy-footed dunnart; 83 captures), S. macroura (striped- faced dunnart; 4 captures), and S. crassicaudata (fat-tailed dunnart; 10 captures).

Piecewise regression was used to identify any thresholds in annual rainfall and temperature at which rodents and dasyurids responded. A threshold relationship was predicted as extreme rains and temperatures (>90th quantile) are more likely to drive small mammal populations in arid Australia than moderate rains or temperatures; linear relationships probably do not occur (Letnic and Dickman 2010). Piecewise regressions use two or more lines, joined at a break point, that can be used to identify thresholds (Toms and

Lesperance 2003). We used piecewise regressions with one knot, or break point, that could occur at any rainfall or temperature value. The break point was estimated from likelihood- ratio statistics (Toms and Lesperance 2003). Confidence intervals were estimated from 1000

39

Chapter 2: Extreme climatic events drive mammal irruptions

bootstrap samples. Piecewise regressions were performed using SiZer 0.1-4 (Sonderegger

2011) in R 2.14.1 (R Development Core Team 2011).

Results

Temperature

Mean minimum annual temperatures showed significant warming over the period of

study at three weather stations (Boulia, Birdsville and Oodnadatta) in most quantiles, with

median temperatures (50th quantile) increasing by up to almost 2° C (Fig. 2.2a,b,c,d). The

median temperature increased also at the fourth weather station, Alice Springs, but no clear

changes were apparent in extreme minimum temperatures (Table 2.1, Fig. 2.2b). Mean

maximum annual temperatures showed significant warming of >1° C over the study period at

two weather stations (Birdsville and Oodnadatta), especially in the 10th, 50th, 90th and 95th

quantiles (Table 2.2, Fig. 2.3a,b). No significant trends were apparent at Alice Springs and

Boulia on the north-western and north-eastern fringes, respectively, of the Simpson Desert

(Table 2.2, Fig. 2.3c,d). There was no difference in the rate of increase in either the extreme

minimum or maximum annual temperatures compared with the rate of increase in median

temperatures (Tables 2.1 and 2.2).

40

Chapter 2: Extreme climatic events drive mammal irruptions

Figure 2.2: Mean minimum annual temperatures (° C) from a) Boulia, b) Alice Springs, c) Birdsville, and d) Oodnadatta weather stations, Simpson Desert, central Australia. Broken grey lines show 95th, 90th, 50th, 10th and 5thquantiles.

41

Chapter 2: Extreme climatic events drive mammal irruptions

Figure 2.3: Mean maximum annual temperatures (° C) from a) Birdsville, b) Oodnadatta, c), Alice Springs and d) Boulia weather stations, Simpson Desert, central Australia. Broken grey lines show 95th, 90th, 50th, 10th and 5thquantiles.

42

Chapter 2: Extreme climatic events drive mammal irruptions

Table 2.1: Quantile regression results for minimum temperature change over time (years), Simpson Desert, central Australia. In each row a significant P-value indicated by * implies that there is a non-zero slope for that quantile. For each weather station (Boulia, Alice Springs, Birdsville, Oodnadatta) the slope estimate for each of the extreme quantiles (0.5, 0.1, 0.9, and 0.95) was compared to the median quantile (0.5). Overlaps in the confidence intervals (CI) indicated similar rates of change for all comparisons.

Quantile Estimate CI t-value P Boulia 0.05 0.02 0.01–0.03 1.75 0.08 0.1 0.02 0.008–0.03 3.49 0.001* 0.5 0.02 0.01–0.02 6.07 <<0.001* 0.9 0.01 0.007–0.02 2.94 0.004* 0.95 0.01 0.009–0.02 3.50 0.001*

Alice Springs 0.05 −0.004 −0.007–0.006 −0.95 0.34 0.1 −0.002 −0.005–0.003 −0.46 0.65 0.5 0.006 −0.00002–0.01 1.83 0.07 0.9 0.004 0.0004–0.01 1.31 0.19 0.95 0.001 −0.003–0.01 0.39 0.70

Birdsville 0.05 0.04 −0.03–0.06 1.98 0.24 0.1 0.02 −0.01–0.06 1.18 0.24 0.5 0.02 0.01–0.03 3.20 0.002* 0.9 0.02 0.006–0.02 2.14 0.04* 0.95 0.009 −0.007–0.03 1.06 0.29

Oodnadatta 0.05 0.02 0.006–0.04 2.31 0.02* 0.1 0.02 0.007–0.02 2.77 0.007* 0.5 0.02 0.015–0.02 6.21 <<0.001* 0.9 0.01 0.007–0.02 4.81 <<0.001* 0.95 0.01 0.01–0.03 2.43 0.02*

43

Chapter 2: Extreme climatic events drive mammal irruptions

Table 2.2: Quantile regression results for maximum temperature change over time (years), Simpson Desert, central Australia. In each row a significant P-value indicated by * implies that there is a non-zero slope for that quantile. For each weather station (Boulia, Alice Springs, Birdsville, Oodnadatta) the slope estimate for each of the extreme quantiles (0.5, 0.1, 0.9, and 0.95) was compared to the median quantile (0.5). Overlaps in the confidence intervals (CI) indicated similar rates of change for all comparisons.

Quantile Estimate CI t-value P Boulia 0.05 0.01 0.003– 0.01 1.62 0.11 0.1 0.007 0.002–0.01 1.65 0.10 0.5 0.003 −0.003–0.005 0.98 0.23 0.9 0.005 −0.003–0.007 0.99 0.32 0.95 −0.002 −0.005–0.004 −0.31 0.76

Alice Springs 0.05 0.002 −0.02–0.02 0.28 0.78 0.1 0.003 −0.01–0.01 0.48 0.63 0.5 0.004 −0.001–0.009 1.29 0.20 0.9 −0.002 −0.004–0.001 −0.61 0.54 0.95 −0.003 −0.004–0.002 −1.07 0.29

Birdsville 0.05 0.003 −0.004–0.06 0.12 0.90 0.1 0.04 0.01–0.06 1.95 0.06 0.5 0.03 0.02–0.04 4.30 <<0.001* 0.9 0.02 0.01–0.05 1.55 0.12 0.95 0.03 0.006–0.04 2.73 0.01*

Oodnadatta 0.05 0 −0.004–0.03 0 1 0.1 0.01 −0.001–0.02 1.97 0.05* 0.5 0.02 0.01–0.03 3.98 <<0.001* 0.9 0.02 0.01–0.03 2.99 0.004* 0.95 0.03 −0.002–0.04 3.50 <<0.001*

Rainfall

The magnitude and direction of change in annual rainfall varied among the long-term

weather stations in central Australia, with some indication of both local and regional trends.

Four weather stations (Bedourie, Glenormiston, Undoolya and Oodnadatta) showed

significant increases in the magnitude of extreme rainfall events (90th and 95th quantiles; Fig.

2.4a,b,c,d), with peak rainfalls roughly doubling in size over the study period; two weather

stations (Birdsville and Marion Downs) showed increases in median rainfall (50th quantile;

Table 2.3, Fig. 2.4e,f). Two weather stations (Birdsville and Alice Springs) showed

44

Chapter 2: Extreme climatic events drive mammal irruptions

significant changes in rainfall during extreme dry years (5th and 10th quantiles), but the

direction of these changes was inconsistent (Table 2.3, Fig. 2.4e,g). Four weather stations

(Boulia, Sandringham, Arltunga and Mt. Dare) showed no significant relationship (Table 2.3,

Fig. A1.11a,b,c,d). In general, extreme annual rainfall events increased in magnitude around

the Simpson Desert whereas increases in the magnitude of median rainfall events were

restricted to the regions. In the west of the Simpson Desert extremely dry years

became drier over the study period, whereas the reverse pattern occurred in the east.

The rate of change (increase) in the magnitude of extreme rainfall events (90th and

95th quantiles) was greater than the median rate of change at four weather stations (Bedourie,

Glenormiston, Undoolya and Oodnadatta; Table 2.3), indicating a desert-wide increase in the size of extremely heavy rainfall events over the last century.

45

Chapter 2: Extreme climatic events drive mammal irruptions

Figure 2.4: Annual rainfall (mm) recorded each year for a) Bedourie, b) Glenormiston, c) Undoolya, d) Oodnadatta, e) Birdsville, f) Marion Downs, and g) Alice Springs weather stations, Simpson Desert, central Australia. Broken grey lines show 95th, 90th, 50th, 10th and 5thquantile. 90 and 95thquantiles represent boom years, 50th is the median and 10th and 5th represent bust years.

46

Chapter 2: Extreme climatic events drive mammal irruptions

Table 2.3: Quantile regression results for annual rainfall changes over time (years) for each weather station, Simpson Desert, central Australia. The estimates in bold indicate that the slope does not overlap the confidence interval (CI) for the median slope (0.5 quantile). Significant P-value indicated by *.

Quantile Estimate CI t-value P

Bedourie 0.05 −0.54 −0.13–1.89 −0.77 0.44 0.1 0.17 −0.83–1.84 0.28 0.78 0.5 0.41 −0.17–1.26 0.65 0.52 0.9 3.78 0.68–5.05 1.83 0.07 0.95 4.15 2.10–5.10 2.03 0.04*

Glenormiston 0.05 0.15 −0.17–0.97 0.41 0.68 0.1 0.40 −0.18–1.10 1.32 0.19 0.5 0.76 −0.12–1.64 1.71 0.09 0.9 1.95 0.36–3.10 1.79 0.07 0.95 2.44 1.06–3.68 2.05 0.04*

Marion Downs 0.05 0.42 −0.15–0.73 1.11 0.27 0.1 0.50 −0.003–1.17 1.18 0.24 0.5 1.15 0.62–1.93 2.37 0.02* 0.9 1.55 −0.08–3.72 1.11 0.27 0.95 2.10 0.10–3.30 1.41 0.16

Boulia 0.05 0.19 −0.90–0.35 0.53 0.59 0.1 −0.16 −0.61–0.52 −0.53 0.59 0.5 0.79 −0.69–1.04 1.52 0.13 0.9 −1.77 −2.37–1.51 −1.25 0.21 0.95 0.07 −2.45–3.26 0.04 0.97

Birdsville 0.05 0.43 −0.15–0.59 1.97 0.05* 0.1 0.31 0.15–0.64 1.74 0.09 0.5 0.89 0.31–1.39 2.48 0.01* 0.9 −0.09 −2.24–1.71 −0.07 0.94 0.95 −0.63 −3.34–1.48 −0.42 0.67

Sandringham 0.05 −0.79 −2.67–2.24 −0.42 0.67 0.1 −1.39 −2.41–2.09 −0.92 0.36 0.5 −1.95 −3.81–1.70 −1.18 0.24 0.9 1.24 −6.90–3.39 0.38 0.70 0.95 0.56 −7.36–26.32 0.16 0.87

Arltunga 0.05 −0.14 −5.17–1.00 −0.31 0.75 0.1 −0.24 −1.32–0.80 −0.57 0.57 0.5 0.10 −0.72–1.32 0.15 0.88 0.9 1.67 0.47–4.55 0.95 0.34 0.95 1.70 0.55–5.54 0.80 0.42

Undoolya 0.05 −0.11 −0.41–2.43 −0.24 0.81 0.1 0.18 −0.33–1.63 0.37 0.71

47

Chapter 2: Extreme climatic events drive mammal irruptions

0.5 0.84 −0.89–0.98 1.38 0.17 0.9 2.32 0.83–3.75 1.77 0.08 0.95 3.70 1.69–4.17 2.45 0.02*

Alice Springs 0.05 −0.48 −0.86– −0.32 −2.41 0.02* 0.1 −0.48 −6.66– −0.01 −1.96 0.05* 0.5 −0.11 −0.31–0.69 −0.36 0.72 0.9 0.86 −0.70–2.66 0.72 0.47 0.95 2.20 0.50–2.51 1.84 0.07

Oodnadatta 0.05 0.32 −0.05–0.74 1.27 0.21 0.1 0.33 −0.09–0.56 1.50 0.14 0.5 0.50 0.18–1.00 1.91 0.06 0.9 1.81 1.16–2.20 3.83 <<0.001* 0.95 1.96 0.97–3.70 2.10 0.04*

Mt Dare 0.05 0.58 −0.98–1.71 0.83 0.41 0.1 0.51 −0.36–1.66 0.74 0.46 0.5 0.75 −0.77–1.56 0.94 0.35 0.9 −2.33 −8.39–3.68 −0.71 0.48 0.95 −2.27 −8.80–10.63 −0.51 0.61

Extreme rainfall return intervals

There was a marked decrease in the number of years between extreme rainfall events

th (>95 quantile) (GLMM estimate: −0.54, SE = 0.21, z value = −2.54, P = 0.01) across the

Simpson Desert over the last 100 years, indicating that such events have become more frequent over time.

Rainfall variability

The coefficient of variation in rainfall decreased over the period of study at three weather stations, Bedourie, Glenormiston and Boulia (Table 2.4). Variability in rainfall decreased in magnitude during extreme dry years at the first two of these stations and showed a decline in the median rainfall at Boulia. All three weather stations are on the eastern edge of the Simpson Desert, suggesting that this trend is not consistent across central Australia.

48

Chapter 2: Extreme climatic events drive mammal irruptions

Table 2.4: Quantile regression results for intra-annual rainfall coefficient of variation for weather stations in the Simpson Desert, central Australia. Significant P-value indicated by *.

Quantile Estimate SE t-value P

Bedourie 0.05 −0.009 0.002 −4.62 <<0.001* 0.1 −0.008 0.002 −3.37 0.001* 0.5 −0.002 0.004 −0.49 0.63 0.9 0.003 0.004 0.63 0.53 0.95 0.007 0.004 1.75 0.09

Glenormiston 0.05 −0.002 0.002 −0.1 0.33 0.1 −0.003 0.001 −2.39 0.02* 0.5 −0.004 0.002 −1.81 0.07 0.9 −0.001 0.004 −0.31 0.76 0.95 −0.001 0.003 −0.41 0.68

Marion Downs 0.05 −0.0004 0.002 −0.17 0.87 0.1 −0.003 0.003 −0.96 0.34 0.5 −0.0007 0.002 −0.08 0.93 0.9 −0.001 0.005 −0.19 0.85 0.95 0.003 0.005 0.64 0.52

Boulia 0.05 −0.002 0.002 −1.1 0.28 0.1 −0.0004 0.002 −0.29 0.78 0.5 −0.003 0.001 −2.83 0.005* 0.9 −0.0002 0.002 −0.07 0.95 0.95 0.0006 0.002 0.31 0.76

Birdsville 0.05 −0.0002 0.002 −0.1 0.92 0.1 0.0008 0.002 0.51 0.61 0.5 0.0003 0.001 0.25 0.81 0.9 −0.005 0.004 −1.23 0.22 0.95 −0.006 0.004 −1.8 0.08

Sandringham 0.05 −0.00002 0.008 −0.002 0.99 0.1 −0.009 0.008 −1.11 0.27 0.5 0.007 0.008 0.93 0.36 0.9 −0.008 0.01 −0.79 0.44 0.95 0.001 0.009 0.12 0.90

Arltunga 0.05 −0.0009 0.002 −0.52 0.61 0.1 −0.002 0.002 −0.85 0.40 0.5 −0.0007 0.002 −0.37 0.71 0.9 −0.0006 0.002 −0.26 0.80 0.95 −0.0003 0.003 −0.09 0.93

Undoolya 0.05 0.0003 0.003 0.09 0.93 0.1 0.002 0.002 1.29 0.20 0.5 0.003 0.002 1.28 0.21

49

Chapter 2: Extreme climatic events drive mammal irruptions

0.9 0.0002 0.004 0.06 0.95 0.95 −0.004 0.006 −0.67 0.51

Alice Springs 0.05 −0.0005 0.0009 −0.58 0.56 0.1 −0.0008 0.0009 −0.84 0.40 0.5 0 0.001 −0.001 1 0.9 0.003 0.002 1.60 0.11 0.95 0.003 0.003 0.98 0.33

Oodnadatta 0.05 −0.003 0.002 −1.06 0.29 0.1 −0.002 0.002 −0.79 0.43 0.5 0.0003 0.001 0.23 0.82 0.9 −0.002 0.005 −0.38 0.70 0.95 0.00008 0.006 0.01 0.99

Mt Dare 0.05 −0.0005 0.006 −0.09 0.93 0.1 0.003 0.004 0.86 0.39 0.5 −0.002 0.004 −0.49 0.63 0.9 −0.008 0.01 −0.72 0.47 0.95 0.003 0.01 0.20 0.84

Small mammals

Rodent captures showed a strongly positive response to rainfall in the preceding year at a threshold of 418 mm (CI: 240–506 mm; Table 2.5; Fig. 2.5a). This represents an extreme

annual rainfall event in the ~95th quantile on Ethabuka Reserve. In contrast, dasyurid captures

showed no response to annual rainfall from the preceding year (Table 2.5; Fig. 2.5b). There

was no significant relationship for rodents or dasyurids and minimum or maximum

temperatures (Table 2.6).

50

Chapter 2: Extreme climatic events drive mammal irruptions

Table 2.5: Piecewise regression results for the response of small mammal captures to annual rainfall from the preceding year, Ethabuka Reserve, Simpson Desert, central Australia. Small mammal captures were standardised per 100 trap nights (TN) and averaged per year (n = 22 years). Significant P-value indicated by *.

Estimate SE t-value P Rodents Intercept 7.60 4.42 1.72 0.10 Line 1 −0.001 0.02 −0.07 0.95 Line 2 0.17 0.05 3.11 0.006*

Dasyurids Intercept 3.24 1.77 1.83 0.08 Line 1 0.02 0.01 1.19 0.25 Line 2 −0.02 0.01 −1.06 0.30

Figure 2.5: The response of a) rodent and b) dasyurid marsupial captures to rainfall in the preceding year. Rainfall was recorded from on-site automatic weather station at Ethabuka (Environdata, Warwick, Queensland). Missing values were filled in from the nearest weather station at Sandringham, 63 km away. Lines depict predicted values from piecewise regression. Rodent captures show a response to rainfall only after a threshold of 418 mm (~ 0.95 quantile) is reached. Labels on (a) indicate year of irruption. No relationship between rainfall and dasyurids was found. Small mammal captures were standardised per 100 trap nights (TN) and averaged per year (n = 22 years).

51

Chapter 2: Extreme climatic events drive mammal irruptions

Table 2.6: Piecewise regression results for the response of small mammal captures to annual minimum and maximum temperatures, Ethabuka Reserve, Simpson Desert, central Australia. Small mammal captures were standardised per 100 trap nights (TN) and averaged per year (n = 22 years). Significant P-value indicated by *.

Estimate SE t-value P Minimum temperature Rodents Intercept 9.81 40.68 0.24 0.81 Line 1 −0.04 2.94 −0.01 0.99 Line 2 1.87 4.60 0.41 0.69

Dasyurids Intercept −38.21 25.42 −1.50 0.15 Line 1 3.78 2.19 1.72 0.10 Line 2 −3.91 2.26 −1.73 0.10

Maximum temperature Rodents Intercept 88.49 83.70 1.06 0.30 Line 1 −2.41 2.60 −0.93 0.37 Line 2 7.10 9.07 0.78 0.44

Dasyurids Intercept −3.70 11.31 −0.33 0.75 Line 1 0.28 0.35 0.81 0.43 Line 2 −0.24 0.71 −0.34 0.74

Discussion

We predicted that recent global warming would be reflected by increases in temperatures at local (individual weather stations) and regional (Simpson Desert) scales, for both median and extreme temperature values. In contrast, we predicted increases in the size and frequency of extreme rainfall events over time in the Simpson Desert, but expected median rainfall to vary inconsistently at both spatial scales. In general, temperatures increased largely as predicted while trends in rainfall showed local and regional variation; the magnitude of heavy rainfall events increased at four weather stations on the eastern and western fringes of the Simpson Desert, and the frequency of extreme rainfall events also increased. Small mammals showed differing responses to rainfall; rodent populations irrupted only after an extreme rainfall event (>95th quantile), whereas dasyurids did not. There was no 52

Chapter 2: Extreme climatic events drive mammal irruptions

response by small mammal populations to temperature. Here we address each climate

variable in turn and then, using small mammals as a case study, we discuss the likely

responses of desert organisms to climate change.

Temperature

Our results confirm that there has been significant warming in the Simpson Desert region over the last century. Mean annual night temperatures have increased, with 6 of the 12 hottest nights occurring since 1970 (90–95th quantiles) and the coolest nights also warming

over the study period (10th quantile; Fig. 2.2a,c,d). This warming has occurred locally and

regionally, suggesting that increases in global temperatures manifest at both scales. These

results are consistent with previous studies in Australia (Hughes 2003, Trewin and Vermont

2010, Hughes 2011) and globally (IPCC 2007, Tabari and Hosseinzadeh Talaee 2011). We

found similar rates of change of both median and extreme annual temperatures, suggesting

that increases in global temperatures have affected median and extreme temperatures equally.

Global temperatures are predicted to increase by 0.6–4° C by the end of the 21st

century and Australian temperatures to increase by 1–5° C by 2070, depending on emission

scenarios (CSIRO and Australian Bureau of Meteorology 2007, IPCC 2007). The north-

western and central regions of Australia are predicted to show the greatest warming in both

minimum and maximum temperatures (CSIRO and Australian Bureau of Meteorology 2007).

Our results showing changes in both the median and extreme temperatures are consistent with

these predictions.

53

Chapter 2: Extreme climatic events drive mammal irruptions

Rainfall

There was variability in long-term annual rainfall trends at local and regional scales.

The Bedourie, Undoolya, Oodnadatta and Glenormiston weather stations showed significant

increases in 95th quantile rainfall events (Table 2.3), suggesting that the magnitudes of these

large rainfalls (>300 mm) have increased in recent decades. Marked increases in median

rainfalls were found for the Marion Downs and Birdsville weather stations and increases in

rainfall during extremely dry years (5th quantile) occurred at Birdsville. In contrast, rainfall at

Alice Springs decreased during extremely dry years (Table 2.3). Rainfall trends in Australia show strong spatial variability (CSIRO and Australian Bureau of Meteorology 2007). At the

continental scale there has been a non-significant increase in rainfall (Hennessy et al. 1999), but regional patterns are more clear (Hennessy et al. 1999, Hughes 2003). Since 1950, annual rainfall has increased in the north-west of the continent and decreased in the east; in central arid regions annual rainfall generally has increased (CSIRO and Australian Bureau of

Meteorology 2007). We found that annual rainfalls have increased during extremely wet years, while increases in the magnitude of median rainfall events were restricted to eastern parts of the Simpson Desert, with some local variability. In the west of the Simpson Desert extremely dry years received reduced annual rainfall, but in the east there were increases.

Higher rainfall has been linked to increases in both heavy rainfall events and number of rain days (Hennessy et al. 1999, Hughes 2003, CSIRO and Australian Bureau of Meteorology

2007), which is consistent with increases in rainfall in the 95th quantile in our study.

As global temperatures rise the moisture-holding capacity of the atmosphere increases and circulation patterns change, with resultant impacts on global rainfall (CSIRO and

Australian Bureau of Meteorology 2007, IPCC 2007). By the end of the 21st century rainfall

is predicted to increase at high latitudes and decrease in subtropical regions (IPCC 2007).

54

Chapter 2: Extreme climatic events drive mammal irruptions

Discrepancies in rainfall predictions for Australia by the end of the 21st century arise largely from the use of different approaches to determine the base level of rainfall (Hughes 2003,

CSIRO and Australian Bureau of Meteorology 2007). Some models predict an increase in rainfall from tropical northern Australia to parts of the arid interior, while others predict a general decline in rainfall across most of Australia (CSIRO and Australian Bureau of

Meteorology 2007). We found that changes in median rainfall over the last century differed across the regional study area. Two weather stations showed a significant increase in annual median rainfall, while the other stations showed no pattern (Table 2.3). The intensity of extreme rainfall events is predicted to increase across most of Australia (CSIRO and

Australian Bureau of Meteorology 2007), and globally (IPCC 2007). Rainfall patterns in

Australia are influenced by the El Niño Southern Oscillation, which is predicted to intensify due to climate change (IPCC 2007). We found increases in 95th quantile rainfall in the

Simpson Desert to have occurred at significantly greater rates than the median; this is consistent with suggestions that past changes in extreme rainfall events will continue into the

21st century, although considerable spatial variability is likely to occur.

In general, the time between extremely large rainfall events (>95th quantile) decreased over time in the Simpson Desert, suggesting that the frequency of extreme rainfall events is increasing. This is consistent with the predictions of global climate change models (IPCC

2007) and with predictions for the future Australian climate (CSIRO and Australian Bureau of Meteorology 2007, 2012).

We predicted that annual rainfall would become more variable due to the impacts of climate change in the Simpson Desert, but found no support for this at either local or regional scales. Instead, we found a decrease in rainfall variability during extremely dry years— implying longer droughts—at two weather stations and for median rainfall at another weather

55

Chapter 2: Extreme climatic events drive mammal irruptions

station. All three weather stations are on the eastern edge of the Simpson Desert, suggesting

that this trend is not consistent across the study region.

Case study: small mammals

The two groups of small mammals in this study showed contrasting responses to

annual rainfall and no response to annual minimum and maximum temperatures. Captures of

rodents increased only after extreme rainfall years (~95th quantile, or >400 mm rain), whereas dasyurid captures showed no response. In arid regions increases in small mammal populations after rainfall are facilitated by increases in the availability of food (Dickman et al. 1999, Letnic et al. 2005, Dickman et al. 2011, Letnic et al. 2011), with rodents exhibiting

the strongest responses (Dickman et al. 1999, Lima and Jaksic 1999, Lima et al. 2002, Lima et al. 2008, Pavey et al. 2008). Dasyurids do not appear to be limited by rainfall, but by other factors such as vegetation cover and life history constraints (Dickman et al. 2001). Most species breed once in the Austral winter—spring and thus have limited opportunity to respond to infrequent, heavy rainfall events (Dickman et al. 2001).

With increases in the magnitude and frequency of extreme rainfall events recorded in at least some local areas of the Simpson Desert in recent decades, associated rodent irruptions are likely to have taken place. Historical records confirm that rodents have frequently irrupted following years receiving 95th quantile rainfall (e.g. Cleland 1918, Newsome and Corbett

1975, Watts and Aslin 1981), and our 22-year database suggests further that the size of

irruptions is related to the amount of rain received (Fig. 2.5a). If the trends that we have

identified continue, rodent irruptions will likely occur more frequently in future. It is possible

also that populations of the constituent rodent species will achieve greater densities during

outbreaks; very large population increases may be dampened by intraspecific competition or

56

Chapter 2: Extreme climatic events drive mammal irruptions

other density-dependent interactions (Lima and Jaksic 1999, Lima et al. 2008), but these are yet to be demonstrated in this study region (Dickman et al. 2010).

Implications of climate change

Continuing increases in the frequency and magnitude of extreme rainfall events will likely have several important ecological consequences for biodiversity in inland Australia.

We draw here on our observations of small mammals to provide focus, but acknowledge that most organisms will be affected directly and indirectly by temperature and rainfall shifts.

Firstly, while irruptions of native rodents can be expected to increase, new species also may be able to establish during resource-rich boom conditions. Introduced rodent species such as the house mouse (Mus musculus), which is largely absent from the study region in dry times, may become more common (Letnic and Dickman 2010). Secondly, novel predators also may use boom conditions to become established. There is considerable evidence that feral cat

(Felis catus) and red fox (Vulpes vulpes) populations can increase in the wake of extreme rainfall events, either by elevating their breeding in situ or by moving in from more mesic environments near the study region in response to elevated populations of rodent prey (Letnic et al. 2005, Letnic and Dickman 2006, Pavey et al. 2008). Top-down processes then may become stronger and more pervasive, reducing small mammal populations to very low levels during droughts (Letnic et al. 2011, Chapter 6). Thirdly, other novel interactions may develop; these include extended floral-pollinator networks that can be supported during periods of high productivity, and hyper-predation of seeds, green vegetation and invertebrates that can arise when rodent populations are high (Brown et al. 1997, Bellard et al. 2012, Popic and Wardle 2012). Finally, the risk of wildfires increases after two consecutive years of rainfall that exceeds the 91st quantile in the study region (Letnic and Dickman 2006,

57

Chapter 2: Extreme climatic events drive mammal irruptions

Greenville et al. 2009). An increase in the magnitude and frequency of extreme rainfall

events thus may promote higher fuel loads, increase fire risk through increased temperatures

and evaporation, and decrease the fire return interval. Increased incidence of wildfires may

favour species that do not require high vegetation cover such as some agamid species, but

disadvantage many others that need dense cover (Haythornthwaite and Dickman 2006,

Pianka and Goodyear 2012). As vegetation also provides a food resource for invertebrates

and protection from predation, its removal by fire could exacerbate the negative effects of

predators on small mammal populations (Letnic and Dickman 2006).

If the identified climatic trends continue, as predicted, rodents will probably experience exaggerated boom and bust cycles that will be characterized by large irruptions and deep and prolonged troughs. These dynamics will likely create conditions that are conducive to stochastic population extinctions at local or regional levels (Xu et al. 2006), and also allow suites of novel ecological interactions to develop that will influence small mammal assemblages in different ways. We conclude that these novel conditions will complicate attempts to conserve biodiversity in arid environments, but suggest also that long-term datasets and analyses of the biotic effects of extreme events will provide important insights into this task.

Acknowledgements

We thank , H. Jukes, G. McDonald, D. Smith and G. Woods for allowing access to the properties in the study region; A. Foster for valuable comments on earlier drafts of the manuscript; members of the Desert Ecology Research Group, B. Tamayo and C.-L. Beh for valuable assistance in the field, and the Australian Research Council and an

Australian Postgraduate Award (to ACG) for providing funding.

58

Chapter 2: Extreme climatic events drive mammal irruptions

References:

Allan, R., J. Lindesay, and D. Parker. 1996. El Niño Southern Oscillation and climatic variability. CSIRO Publishing., Melbourne, Australia. Bates, D., M. Maechler, and B. Bolker. 2011. lme4: Linear mixed-effects models using S4 classes. R package version 0.999375-42. http://CRAN.R-project.org/package=lme4. Bellard, C., C. Bertelsmeier, P. Leadley, W. Thuiller, and F. Courchamp. 2012. Impacts of climate change on the future of biodiversity. Ecology Letters 15: 1-13. Brown, J. H., T. J. Valone, and C. G. Curtin. 1997. Reorganization of an arid ecosystem in response to recent climate change. Proceedings of the National Academy of Sciences 94: 9729-9733. Bureau of Meteorology. 2012. Climate data online, accessed 20 June 2012, http://www.bom.gov.au/climate/data/ Chamaillé-Jammes, S., H. Fritz, and F. Murindagomo. 2007. Detecting climate changes of concern in highly variable environments: quantile regressions reveal that droughts worsen in Hwange National Park, Zimbabwe. Journal of Arid Environments 71: 321- 326. Cleland, J. B. 1918. Presidential address to the Royal Society of NSW. Journal and Proceedings of the Royal Society of NSW 52: 1-165. CSIRO and Australian Bureau of Meteorology. 2007. Climate change in Australia. Technical report, CSIRO and Australian Bureau of Meteorology for the Australian Climate Change Science Programme., www.climatechangeinaustralia.gov.au CSIRO and Australian Bureau of Meteorology. 2012. State of the climate 2012, CSIRO and the Australian Bureau of Meteorology, www.climatechangeinaustralia.gov.au Dickman, C. R., A. C. Greenville, C.-L. Beh, B. Tamayo, and G. M. Wardle. 2010. Social organization and movements of desert rodents during population “booms” and “busts” in central Australia. Journal of Mammalogy 91: 798-810. Dickman, C. R., A. C. Greenville, B. Tamayo, and G. M. Wardle. 2011. Spatial dynamics of small mammals in central Australian desert habitats: the role of drought refugia. Journal of Mammalogy 92: 1193-1209. Dickman, C. R., A. S. Haythornthwaite, G. H. McNaught, P. S. Mahon, B. Tamayo, and M. Letnic. 2001. Population dynamics of three species of dasyurid marsupials in arid central Australia: a 10-year study. Wildlife Research 28: 493-506.

59

Chapter 2: Extreme climatic events drive mammal irruptions

Dickman, C. R., P. S. Mahon, P. Masters, and D. F. Gibson. 1999. Long-term dynamics of rodent populations in arid Australia: the influence of rainfall. Wildlife Research 26: 389-403. Donohue, R. J., T. R. McVicar, and M. L. Roderick. 2010. Assessing the ability of potential evaporation formulations to capture the dynamics in evaporative demand within a changing climate. Journal of Hydrology 386: 186-197. Epps, C. W., D. R. McCullough, J. D. Wehausen, V. C. Bleich, and J. L. Rechel. 2004. Effects of climate change on population persistence of desert-dwelling mountain sheep in California. Conservation Biology 18: 102-113. Fensham, R. J. and J. E. Holman. 1999. Temporal and spatial patterns in drought-related tree dieback in Australian savanna. Journal of Applied Ecology 36: 1035-1050. Friend, G. R., G. T. Smith, D. S. Mitchell, and C. R. Dickman. 1989. Influence of pitfall and drift fence design on capture rates of small vertebrates in semi-arid habitats of Western Australia. Australian Wildlife Research 16: 1-10. Geiser , F. 2004. The role of torpor in the life of Australian arid zone mammals. Australian Mammalogy 26: 125-134. Gelman, A. and J. Hill. 2007. Data analysis using regression and multilevel/hierarchical models. Cambridge University Press, Cambridge. Greenville, A. C., C. R. Dickman, G. M. Wardle, and M. Letnic. 2009. The fire history of an arid grassland: the influence of antecedent rainfall and ENSO. International Journal of Wildland Fire 18: 631-639. Haythornthwaite, A. S. and C. R. Dickman. 2006. Distribution, abundance, and individual strategies: a multi-scale analysis of dasyurid marsupials in arid central Australia. Ecography 29: 285-300. Hennessy, K. J., R. Suppiah, and C. M. Page. 1999. Australian rainfall changes, 1910-1995. Australian Meteorological Magazine 48: 1-13. Hoffman, T. and C. Vogel. 2008. Climate change impacts on African rangelands. Rangelands 30: 12-17. Holmgren, M., P. Stapp, C. R. Dickman, C. Gracia, S. Graham, J. R. Gutierrez, C. Hice, F. Jaksic, D. A. Kelt, M. Letnic, M. Lima, B. C. Lopez, P. L. Meserve, W. B. Milstead, G. A. Polis, M. A. Previtali, R. Michael, S. Sabate, and F. A. Squeo. 2006. Extreme climatic events shape arid and semiarid ecosystems. Frontiers in Ecology and the Environment 4: 87-95.

60

Chapter 2: Extreme climatic events drive mammal irruptions

Hughes, L. 2003. Climate change and Australia: trends, projections and impacts. Austral Ecology 28: 423-443. Hughes, L. 2011. Climate change and Australia: key vulnerable regions. Regional Environmental Change 11: 189-195. IPCC. 2007. Climate change 2007: Synthesis report. Contribution of Working Groups I, II and III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, Geneva. Jentsch, A., J. Kreyling, M. Elmer, E. Gellesch, B. Glaser, K. Grant, R. Hein, M. Lara, H. Mirzae, S. E. Nadler, L. Nagy, D. Otieno, K. Pritsch, U. Rascher, M. Schädler, M. Schloter, B. K. Singh, J. Stadler, J. Walter, C. Wellstein, J. Wöllecke, and C. Beierkuhnlein. 2011. Climate extremes initiate ecosystem-regulating functions while maintaining productivity. Journal of Ecology 99: 689-702. Koenker, R. W. 1994. Confidence intervals for regression quantiles., Pages 349–359 in P. Mandl and M. Huskova, editors. Asymptotic statistics. Springer-Verlag, New York. Koenker, R. W. 2011. quantreg: Quantile Regression. R package version 4.76. Letnic, M. and C. R. Dickman. 2006. Boom means bust: interactions between the El Niño/Southern Oscillation (ENSO), rainfall and the processes threatening mammal species in arid Australia. Biodiversity and Conservation 15: 3847-3880. Letnic, M. and C. R. Dickman. 2010. Resource pulses and mammalian dynamics: conceptual models for hummock grasslands and other Australian desert habitats. Biological Reviews 85: 501-521. Letnic, M., P. Story, G. Story, J. Field, O. Brown, and C. R. Dickman. 2011. Resource pulses, switching trophic control, and the dynamics of small mammal assemblages in arid Australia. Journal of Mammalogy 92: 1210-1222. Letnic, M., B. Tamayo, and C. R. Dickman. 2005. The responses of mammals to La Niña (El Niño Southern Oscillation); associated rainfall, predation, and wildfire in central Australia. Journal of Mammalogy 86: 689-703. Lima, M., S. K. M. Ernest, J. H. Brown, A. Belgrano, and N. C. Stenseth. 2008. Chihuahuan Desert kangaroo rats: nonlinear effects of population dynamics, competition and rainfall. Ecology 89: 2594-2603. Lima, M. and F. M. Jaksic. 1999. Population rate of change in the leaf-eared mouse: the role of density-dependence, seasonality and rainfall. Australian Journal of Ecology 24: 110-116.

61

Chapter 2: Extreme climatic events drive mammal irruptions

Lima, M., N. C. Stenseth, and F. M. Jaksic. 2002. Food web structure and climate effects on the dynamics of small mammals and owls in semi-arid Chile. Ecology Letters 5: 273- 284. Low, T. 2011. Climate change and Queensland biodiversity. Queensland Department of Environment and Resource Management, Brisbane, Australia. McKechnie, A. E., P. A. R. Hockey, and B. O. Wolf. 2012. Feeling the heat: Australian landbirds and climate change. Emu 112: 1-7. Morton, S. R., D. M. Stafford Smith, C. R. Dickman, D. L. Dunkerley, M. H. Friedel, R. R. J. McAllister, J. R. W. Reid, D. A. Roshier, M. A. Smith, F. J. Walsh, G. M. Wardle, I. W. Watson, and M. Westoby. 2011. A fresh framework for the ecology of arid Australia. Journal of Arid Environments 75: 313-329. Newsome, A. E. and L. K. Corbett. 1975. Outbreaks of rodents in semi-arid and arid Australia: causes, preventions and evolutionary considerations, Pages 117-153 in I. Prakash and P. K. Ghosh, editors. Rodents in desert environments. Dr. W. Junk Publishers, The Hague. Noy-Meir, I. 1973. Desert ecosystems: environment and producers. Annual Review of Ecology and Systematics 4: 25-51. Pavey, C. R., S. R. Eldridge, and M. Heywood. 2008. Population dynamics and prey selection of native and introduced predators during a rodent outbreak in arid Australia. Journal of Mammalogy 89: 674-683. Peters, D. P. C., B. T. Bestelmeyer, and A. K. Knapp. 2011. Perspectives on global change theory, Pages 261-281 in S. M. Scheiner and M. R. Willig, editors. The Theory of Ecology. The University of Chicago Press, Chicago. Pianka, E. R. and S. E. Goodyear. 2012. Lizard responses to wildfire in arid interior Australia: long-term experimental data and commonalities with other studies. Austral Ecology 37: 1-11. Popic, T. J. and G. M. Wardle. 2012. Extremes: understanding flower – visitor interactions in a changing climate., Pages 99-106 in D. Lunney and P. Hutchings, editors. Wildlife and climate change: towards robust conservation strategies for Australian fauna. Royal Zoological Society of New South Wales, Mosman, NSW, Australia. Purdie, J. L. 1984. Land systems of the Simpson Desert region. Natural resources series no. 2. CSIRO Division of Water and Land Resources, Melbourne, Australia. R Development Core Team. 2011. R: A language and environment for statistical computing. R Foundation for Statistical Computing., Vienna, Austria. 62

Chapter 2: Extreme climatic events drive mammal irruptions

Shephard, M. 1992. The Simpson Desert: natural history and human endeavour. Royal Geographical Society of Australasia, Adelaide, Australia. Smith, F. A., H. Browning, and U. L. Shepherd. 1998. The influence of climate change on the body mass of woodrats Neotoma in an arid region of New Mexico, USA. Ecography 21: 140-148. Smith, M. D. 2011a. An ecological perspective on extreme climatic events: a synthetic definition and framework to guide future research. Journal of Ecology 99: 656-663. Smith, M. D. 2011b. The ecological role of climate extremes: current understanding and future prospects. Journal of Ecology 99: 651-655. Sonderegger, D. 2011. SiZer: SiZer: Significant Zero Crossings. R package version 0.1-4. Tabari, H. and P. Hosseinzadeh Talaee. 2011. Analysis of trends in temperature data in arid and semi-arid regions of Iran. Global and Planetary Change 79: 1-10. Toms, J. D. and M. L. Lesperance. 2003. Piecewise regression: a tool for identifying ecological thresholds. Ecology 84: 2034-2041. Trewin, B. and H. Vermont. 2010. Changes in the frequency of record temperatures in Australia, 1957-2009. Australian Meteorological and Oceanographic Journal 60: 113-119. Van Etten, E. J. B. 2009. Inter-annual rainfall variability of arid Australia: greater than elsewhere? Australian Geographer 40: 109-120. Vincenzi, S., G. A. De Leo, and M. Bellingeri. 2012. Consequences of extreme events on population persistence and evolution of a quantitative trait. Ecological Informatics 8: 20-28. Ward, D. 2009. The biology of deserts. Oxford University Press, New York. Watts, C. H. S. and H. J. Aslin. 1981. The rodents of Australia. Angus & Robertson, Sydney, Australia. Welbergen, J. A., S. M. Klose, N. Markus, and P. Eby. 2008. Climate change and the effects of temperature extremes on Australian flying-foxes. Proceedings of the Royal Society B: Biological Sciences 275: 419-425. Xu, D., Z. Feng, L. J. S. Allen, and R. K. Swihart. 2006. A spatially structured metapopulation model with patch dynamics. Journal of Theoretical Biology 239: 469- 481.

63

64

Chapter 3: Rat plagues in central Australia

Chapter 3: Extreme rainfall events predict irruptions of

rat plagues in central Australia

“…we remained there (Camp 63) until the 5th when we were driven out by the rats…” —Burke (1860). Photograph by Aaron Greenville.

A version of this chapter is published as Greenville A.C., Wardle G.M., Dickman C.R. 2013. Extreme rainfall events predict irruptions of rat plagues in central Australia. Austral Ecology 38: 754–764. My contribution was very substantial, and included conceptualisation of ideas, engagement in fieldwork, completion of data manipulations and analyses, and writing and editing drafts of all the chapters, in consultation with my co-authors. A version of this chapter was also published as a popular science article for Biology News (Appendix 6).

65

Chapter 3: Rat plagues in central Australia

Abstract

A general understanding of the factors that regulate populations remains a central goal

for ecologists, and species that increase rapidly to plague numbers followed by equally

sudden crashes in abundance are particularly intriguing. Throughout central Australia, the

long-haired rat, Rattus villosissimus, is largely absent in dry times, but can irrupt into plague

numbers after heavy rains. Using historical data we firstly relate the occurrence of plagues to

exceptionally high rainfalls, and using live-trapping records from rat irruptions in 1991 and

2011 in the eastern Simpson Desert, we then investigate the population structure at the plague front. In doing so, we ask if long-haired rat plagues irrupt from multiple refugia, if young males are first to disperse and whether the rate of dispersal is rapid. Annual rainfall in the year preceding a plague successfully predicted the probability of an outbreak, with an 80% chance of a plague occurring after an annual rainfall of 750 mm. Trapping grids closest to

drainage lines to the south of the study area had greater captures of long-haired rats.

Enhanced rat numbers in proximity to drainage areas suggested that the irruptive process

unfolded from the south and demonstrated the existence of multiple refugia. In contrast to the

resident fitness hypothesis, there was no difference in captures between the sexes or age

classes of long-haired rats. Heavier animals were caught at the invasion front, presumably

because they could travel larger distances than smaller individuals, suggesting that larger

individuals were maximising their fitness by dispersing first to new resources. Captures of

long-haired rats were highest at the beginning of irruptions in both 1991 and 2011, suggesting

rapid colonisation. Although rare in occurrence, rodent population irruptions and the resulting dispersal across large areas, which may persist for a few years, can thus extend the domain of influence of exceptional rain years well into the future.

66

Chapter 3: Rat plagues in central Australia

Introduction

One of the most challenging but elusive goals in ecology is to develop a general understanding of the factors that interact to regulate population size and dynamics. This challenge is of particular interest for populations that increase rapidly to plague numbers and then crash suddenly to low levels. For populations of pathogen, pest or invasive species, there is also considerable epidemiological or applied interest in understanding the factors that trigger irruptions. In animals, classically, both biotic and abiotic factors have been posited as responsible for driving cyclic population dynamics (Uvarov 1931, Nicholson 1933, Chitty

1960). Biotic factors include processes such as intraspecific changes in social organisation

(e.g. Cockburn 1988) or interspecific interactions arising from competition or predation (Kelt

2011, Hernández Plaza et al. 2012), whereas abiotic processes include changes in nutrient flows, edaphic conditions or weather (White 2008, Letnic and Dickman 2010). As the global climate changes, attention is focusing increasingly on the role of extreme climate events in driving changes in the size, structure and dynamics of animal populations (Smith 2011,

Greenville et al. 2012, Chapter 2).

For many irruptive animals, such as certain species of locusts and rodents, flooding rainfall events have been proposed to be the major driver of population increases (Newsome and Corbett 1975, Farrow 1979, Singleton et al. 2005). Such rainfall events stimulate increases in primary production, in turn providing pulses of resources that can elevate the reproductive success and survival of primary consumers (Farrow 1979, Pavey and Nano

2013). Extraordinary rainfall events are—by definition—infrequent and thus witnessing a plague event can be a rare experience. To complicate matters further for researchers, not all rainfall events necessarily lead to irruptions; this suggests that, for many irruptive species, combinations of factors may be at play (Carstairs 1974). In this study we use both the

67

Chapter 3: Rat plagues in central Australia historical record and field studies following exceptional rains that fell in central Australia in

1991, 2010 and 2011 to investigate the population dynamics of a native plague rodent, the long-haired rat Rattus villosissimus.

Although it has a broad distribution in the arid regions of Australia, R. villosissimus is largely absent in dry times and persists in small and localized refugia. The species is not physiologically well adapted to dry conditions, so the refugia are crucial in providing a reliable supply of green food and moist areas for shelter and digging burrows (Watts 1970,

Carstairs 1974, Newsome and Corbett 1975). After heavy rainfall rats emerge from their refuge sites and, if an irruption ensues, may penetrate a long way into the broader landscape

(Plomley 1972, Carstairs 1974, Newsome and Corbett 1975). Refugia clearly are a cornerstone of the irruption biology of R. villosissimus, but their locations remain elusive.

Plomley (1972) suggested the existence of a single major refuge area in the Barkly

Tablelands and proposed that, as rats irrupt, they disperse south along river channels into more-arid regions. Others have suggested that multiple refugia exist in the Barkly Tablelands, the Georgina-Diamantina drainage system and other rivers flowing towards

(Parker 1973, Carstairs 1974, Newsome and Corbett 1975, 1977).

In addition to uncertainty about the location of refugia, little is known of the population structure of the long-haired rat at the plague front. Such knowledge can provide important insight into the ecological processes that are associated with animal movements and plague formation. It can, for example, indicate whether dispersal is sex- or age-biased

(Wolff 1994), whether dispersers are reproductively active as they move, and even uncover evolutionary processes such as selection for increased leg length that may speed the rate of movement (Phillips et al. 2006). In R. villosissimus, the rate of dispersal has been suggested to range from a slow and diffuse drift of 1–3 km/day (Finlayson 1939, Crombie 1944) to as much as 8 km/day (Crombie 1944, Plomley 1972). There is some evidence that young

68

Chapter 3: Rat plagues in central Australia

animals are well represented at the plague front (Predavec and Dickman 1994) and that the

sex ratio is male-biased (Crombie 1944, Carstairs 1976, Predavec and Dickman 1994). These

observations accord with Anderson's (1989) resident fitness hypothesis that young males are

the predominant dispersers in rodent populations, although the distances moved by R.

villosissimus during plagues are considerably greater than those in most species that have

been studied.

The ‘greening’ of arid Australia in 2010 and 2011 stimulated widespread irruptions of

long-haired rats over much of the continental interior and provided a rare opportunity to

investigate the species invasion dynamics. Here, we firstly collate historical records of rat

plagues from central Australia and examine the effects of rainfall in triggering them. We then

use live-trapping data from the eastern Simpson Desert to describe the population structure of

plagues. We test the following predictions:

1) Large rainfall events drive long-haired rat irruptions,

2) Long-haired rats irrupt from multiple refugia,

3) Colonisation of new sites is rapid, and

4) Young males arrive first at new sites.

Our primary field data derive largely from an irruption of R. villosissimus that began in May

2011, but we also use data from an earlier irruption that began in May 1991 (Predavec and

Dickman 1994).

Methods

Study region

The study was carried in the Simpson Desert, central Australia (Fig. 3.1). This region

occupies 170 000 km2; dune fields comprise 73% of the region, with smaller areas consisting

69

Chapter 3: Rat plagues in central Australia of clay pans, rocky outcrops and gibber flats (Shephard 1992). The sand dunes run parallel in a north-south direction aligned with the prevailing southerly wind. The dunes are up to 10 m high and spaced 0.6–1 km apart (Purdie 1984). Vegetation in the interdune swales and on dune sides is predominantly spinifex grassland (Triodia basedowii) with patchy cover of small stands of gidgee trees (Acacia georginae) and other woody Acacia shrubs or mallee eucalypts; low-lying clay pans fill with water temporarily after heavy rain.

During summer, daily temperatures usually exceed 40º C and minima in winter often fall below 5º C (Purdie 1984). Highest rainfall occurs in summer, but heavy rains can fall locally or regionally throughout the year (Chapter 2). Long-term weather stations in the study area are at Glenormiston (1890–2011), Boulia (1888–2011) and Birdsville (1954–2011) (Fig.

3.1), and have median annual rainfalls of 186 mm (n = 122 years), 216.2 mm (n = 126 years), and 153.1 mm (n = 120 years), respectively (Bureau of Meteorology 2012). For all statistical tests a significance value of P < 0.05 was used.

70

Chapter 3: Rat plagues in central Australia

Figure 3.1: Location of study sites across Cravens Peak Reserve, Carlo Station and Ethabuka Reserve, Simpson Desert. Insert shows locations of Bureau of Meteorology weather stations used for predicting rat plagues.

71

Chapter 3: Rat plagues in central Australia

Prediction 1: Large rainfall events drive long-haired rat irruptions

The years when long-haired rat irruptions occurred (defined as animals expanding their range and becoming abundant enough for people to report their occurrence) were collated from published papers (Crombie 1944, Plomley 1972, Carstairs 1974, Newsome and

Corbett 1975, Carstairs 1976, Predavec and Dickman 1994, Letnic and Dickman 2006), technical reports, newspaper articles, local informants and anecdotal records from the Arid

Zone Research Chapter, Ecological Society of Australia. An irruption-year was scored as “1” and a non-irruption year was scored as “0” for each of five weather stations closest to the records of the outbreak (Boulia (1888–2011), Alice Springs (1878–2011), Birdsville (1954–

2011), Glenormiston (1890–2011) and Ethabuka (1995–2011)). To avoid false negative results (i.e. irruption-years that were not reported), we searched newspapers and other local sources in plague areas, focusing in particular on periods of high rainfall when pulses of primary productivity would be expected. We tested for an irruption-rainfall relationship using a quasi-binomial generalised linear model with a logit-link function, regressing annual rainfall in the year prior to an irruption against the occurrence (1) or absence (0) of an irruption. The quasi-binomial distribution was chosen to account for overdispersion (Zuur

2009) and tested using the F-statistic in an Analysis of Deviance (Zuur 2009). Analyses were performed in R 2.14.1 (R Development Core Team 2011).

Prediction 2: Long-haired rats irrupt from multiple refugia

Live-trapping was carried out at 12 sites on Carlo Station, Tobermorey Station,

Cravens Peak and Ethabuka Reserves in the Simpson Desert, south-western Queensland,

Australia (Fig. 3.1). Long-haired rats were live-trapped using pitfalls (16 cm diameter, 60 cm deep) each equipped with a 5 m drift fence of aluminium flywire to increase trap efficiency

72

Chapter 3: Rat plagues in central Australia

(Friend et al. 1989). Pitfalls were arranged on grids comprising six lines of six pitfall traps

spaced 20 m apart to cover 1 ha (i.e. 36 pitfall traps per grid). The top line of traps was

positioned on a dune crest and the bottom line 100 m distant in the swale so that each grid

sampled the topography of the dune field. Additional grids comprising three lines of pitfall

traps located in patches of gidgee (Acacia georginae) woodland in the swales were also

sampled. Sites contained 2–12 grids and were spaced at least 10 km apart (Fig. 3.1). Grids within sites were set 0.5–2 km apart in randomly chosen positions along access tracks.

Traps were opened for 2–6 nights, 4–6 times a year from 1990–2012 (see Dickman et al. 2010, 2011 for details), but long-haired rats were only captured in 1991 and 2011..

Sampling in 1991 was carried out at one site (Main Camp, total 12 grids) in April, May, July

and October, and in 2011 at four sites (total 10 grids) in May, seven sites (total 14 grids) in

June, two sites (total four grids) in August, seven sites (total 33 grids) in September and four

sites (total eight grids) in November. On each sampling occasion animals were identified,

weighed, individually marked by ear notching, and sex and reproductive condition were

determined as in Dickman et al. (1999, 2001). Recaptures were removed and minimum

numbers of animals alive were used in the analyses.

Long-haired rats use drainage lines and water features to move through the

landscape from refuge sites (Plomley 1972, Newsome and Corbett 1975). Using the 2011

data, we calculated the distance and bearing of each trapping grid opened in May, June and

September to the nearest drainage line or clay pan in ArcGIS 10 (ESRI 2010). Only data from

2011were used as multiple sites were required for this aim. To test the prediction that long-

haired rats irrupt from multiple refugia, we then used a negative binomial generalised linear

model (GLM) and log-link function to regress captures against the predictors Trip and

Distance to drainage for grids trapped in 2011. An offset was used for the response variable

(captures) that retained the original scale and integer values (a requirement for the negative

73

Chapter 3: Rat plagues in central Australia

binomial distribution), while accounting for unequal trapping effort (Zuur 2009). In addition,

to assess if long-haired rats moved from multiple directions into the study region a linear

regression, with an offset for number of trapping nights, was used to regress captures against

the predictors Trip and Angle to drainage. Angles were converted to radians for analysis.

Inspection of diagnostic plots indicated that the assumptions of the models were met, and

graphing was performed using the Circular 0.4-3 package (Agostinelli and Lund 2011) in R

2.14.1 (R Development Core Team 2011).

Live-trapping results from 1991 were also interrogated to assess movement of rats

into the study region on the finer spatial scale afforded by sampling at a single site, Main

Camp. Here, the 12 available grids were arranged in a north-south direction and spaced 1–2

km apart over ~14 km. Captures of rats were tallied for each grid and a chi-squared test with

1000 randomisations was used to test if there was a difference in captures between grids.

Analysis was performed in R 2.14.1 (R Development Core Team 2011).

Prediction 3: Colonisation of new sites will be rapid

Captures in 1991 and 2011 were used to assess the rate of colonisation of new sites.

As trapping in 1991 occurred repeatedly on the same grids at one site, the colonisation rate

could be assessed on a fine spatial and temporal scale. However, as different sites were

trapped in 2011 the colonisation rate may have been confounded by spatial differences in

when rats first arrived. To reduce bias we pooled captures for each trapping month for all

sites in 2011. To investigate if there was a difference in captures between trapping months we

used a negative binomial GLM and log-link function. The number of trapping nights per grid

was used as an offset to account for unequal trapping effort between sites and trips. The years

1991 and 2011 were analysed separately, and inspection of the diagnostic plots indicated that

74

Chapter 3: Rat plagues in central Australia the appropriate model was chosen. Analysis was performed using the MASS 7.3-16 package

(Venables and Ripley 2002), in R 2.14.1 (R Development Core Team 2011).

Prediction 4: Young males arrive first at new sites

Captures of male and female rats were collated for each trapping grid in May 2011 as animals appeared to be first arriving in the study area. We designated animals as juveniles, sub-adults or adults based on size. Taylor and Horner (1973a) noted that laboratory- maintained female R. villosissimus can produce litters at about two months of age and that males can sire young from 64 days; animals then weigh ~60 g. However, Taylor and Horner

(1973b) gave 105 g as the minimum mass for adult R. villosissimus. Our raw trapping data confirmed that both sexes could be reproductively active (pregnant/parous or with descended testes) from ~60 g, but also that such early reproduction occurred very rarely; most animals

>100 g, by contrast, were or had been reproductively active. Hence, we designated animals weighing ≤60 g as juveniles, animals weighing 60–100 g as sub-adults and ≥100 g as adults.

The same criteria were adopted by D'Souza et al. (2013). We also scored reproductive condition separately as non-parous (animal has never bred), parous (animal has bred before) and pregnant/lactating for females, and non-scrotal (testes fully retracted), semi-scrotal (testes partially descended) and scrotal (testes fully descended) for males. Three separate negative binomial GLMs were used to regress the numbers of captures against 1) sex (male or female) and site; 2) age class (juvenile + sub-adult or adult) and site; and 3) reproductive category and site. Diagnostic plots were inspected to confirm model fit in each GLM.

Captures also were pooled across sites for each trip to investigate if the mean mass of individual long-haired rats was smaller in May 2011 than in subsequent months. As mass differs between the sexes, a two-factor ANOVA, with Trip and Sex as fixed factors, was used

75

Chapter 3: Rat plagues in central Australia to test for any significant relationships. As above, diagnostic plots were inspected to assess the model fit and the analyses were performed in R 2.14.1 (R Development Core Team

2011).

Results

Prediction 1: Large rainfall events drive long-haired rat irruptions

Long-term records (~100 years) from the regions around Alice Springs, Birdsville,

Boulia and Glenormiston revealed eight, two, eight and four irruptions of long-haired rats, respectively (Table 3.1). Two irruptions occurred on Ethabuka, in 1991 and 2011, with only a few animals (< 3) caught on Ethabuka in 2000 (Dickman et al. unpublished data). Annual rainfall prior to an outbreak was a strong predictor of an irruption (GLM ANODEV: F = 179,

P << 0.001). The probability of an irruption increased greatly after 600 mm rainfall and at

750 mm there was an 80% chance of an outbreak (Fig. 3.2).

76

Chapter 3: Rat plagues in central Australia

Figure 3.2: Predicted probability of a long-haired rat, Rattus villosissimus, irruption from annual rainfall the year before. Rat plagues were collated from records within the last 100 years for Alice Springs, Glenormiston, Boulia and Ethabuka Reserve. Predicted results from a quasi-binomial generalised linear model. Grey triangle represent 2010 rainfall event. Dashed lines represent standard errors.

77

Chapter 3: Rat plagues in central Australia

Table 3.1: Year and location of each long-haired rat, Rattus villosissimus, irruption from historical records. ESA AZRC = Ecological Society of Australia Arid Zone Research Chapter.

Year Weather Station Rainfall Source (mm)

1887 Boulia 655.3 Plomley (1972) 1917 Boulia 574.7 Plomley (1972) 1920 Boulia 644.4 Crombie (1944) 1939 Boulia 308.7 Plomley (1972) 1950 Boulia 798.6 Plomley (1972) 1991 Boulia 410.4 Predavec and Dickman (1994) 2000 Boulia 746.8 Local residents 2010 Boulia 363.5 Local residents 1904 Alice Springs 335.1 Plomley (1972) 1968 Alice Springs 502.3 Carstairs (1976) 1974 Alice Springs 902.9 ESA AZRC 1975 Alice Springs 684 ESA AZRC 1986 Alice Springs 432.6 ESA AZRC 1995 Alice Springs 353.4 ESA AZRC 2001 Alice Springs 741.2 ESA AZRC 2010 Alice Springs 769.6 ESA AZRC 1916 Birdsville 541.8 Crombie (1944) 1966 Birdsville 311.9 Plomley (1972) 1991 Ethabuka Station 513.7 Predavec and Dickman (1994) 2000 Ethabuka Station 650.6 Dickman et al. unpublished data 2011 Ethabuka Station 574.2 Dickman et al. unpublished data 1974 Glenormiston 677.2 Letnic and Dickman (2006) 1991 Glenormiston 665.4 local residents 2000 Glenormiston 632.9 local residents 2010 Glenormiston 578.5 local residents CRD

78

Chapter 3: Rat plagues in central Australia

Prediction 2: Long-haired rats irrupt from multiple refugia

Captures of long-haired rats in May and July 1991 averaged three and four

individuals, respectively, and October 1991 averaged one individual per 100 trap nights. In

May, June, August, September and November 2011 there were 3, 42, 17, 4 and 5 individuals

per 100 trap nights, respectively. At the beginning of the irruption in May 1991, more long-

haired rats were captured on southerly than on northerly grids (χ2 = 18.55, P = 0.04, Fig. 3.3).

There was a significant interaction with trip and distance to nearest drainage line (Table 3.2).

In May 2011 there was a weak negative relationship with captures and distance to drainage

lines; this became strong in June 2011 but showed no relationship in September 2011 (Fig.

3.4a). Thus, there is some support for long-haired rats using drainage lines earlier in an irruption. There was also an interaction between trip and direction to drainage line (Table

3.3). Captures on southern grids that were closest to drainage lines were slightly higher in

May 2011 and much higher on these grids in June 2011; no relationship was evident in

September 2011 (Fig. 3.4b). This suggests that rats moved into the study region mostly from the south as the irruption began in 2011.

79

Chapter 3: Rat plagues in central Australia

Table 3.2: Analysis of deviance results for the effect of distance to nearest drainage line on live-trapping captures of the long-haired rat, Rattus villosissimus, Simpson Desert, central Australia in 2011. A negative binomial generalized linear model was used, with log-link function. Deviance follows the chi-square distribution. * indicates significance P<0.05.

Factor df Deviance Residual df Residual deviance P Distance Trip 2 16667 45 21716 <<0.01* Distance 1 187 44 4863 <<0.01* Trip × Distance 2 126 42 4736 <<0.01*

Figure 3.3: Captures of long-haired rats, Rattus villosissimus, for each grid at Main Camp in May 1991, Simpson Desert, Queensland. Higher captures were found on grids in the south at the beginning of the 1991 irruption.

80

Chapter 3: Rat plagues in central Australia

Figure 3.4: Captures of long-haired rat (per grid), Rattus villosissimus, from a) distance and b) direction (angle) to nearest drainage line, Simpson Desert, Queensland. Points in (a) represents observed captures and lines represent predicted captures from GLM. Points on circle in (b) represent trapping grids and length of arrows are proportional to the numbers of captures for each grid.

81

Chapter 3: Rat plagues in central Australia

Table 3.3: Analysis of variance results for the effect of direction (angle) to nearest drainage line on live-trapping captures of the long-haired rat, Rattus villosissimus, Simpson Desert, central Australia. * indicates significance P<0.05.

Factor df SS F-value P-value Trip 2 11338 13.3 0.008* angle 1 3277 7.7 <<0.01* Trip × Angle 2 3334 3.9 0.03*

Prediction 3: colonisation of new sites will be rapid

Capture numbers differed between trips in 1991 (deviance = 392, df = 2, residual

deviance = 6687, df = 32, P << 0.001) and in 2011 (deviance = 12597, df = 4, residual deviance = 6997, df = 59, P << 0.001). In 1991 captures were greatest at the beginning of the plague and declined throughout the year (Fig. 3.5a). However, as all the southern-most grids that captured rats in May 1991 were trapped in April 1991, full colonisation clearly occurred within a month. In general, captures were low at the start of the plague in 2011 and increased rapidly by June 2011 (Fig. 3.5b).

82

Chapter 3: Rat plagues in central Australia

Figure 3.5: Predicted captures per grid from negative binomial generalised linear models for captures of long-haired rats, Rattus villosissimus, across each trip in a) 1991 irruption and b) 2011 irruption, Simpson Desert, Queensland. Lines indicate standard errors.

Prediction 4: Young males arrive first at new sites

More adult than sub-adult or juvenile long-haired rats were captured in May 2011, and most captures occurred at one site, Field River South (Table 3.4). There was no significant difference in captures of long-haired rats by trip or between the sexes (Table 3.5).

83

Chapter 3: Rat plagues in central Australia

However, there was a site effect for females when animals were categorised by reproductive condition (Table 3.6, Fig. 3.6a), and also more scrotal than non-scrotal or semi-scrotal males at all sites (Table 3.6, Fig. 3.6b). The body masses of long-haired rats differed by trip and by sex (Table 3.7). Males were heaviest in all months, and heavier individuals occurred at the beginning of the irruption in May 2011 compared to later months in the year (Fig. 3.7).

Table 3.4: ANOVA results for the influence of age class (adult or sub-adult) and site on captures of the long-haired rat, Rattus villosissimus, Simpson Desert, central Australia. *indicates significance P<0.05.

Factor df Deviance df Residual Residual deviance P Age class 1 5.31 18 20.57 0.02*

Site 3 7.89 15 12.69 0.05*

Age class × Site 3 7.00 12 5.69 0.07

Table 3.5: Analysis of deviance results for the influence of sex (male or female) and site on captures of the long-haired rat, Rattus villosissimus, Simpson Desert, central Australia.

Factor df Deviance df Residual Residual deviance P Sex 1 1.01 18 27.16 0.32

Site 3 5.30 15 21.86 0.15 Sex × Site 3 3.01 12 18.84 0.39

84

Chapter 3: Rat plagues in central Australia

Table 3.6: Analysis of deviance results for reproductive condition of female and male long- haired rats, Rattus villosissimus. Reproductive condition non-parous, parous and pregnant/lactating female captures and non-scrotal, semi-scrotal and scrotal male captures in May 2011. *indicates significance P<0.05.

Factor df Deviance df Residual Residual P deviance Females Reproductive condition (RC) 2 3.14 18 25.76 0.08 Site 3 9.00 15 16.75 0.03* Site × RC 6 4.98 12 11.77 0.17

Males Reproductive condition (RC) 2 15.55 27 11.98 <<0.01* Site 3 1.20 24 10.79 0.63 Site × RC 6 3.18 18 1.61 0.79

Figure 3.6 : Mean captures (per grid) of each category of reproductive condition for a) female and b) male long-haired rats, Rattus villosissimus, for each site at the beginning of the 2011 irruption (May 2011), Simpson Desert, Queensland. Each site sampled two grids, except Main Camp which had four. Lines indicate standard errors. 85

Chapter 3: Rat plagues in central Australia

Table 3.7: ANOVA results investigating if body mass of the long-haired rat, Rattus villosissimus, differs across trips and between sex, Simpson Desert, central Australia. *indicates significance P<0.05.

Factor df SS F-value P Trip 4 141801 24.71 <<0.01* Sex 2 186904 65.13 <<0.01*

Trip × Sex 5 12943 1.80 0.11

Figure 3.7: Mean mass (g) of male and female long-haired rats, Rattus villosissimus, captured during the 2011 irruption, Simpson Desert, Queensland. Sites pooled for each trip. Numbers above bars represent total captures of males and females. Lines indicate standard errors.

86

Chapter 3: Rat plagues in central Australia

Discussion

This study aimed to quantify one of the drivers of R. villosissimus irruptions in central

Australia and describe the population structure of a rat plague in the eastern Simpson Desert.

In doing so, we tested the following predictions: 1) large rainfall events drive long-haired rat

irruptions, 2) long-haired rats irrupt from multiple refugia, 3) colonisation of new sites will be

rapid, and 4) young males arrive first at new sites. Following from 4) we predicted that non-

breeding and smaller individuals (lower body mass) would arrive first in the study region. We

found that long-haired rat irruptions were rare events during the last 100 years, and that they

were driven by flooding rain events in central Australia.

Prediction 1: large rainfall events drive long-haired rat irruptions

Long-term records (max 133 years) from the region around Alice Springs, Birdsville,

Boulia and Glenormiston recorded eight, two, eight and four irruptions, respectively. Live-

trapping records (22 years) have recorded two irruptions in the eastern Simpson Desert.

Taken together, the historical and observational records demonstrate the rare nature of

irruptive events in central Australia. Annual rainfall the year before the plague significantly

predicted the probability of an irruption, with the probability increasing rapidly after a 600

mm rainfall event and at 750 mm there was an 80% chance of a plague. A rainfall event of

this size represents an extreme event (>95th quantile) for central Australia (Greenville et al.

2012, Chapter 2).

Past studies have found an inconsistent relationship between captures and rainfall.

Carstairs (1974) found that not all large rainfall events lead to rat irruptions. In contrast, however, Newsome and Corbett (1975) and Letnic and Dickman (2006) identified that large

87

Chapter 3: Rat plagues in central Australia

rainfall events were associated with rat plagues. This study supports the notion that flooding

rainfall events are good predictors of rat irruptions, and proposes that the observations of

Carstairs (1974) may be due to some rainfall events not reaching a critical threshold for the population to respond. For example, rodent irruptions in the Simpson Desert occurred only

after annual rainfall events above the 95th quantile (Greenville et al. 2012, Chapter 2).

Prediction 2: long-haired rats irrupt from multiple refugia

Plomley (1972) suggested that the long-haired rat irrupts from a single refuge, the

Barkly Tablelands, to the north of the study region and disperses along drainage lines and desert rivers into drier parts of central Australia (Plomley 1972, Newsome and Corbett 1975).

There was a significant interaction between distance to drainage lines and trip on captures of

long-haired rats, with the relationship strongest at the beginning and peak of the irruption,

providing support for the use of drainage lines. There was a significant interaction between

bearing (direction) and trip on captures of long-haired rats, with trapping grids in the south of

the study area that were close to drainage lines recording the highest captures in the

beginning and peak of the plague. In addition, trapping records from 1991 were used to study

movements at a finer spatial scale, and showed that rats moved into the study region from the

south. Taking these observations together, it is apparent that long-haired rats dispersed along

drainage lines and entered the study region from the south. This movement pattern suggests

that additional refugia exist to those in the north of the study region that were proposed by

Plomley (1972).

88

Chapter 3: Rat plagues in central Australia

Prediction 3: colonisation of new sites will be rapid

In general, the 2011 long-haired rat plague started with low captures in the eastern

Simpson Desert region, but within a month total captures increased some 16 times (Fig 3.5).

The previous plague in 1991 also recorded the highest captures at the beginning of the

irruption, suggesting that colonisation is rapid.

Little is known about the number of individuals that are present at the plague front and at what speed this species moves across the landscape. For example, Crombie (1944) suggested that rats move 1.6 km per day, Finlayson (1939) recorded approximately 3 km per day and Plomley (1972) suggested rapid movements of up to 8 km per day. Long-haired rats irrupted quickly after the flooding rains of 2010 and were recorded on the Mulligan River in

May 2010, 17 km east of the Simpson Desert (M. Tischler, unpublished data, 2010),

suggesting local irruptions or rapid movement along the river channels. Approximately one

year later this species was recorded in the sand dunes of the Simpson Desert (this study),

suggesting a gradual drift along drainage lines into the sand dunes. Consequently, we suggest

that large numbers of individuals are at the plague front, and movements may be rapid, but

the tempo of movements will be influenced by the availability of suitable habitat or corridors

and thus may vary for each event and location.

Prediction 4: young males arrive first at new sites.

In contrast to our predictions, adult long-haired rats were captured in greater numbers

than sub-adults at the beginning of the plague (May 2011). Larger (greater body mass)

individuals were more numerous at the beginning of the plague, but there was no support for

a difference in the sex of arriving rats. In early stages of the plague, scrotal males were

89

Chapter 3: Rat plagues in central Australia

captured in greater numbers, in contrast to females that did not differ in reproductive

condition. Similar demographic results were reported by D'Souza et al. (2013) for a long-

haired rat irruption near Longreach in 2011.

Past studies have showed contrasting results for the sex of dispersing individuals.

Carstairs (1976) found a male bias in dispersers, however Predavec and Dickman (1994)

found no bias between numbers of male or female dispersers, which is consistent with our

findings. In addition, large adult long-haired rats were caught in greater numbers at the

beginning of the irruption throughout the study region, presumably because they are able to

move larger distances than smaller individuals. Such differences in size of individuals at an

invasion front have been reported for other species, such as the invasive cane toad (Rhinella

marina), with longer legged individuals being first to be observed in new areas (Phillips et al.

2006). Given the above, it is unlikely that individuals were dispersing from refugia to avoid

inbreeding (Wolff 1994) and this is not consistent with Anderson's (1989) resident fitness hypothesis, but most likely because larger individuals could maximise fitness by taking advantage of the growth of new vegetation and burrowing sites created by the flooding rains.

Although rare in occurrence, rodent population irruptions and the resulting dispersal

across large areas, which may persist for a few years, can thus extend the domain of influence

of exceptional rain years well into the future.

Acknowledgements

We thank Bush Heritage Australia, H. Jukes, G. McDonald, D. Smith and G. Woods

for allowing access to the properties in the study region; M. Tischler for valuable discussions

of rat movements; members of the Desert Ecology Research Group, B. Tamayo and C.-L.

90

Chapter 3: Rat plagues in central Australia

Beh and many volunteers for valuable assistance in the field. Funding was provided by the

Australian Research Council and an Australian Postgraduate Award (to ACG) and the

Australian Government’s Terrestrial Ecosystems Research Network (www.tern.gov.au), an

Australian research infrastructure facility established under the National Collaborative

Research Infrastructure Strategy and Education Infrastructure Fund - Super Science Initiative through the Department of Industry, Innovation, Science, Research and Tertiary Education.

91

Chapter 3: Rat plagues in central Australia

References:

Agostinelli, C. and U. Lund. 2011. R package 'circular': circular statistics (version 0.4-3). Anderson, P. K. 1989. Dispersal in rodents: a resident fitness hypothesis. American Society of Mammalogists, Provo, Utah. Bureau of Meteorology. 2012. Climate data online, accessed 20 June 2012, http://www.bom.gov.au/climate/data/ Burke, R. O. 1860. Robert O'Hara Burke's dispatch, Cooper's Creek, 13 December. MS 13071, Box 2082/1a, Item 13. Records of the Burke & Wills Expedition. Australian Manuscripts Collection, State Library of Victoria, Victoria. Carstairs, J. 1974. The distribution of Rattus villosissimus (Waite) during plague and non- plague years. Wildlife Research 1: 95-106. Carstairs, J. 1976. Population dynamics and movements of Rattus villosissimus (Waite) during the 1966-69 plague at Brunette Downs, N.T. Wildlife Research 3: 1-9. Chitty, D. 1960. Population processes in the vole and their relevance to general theory. Canadian Journal of Zoology 38: 99-113. Cockburn, A. 1988. Social behaviour in fluctuating populations, Croom Helm, London. Crombie, A. C. 1944. Rat plagues in western Queensland. Nature 152: 803-804. D'Souza, J. B., A. Whittington, C. R. Dickman, and L. K. P. Leung. 2013. Perfect storm: demographic responses of an irruptive desert mammal to prescribed burns following flooding rain. Austral Ecology 38: 765–776. Dickman, C. R., A. C. Greenville, C.-L. Beh, B. Tamayo, and G. M. Wardle. 2010. Social organization and movements of desert rodents during population “booms” and “busts” in central Australia. Journal of Mammalogy 91: 798-810. Dickman, C. R., A. C. Greenville, B. Tamayo, and G. M. Wardle. 2011. Spatial dynamics of small mammals in central Australian desert habitats: the role of drought refugia. Journal of Mammalogy 92: 1193-1209. Dickman, C. R., A. S. Haythornthwaite, G. H. McNaught, P. S. Mahon, B. Tamayo, and M. Letnic. 2001. Population dynamics of three species of dasyurid marsupials in arid central Australia: a 10-year study. Wildlife Research 28: 493-506. Dickman, C. R., P. S. Mahon, P. Masters, and D. F. Gibson. 1999. Long-term dynamics of rodent populations in arid Australia: the influence of rainfall. Wildlife Research 26: 389-403. ESRI. 2010. ArcGIS 10.0. ESRI Inc, California.

92

Chapter 3: Rat plagues in central Australia

Farrow, R. 1979. Population dynamics of the Australian plague locust, Chortoicetes terminifera (Walker), in central western New South Wales. I. Reproduction and migration in relation to weather. Australian Journal of Zoology 27: 717-745. Finlayson, H. 1939. On mammals of the . Part V. Transactions of the Royal Society of 63: 348-353. Friend, G. R., G. T. Smith, D. S. Mitchell, and C. R. Dickman. 1989. Influence of pitfall and drift fence design on capture rates of small vertebrates in semi-arid habitats of Western Australia. Australian Wildlife Research 16: 1-10. Greenville, A. C., G. M. Wardle, and C. R. Dickman. 2012. Extreme climatic events drive mammal irruptions: regression analysis of 100-year trends in desert rainfall and temperature. Ecology and Evolution 2: 2645-2658. Hernández Plaza, E., L. Navarrete, C. Lacasta, and J. L. González-Andujar. 2012. Fluctuations in plant populations: role of exogenous and endogenous factors. Journal of Vegetation Science 23: 640-646. Kelt, D. A. 2011. Comparative ecology of desert small mammals: a selective review of the past 30 years. Journal of Mammalogy 92: 1158-1178. Letnic, M. and C. R. Dickman. 2006. Boom means bust: interactions between the El Niño/Southern Oscillation (ENSO), rainfall and the processes threatening mammal species in arid Australia. Biodiversity and Conservation 15: 3847-3880. Letnic, M. and C. R. Dickman. 2010. Resource pulses and mammalian dynamics: conceptual models for hummock grasslands and other Australian desert habitats. Biological Reviews 85: 501-521. Newsome, A. E. and L. K. Corbett. 1975. Outbreaks of rodents in semi-arid and arid Australia: causes, preventions and evolutionary considerations, Pages 117-153 in I. Prakash and P. K. Ghosh, editors. Rodents in desert environments. Dr. W. Junk Publishers, The Hague. Newsome, A. E. and L. K. Corbett. 1977. The elusive rat plague ECOS 1977. ECOS 10-13. Nicholson, A. J. 1933. Supplement: the balance of animal populations. Journal of animal Ecology 2: 131-178. Parker, S. A. 1973. An annotated checklist of the native land mammals of the . Records of the South Australian Museum 16: 1-57. Pavey, C. R. and C. E. M. Nano. 2013. Changes in richness and abundance of rodents and native predators in response to extreme rainfall in arid Australia. Austral Ecology 38: 777–785. 93

Chapter 3: Rat plagues in central Australia

Phillips, B. L., G. P. Brown, J. K. Webb, and R. Shine. 2006. Invasion and the evolution of speed in toads. Nature 439: 803 - 803. Plomley, N. J. B. 1972. Some notes on plagues of small mammals in Australia. Journal of Natural History 6: 363-384. Predavec, M. and C. R. Dickman. 1994. Population dynamics and habitat use of the long- haired rat (Rattus villosissimus) in south-western Queensland. Wildlife Research 21: 1-10. Purdie, J. L. 1984. Land systems of the Simpson Desert region. Natural resources series no. 2. CSIRO Division of Water and Land Resources, Melbourne, Australia. R Development Core Team. 2011. R: A language and environment for statistical computing. R Foundation for Statistical Computing., Vienna, Austria. Shephard, M. 1992. The Simpson Desert: natural history and human endeavour. Royal Geographical Society of Australasia, Adelaide, Australia. Singleton, G. R., P. R. Brown, R. P. Pech, J. Jacob, G. J. Mutze, and C. J. Krebs. 2005. One hundred years of eruptions of house mice in Australia – a natural biological curio. Biological Journal of the Linnean Society 84: 617-627. Smith, M. D. 2011. The ecological role of climate extremes: current understanding and future prospects. Journal of Ecology 99: 651-655. Taylor, J. M. and B. E. Horner. 1973a. Reproductive characteristics of wild native Australian Rattus (Rodentia : Muridae). Australian Journal of Zoology 21: 437-475. Taylor, J. M. and B. E. Horner. 1973b. Results of the Archbold Expeditions. No. 98 Systematics of native Australian Rattus (Rodentia, Muridae). Bulletin of the American Museum of Natural History 150: 1-130. Uvarov, B. P. 1931. Insects and climate. Transactions of the Royal Entomological Society of London 79: 1-232. Venables, W. N. and B. D. Ripley. 2002. Modern Applied Statistics with S. Fourth Edition edition. Springer, New York. Watts, C. H. S. 1970. The foods eaten by some Australian desert rodents. South Australian Naturalist 44: 71-74. White, T. C. R. 2008. The role of food, weather and climate in limiting the abundance of animals. Biological Reviews 83: 227-248. Wolff, J. O. 1994. More on juvenile dispersal in mammals. Oikos 71: 349-352. Zuur, A. F. 2009. Mixed effects models and extensions in ecology with R. Springer, New York ; London. 94

Chapter 4: Mammal population dynamics

Chapter 4: Spatial and temporal dynamics in mammal

populations: the role of intrinsic and extrinsic factors

The mulgara (Dasycercus blythi) is a keystone predator in the Simpson Desert. Photograph by Aaron Greenville.

95

Chapter 4: Mammal population dynamics

Abstract

Intrinsic and extrinsic forces can influence the population dynamics of species across space such that sub-populations either fluctuate synchronously or exhibit distinct sub- population signatures. In this study we test five possible spatial population structures of five species of small mammals. We then use the best fitting spatial models to incorporate drivers that may regulate these populations. Using Moran’s theorem, we predicted that species with synchronous spatial dynamics are driven by factors that operate at the landscape scale, such as resource-pulses from large-scale rainfall events or wildfires, whereas species with asynchronous sub-populations will be influenced largely by factors operating at local scales.

We use multivariate state-space models to investigate the spatial population structures of the study species using 17–22 years of intensive live-trapping data from nine spatially distinct sites in central Australia. Sub-populations were synchronous or had two structures if driven by large-scale processes, but were asynchronous if driven by local events. Density dependence was detected in all species, but was weakest in insectivorous dasyurid marsupials. Our findings suggest that local environmental stochasticity is more important than intrinsic factors in driving the population dynamics of small dasyurid species. In contrast, populations of rodents and a large carnivorous dasyurid were driven by both extrinsic and intrinsic factors that operate at the landscape scale, confirming predictions derived from Moran’s theorem.

96

Chapter 4: Mammal population dynamics

Introduction

Understanding the temporal and spatial dynamics of species’ populations is important

both to uncover the drivers that influence demography and to identify populations that are

vulnerable to extinction (Ward et al. 2010). The causes of temporal variation in population

size have been much studied (Dickman et al. 1999, 2001, Shenbrot et al. 2010), but in widely distributed species these drivers may vary across space so that distinct sub-population structures occur regionally or across the geographical range (Hinrichsen and Holmes 2009,

Ward et al. 2010). Spatial sub-structuring of populations may be dictated extrinsically by the local environment or by factors intrinsic to local populations; sub-structuring may persist even in meta-populations where dispersal occurs (Ranta et al. 2006). In contrast, if a driver is common and widespread, all sub-populations may be synchronized over large regional areas; this is the Moran effect (Moran 1953, Bjørnstad et al. 1999).

How intrinsic and extrinsic factors interact to influence species’ population dynamics is important for understanding spatial synchrony. For example, intrinsic factors, such as density dependence, commonly influence population growth rate at local scales (Bateman et al. 2011). Local environmental conditions, such as access to drought refugia (Dickman et al.

2011), can drive population dynamics such that sub-populations are independent from each other. However, multiple sub-populations can also interact with each other if they are close enough and share similar environmental conditions. For example, populations that fall within the same rainfall zone may experience similar levels of productivity (Woodman et al. 2006), or populations that experience the same disturbance event, such as wildfire (Letnic et al.

2004, Evans et al. 2010), may have similar population trajectories. Populations that occur in locations with similar resources, such as near ephemeral water, may exhibit increases in survival or support greater densities of animals than populations in more xeric sites (Kok and

97

Chapter 4: Mammal population dynamics

Nel 1996), and this in turn could be expected to influence each sub-population in a similar way. Lastly, if a driver such as a widespread climatic event is experienced across a species’ geographical range, all sub-populations may exhibit the same dynamics. For example, rodent populations often show ‘boom’ and ‘bust’ dynamics in arid regions, driven by rapid shifts in primary productivity from flooding rains and subsequent drought events (Dickman et al.

1999, Greenville et al. 2013, Chapter 3).

Once the spatial structure of a population is identified, covariates can be investigated

to explore their contribution to population sub-structuring. For example, vegetation cover can

provide foraging sites or refuges from predation for terrestrial prey species (Masters 1998,

Dickman et al. 2001, Körtner et al. 2007), and subtle differences in this resource between

sub-populations may have a large influence on their growth rates, particularly if relationships

are non-linear. Measures of food availability, such as seed production for granivores, can be

important in predicting population size (Predavec 1994, 2000). Species interactions can also

affect population parameters. For example, competition from dominant species (Harris and

Macdonald 2007) and predation (Previtali et al. 2009, Brodie et al. 2013) can exert

downward pressure on populations of subordinate and prey species, respectively, and lead to

spatial synchrony of population dynamics (Ims and Andreassen 2000).

Developing accurate estimates of population size and dynamics across species’ ranges

has long been recognized as a challenge for ecologists (Clark and Bjørnstad 2004). Variation

in estimates can come about from process noise, uncertainty in population growth rates due to

environmental or demographic stochasticity, and variation due to observation error (Ward et

al. 2010, Knape et al. 2013). In addition, long-term monitoring of populations is difficult due

to short-term funding cycles or because research sites are not always accessible, thus leading

to gaps (missing data) in time series. Long-term monitoring often involves multiple observers

over the duration of the time series, which can also influence observation error (Nuno et al.

98

Chapter 4: Mammal population dynamics

2013). By partitioning both process and observation errors and incorporating these into population models, more-accurate estimates of population parameters can be made (Clark and

Bjørnstad 2004, Ward et al. 2010). A statistical framework that incorporates both types of error and handles missing data is provided by state-space models (Ward et al. 2010, Holmes et al. 2012b). These models also can be expanded to incorporate multiple sub-populations and thus used to test hypotheses about spatial structure and its determinants (Hinrichsen and

Holmes 2009, Ward et al. 2010).

In this study we test five potential spatial population structures of five species of desert-dwelling small mammals, and for each structure, include models with and without density dependence. Using multivariate state-space models, we first explore whether sub- populations of mammals are asynchronous (independent hypothesis), form two sub- populations at either ephemeral water sources or open desert sites (oasis hypothesis) or at burnt versus unburnt sites (wildfire hypothesis), form three sub-populations organized by shared rainfall gradients (productivity hypothesis), or if all sub-populations are synchronous

(single population hypothesis). Secondly, we use the best fitting models to identify potential drivers, such as local rainfall, plant cover and species interactions that may regulate the small mammal populations. Moran’s theorem states that sub-populations sharing a common structure of density dependence should be synchronized by a spatially correlated density- independent factor (Moran 1953, Ranta et al. 1995). Thus, we predict that species with synchronous spatial dynamics will be driven by factors that operate at the landscape scale, such as resource-pulses from floods or large-scale rainfall events or wildfires, whereas species with asynchronous sub-populations will be influenced largely by factors operating at local scales.

99

Chapter 4: Mammal population dynamics

Material and Methods

Study region

The study was carried out in the Simpson Desert, central Australia (Appendix 2.1,

Fig. A2.1). This region occupies 170 000 km2; dune fields comprise 73% of the region, with

smaller areas consisting of clay pans, rocky outcrops and gibber flats (Shephard 1992). The

sand dunes run parallel in a north-south direction aligned with the prevailing southerly wind,

are up to 10 m high, and spaced 0.6–1 km apart (Purdie 1984). Vegetation in the interdune

swales and on dune sides is predominantly spinifex (Triodia basedowii) grassland with small stands of gidgee trees (Acacia georginae) and other woody Acacia shrubs or mallee eucalypts; low-lying clay pans fill with water temporarily after heavy rain.

During summer, daily temperatures usually exceed 40º C and minima in winter often fall below 5º C (Purdie 1984). Highest rainfall occurs in summer, but heavy rains can fall locally or regionally throughout the year. Long-term weather stations in the study region are at Glenormiston (1890–2011), Boulia (1888–2011) and Birdsville (1954–2011), and have median annual rainfalls of 186 mm (n = 121 years), 216.2 mm (n = 123 years), and 153.1 mm

(n = 57 years), respectively (Bureau of Meteorology 2012). In general, higher rainfall is

experienced in the north of the study region than in the south, and there is a weak rainfall

gradient from east to west (Bureau of Meteorology 2012).

Small mammals

Live-trapping was carried out at nine sites across Carlo Station, Tobermorey Station,

Cravens Peak and Ethabuka Reserves, and covered an 8000 km2 area in the north-eastern

Simpson Desert in south-western Queensland (Fig. A2.1). Dickman et al. (1995) found that

100

Chapter 4: Mammal population dynamics

the maximum dispersal distance for rodents was 14 km over a 2.5 year study (mean 6.34 km),

and 12 km (mean 1.04 km) for dasyurid marsupials. Thus, each site was set > 20 km apart to

limit the effect of dispersal between sites (Fig. A2.1). Small mammals were live-trapped using pitfall traps (16 cm diameter, 60 cm deep) each equipped with a 5 m drift fence of aluminium flywire to increase trap efficiency (Friend et al. 1989). Pitfalls were arranged on grids comprising six lines of six traps spaced 20 m apart to cover 1 ha (i.e. 36 pitfall traps per grid). The top line of traps was positioned on a dune crest and the bottom line 100 m distant in the swale so that each grid sampled the topography of the dune field. Sites contained 2–12 grids and grids within sites were set 0.5–2 km apart in randomly chosen positions.

Traps were opened 2–6 times a year from 1990–2012 at one site (Main Camp) and from 1995–2012 at eight more sites (Shitty Site, Field River South, Field River North, South

Site, Kunnamuka Swamp East, Carlo, Tobermorey East and Tobermorey West). Each trap was opened for 2–6 nights. To account for unequal trapping effort, live-capture counts were standardized per 100 trap nights (TN: trap nights = traps × nights opened) and averaged for each trip. Not all sites were opened on every trip, but were included in the time-series as a missing value. Live-trapping data were log-transformed (log+1) as the population models used below are in log-space.

Long-term (17–22 years; 130 sampling trips) live-trapping data (205 524 trap nights) yielded six species of rodents: Pseudomys hermannsburgensis (sandy inland mouse; 7878 captures), Notomys alexis (spinifex hopping mouse; 5146 captures), Pseudomys desertor

(desert mouse; 1299 captures), Rattus villosissimus (long-haired rat; 1170 captures), Mus musculus (house mouse; 1120 captures), and Leggadina forresti (short-tailed mouse; 4 captures), and eight species of dasyurid marsupials: Dasycercus blythi (brush-tailed mulgara;

853 captures), Ningaui ridei (wongai ningaui; 863 captures), Sminthopsis youngsoni (lesser hairy-footed dunnart; 2491 captures), Sminthopsis hirtipes (hairy-footed dunnart; 334

101

Chapter 4: Mammal population dynamics

captures), Sminthopsis macroura (striped-faced dunnart; 17 captures), Sminthopsis crassicaudata (fat-tailed dunnart; 14 captures), Planigale gilesi (Giles’ planigale; 2 captures), and Planigale tenuirostris (narrow-nosed planigale; 12 captures). Of these species, only P. hermannsburgensis (12 g) and N. alexis (35 g), the carnivorous D. blythi (100 g), and the insectivorous S. youngsoni (10 g) and N. ridei (8 g) had sufficient time-series in both length and numbers of captures across sites to run the population models below.

Wildfire

Large scale wildfires (> 1000 km2) have occurred three times in the study region since

1972, and the mean wildfire return interval for the region is 26 years (Greenville et al. 2009,

NAFI 2013). To investigate whether wildfire affected the trajectory of sub-populations of the

study species, the wildfire history at the site of each sub-population was classified according to whether it had experienced a wildfire (six sites, see Appendix A2.1) or remained unburnt

(three sites, Appendix A2.1) across the 22-year study period.

Rainfall

Daily data from automatic weather stations (Environdata, Warwick, Queensland) at each site (Fig. 1.2) were used to calculate total rainfall for each trapping month, number of

rain days, mean rainfall per day, mean and maximum event size, and rainfall lags for each

month up to a year. Cumulative rainfall in the previous two months up to that in the previous

12 months was also calculated. Weather stations were active from 1995–2012. To obtain

rainfall data pre-1995 for Main Camp, where trapping started in 1990, we averaged daily

rainfall records from the closest weather stations at Glenormiston, Sandringham, Boulia,

102

Chapter 4: Mammal population dynamics

Bedourie and Birdsville (Bureau of Meteorology 2012). As rodents and dasyurids respond

differently to rainfall (Greenville et al. 2012, Chapter 2), to avoid a “one size fits all” approach, we selected rainfall metrics for covariates with the highest Pearson correlation with the trapping rates for each species. Thus, we used 12 months cumulative rainfall for P. hermannsburgensis and N. alexis (r = 0.48 and 0.50, respectively), eight months cumulative rainfall for N. ridei (r = −0.38) and mean rainfall event size two months prior for S. youngsoni (r = 0.41); no rainfall variables were correlated with captures of D. blythi.

Spinifex cover and seed

To measure cover of the dominant vegetation, spinifex (T. basedowii), we scored

ground cover visually as a percentage in 2.5 m radius plots around six randomly selected

traps on each small mammal trapping grid. In addition, a seed productivity index (0–5, where

0 represents no seeding and 5 represents all plants seeding profusely) was used to score the

degree of spinifex seeding in each plot. We chose spinifex seed due to its dominance in the

landscape and because it is a key component in the diet of the study rodents, representing up

to 52% of their diet by frequency of occurrence (Murray and Dickman 1994). Estimates of

cover and seed productivity were pooled for each site, per trip, over 17–22 years for each of

the nine sites.

Multivariate autoregressive state-space models

We used multivariate autoregressive state-space (MARSS) models to analyse live-

trapping data from our nine sites. MARSS models are based on the Gompertz population

growth model and assume that sub-population growth rate varies exponentially with sub-

103

Chapter 4: Mammal population dynamics

population size and that meta-populations are closed to immigration and emigration

(Hinrichsen and Holmes 2009). The MARSS framework is hierarchical and allows modelling of different spatial population structures and parameters, such as density dependence, while including both process (state) and observation variability (Ward et al. 2010). Process variability represents temporal variability in population size due to environmental and demographic stochasticity (Ward et al. 2010). Observation variability includes sampling error

(e.g. temporal changes in detectability or error resulting in only a sub-sample of the population being counted) (Hinrichsen and Holmes 2009, Ward et al. 2010). MARSS models can incorporate missing values and estimate them based on model parameters (Holmes et al.

2012b).

The process component is a multivariate first-order autoregressive process and is written in log-space (Holmes et al. 2012a, 2012b) as:

= + + ; ~ (0, ) (1)

퐗t 퐁퐗t−1 퐮 퐰푡 풘푡 푀푀푀 푸 where Xt represents a vector of all m sub-populations (up to nine sub-populations) at time t

(trip), and u is a vector of length m. B and Q are matrices that denote process parameters. The

B diagonal elements (Bii) represent the coefficients of autoregression in the populations through time and represent the strength of density dependence (diagonal element Bii = 1

represents density independence, Bii < 1 = density dependence). The off-diagonal elements in the B matrix allow for interactions between processes, such as between populations or covariates (Holmes et al. 2012a). The parameter u describes the mean growth rate of the sub- population. We allowed u to vary, as we assumed growth rates may differ across sites. wt

denotes process errors which we assumed to be independent and to follow a multivariate

normal distribution with a mean of zero and variance-covariance matrix Q (i.e. Q represents

104

Chapter 4: Mammal population dynamics

how the sub-populations are correlated to one another). We allowed Q to have different

variances across sites, as process variation was assumed to differ across space.

The observation component, written in log-space (Holmes et al. 2012a, 2012b), is:

= + + ; ~ (0, ) (2)

퐘푡 퐙퐗t 퐚 퐯푡 퐯푡 푀푀푀 퐑 where Yt represents a vector of all observations at n sites at time t, a denotes the mean bias

between sites, and Z is an n × m matrix of 0s and 1s that assigns observations to a sub-

population structure. The number of sites (n) may be different from the number of sub-

populations (m) at time t (Ward et al. 2010). Observation errors, vt,, are assumed to be

uncorrelated and follow a multivariate normal distribution, with a mean of zero and a

variance-covariance matrix R. We set R to have equal variance across sites, as the same trapping methods were employed throughout the study and thus we did not expect differences

in trappability. Equations (1) and (2) comprise the MARSS model.

The modelling approach was undertaken in two steps to reduce computational difficulties. Firstly, for each species, we tested five possible spatial population structures: independent, oasis, productivity, wildfire and single population hypotheses (each hypothesis was modelled by assigning a matrix of 1s and 0s via the Z-matrix; see Appendices A2.1 and

A2.2 for site selection and model formulation). Each hypothesis was tested with and without density dependence (with Bii set to one, assuming that density dependence was equal for all sub-populations to minimize computation when estimating many variables) in equation 1); thus ten models were compared. The best fitting model was chosen using Akaike’s

Information Criterion (AICc), adjusted for small sample size (Burnham and Anderson 2002).

Models were run with the Expectation-Maximization (EM) algorithm; if they did not

converge, the EM parameters were used as starting points for the quasi-Newton (BFGS)

algorithm to find the best estimates (Holmes et al. 2012a).

105

Chapter 4: Mammal population dynamics

Secondly, the best fitting model thus selected was expanded to include covariates that

might influence population dynamics. For P. hermannsburgensis and N. alexis, spinifex

cover, the seed productivity index from the prior trip, 12 month cumulative rainfall and

capture rates of the predatory D. blythi were included in the model. For the dasyurids D. blythi and S. youngsoni, spinifex cover was included in both models. For D. blythi, total

rodent population size was entered as rodents are a prey source (Chen et al. 1998) and for S.

youngsoni, D. blythi was entered as a potential predator and competitor (Dickman 2003). In

addition, mean rainfall event size two months prior was included in the S. youngsoni model.

For N. ridei, spinifex cover and eight months cumulative rainfall were entered as covariates.

All variables were standardized by subtracting the mean and dividing by the standard

deviation (z-scored) to allow direct comparisons between covariates to be made. Covariates

measured for each sub-population were modelled as separate processes (states) to account for

missing values and observation error, and their interaction with each species sub-population

was assessed by examining the off-diagonals of the B matrix (Holmes et al. 2012a). The

diagonal elements of the B matrix (Bii) were assessed for density dependence as above. In all

models that included covariates, we set u = 0 as numeric difficulties can arise when

estimating both u and B within the same model and each dataset has been de-meaned (i.e. via

z-score transformation so u should equal 0) (Holmes et al. 2012a). In addition, by de-

meaning the dataset and setting u = 0, we focused specifically on the effects of the covariates

on sub-population population estimates (see Appendix A2.2 for an example of model

formulation). As above, models were run with the Expectation-Maximization (EM)

algorithm, and if they did not converge the EM parameters were used as starting points for

the quasi-Newton (BFGS) algorithm (Holmes et al. 2012a).

We calculated 95% confidence intervals for each parameter using the Hessian

matrix, and considered these significant if they did not cross zero. Inspection of diagnostic

106

Chapter 4: Mammal population dynamics

plots indicated that all models met statistical assumptions. Analyses were conducted in R

2.15.3 (R Core Team 2014), using the MARSS 3.4 package (Holmes et al. 2013).

Measures of synchrony

To measure the degree of regional synchrony and compare results to the MARSS models, we calculated the zero-lag cross-correlation for the abundance (log) of the nine sub- populations for each species. The mean cross-correlation coefficients were calculated with

1000 resamples using the ncf 1.1-5 package (Bjornstad 2013), in R 3.02 (R Core Team 2014).

Results

In total, 16 617 captures of rodents and 4586 captures of dasyurids were made over the period of study, yielding trap success rates of 8.09% and 2.23%, respectively.

Rodents

Both P. hermannsburgensis and N. alexis exhibited similar spatial population structure. The best fitting model predicted that sub-populations of these rodents showed density dependence and were synchronous across all nine sites, thus conforming to the single population hypothesis (Table 4.1; Fig. 4.1a and b). In addition, the mean cross-correlation coefficients were 0.64 (CI: 0.56–0.70) and 0.65 (CI: 0.54–0.74), respectively, suggesting that regional synchrony with this effect did not decrease with distance (Fig. A2.2). Both species had positive population growth rates (u) with evidence of moderate to weak density dependence (Table 4.2). Both species also had similar process and observation errors (Table

4.2), with larger observation errors indicating similarity in population dynamics and 107

Chapter 4: Mammal population dynamics trappability. Populations of both rodents were affected positively by spinifex seed productivity on the previous trip and by 12 months cumulative rainfall, but were affected negatively by spinifex cover; N. alexis also was associated positively with the predatory mulgara (Table 4.3). The density dependence term (Bii) was stronger with covariates included for both species of rodents than without (Tables 4.2 and 4.3).

108

Chapter 4: Mammal population dynamics

Table 4.1: AICc values for MARSS models describing five possible spatial sub-population structures, with and without density dependence, for five species of small mammals in the Simpson Desert, central Australia. Data are based on 17–22 years of live-trapping from nine sites. Smallest AICc, highlighted in bold, indicates the best-fitting model.

Model Without density dependence With density dependence P. hermannsburgensis Independent 1070.73 971.59 Oasis 888.18 871.66 Productivity 933.72 910.72 Wildfire 851.37 851.37 Single population 819.244 812.57

N. alexis Independent 993.25 942.78 Oasis 860.58 849.67 Productivity 883.51 867.02 Wildfire 848.18 837.58 Single population 804.68 799.69

D. blythi Independent 623.54 570.9 Oasis 535.25 519.63 Productivity 552.06 532.22 Wildfire 518.69 495.8 Single population 506.32 495.47

S. youngsoni Independent 656.77 621.26 Oasis 644.89 638.5 Productivity 653.61 643.22 Wildfire 646.44 637.02 Single population 649.63 647.34

N. ridei Independent 465.48 444.83* Oasis 474.96 474.84 Productivity 474.16 466.42 Wildfire 491.73 488.15 Single population 491.98 497.22 *Process error (Q) set to equal across all sub-populations

109

Chapter 4: Mammal population dynamics

Table 4.2: Model estimates (95% confidence intervals) for best fitting MARSS population models (lowest AICc) for captures of small mammals (100 trap nights, log +1 transformed), Simpson Desert, central Australia. Density dependence occurs if the diagonal element of B (Bii) is less than one. There is only one B element, as density dependence was assumed to be equal for each sub-population.

Model Growth rate (u) Density dependence Process error (Qii) Observation (Bii) error (Rii) P. hermannsburgensis Single population 0.18 (0.04–0.31) 0.86 (0.76–0.96) 0.11 (0.07–0.15) 0.23 (0.2–0.26)

N. alexis Single population 0.09 (0.001–0.18) 0.89 (0.8–0.98) 0.09 (0.05–0.13) 0.23 (0.2–0.26)

D. blythi Wildfire: burnt 0.14 (0.05–0.22) 0.74 (0.56–0.86)* 0.04 (0.02–0.06) 0.11 (0.09–0.13)* unburnt 0.11 (0.03–0.19) 0.08 (0.03–0.13) Single population 0.12 (0.03–0.21) 0.75 (0.59–0.92) 0.04 (0.02–0.06) 0.13 (0.11–0.15)

S. youngsoni Independent: population 1 0.08 (0.06–0.16) 0.91 (0.83–0.93)* 0.01 (0–0.03) 0.14 (0.12–0.17)* population 2 0.08 (0.06–0.16) 0.02 (0–0.06) population 3 0.05 (0.03–0.1) 0.007 (0–0.02) population 4 0.04 (0.01–0.1) 0.01 (0–0.04) population 5 0.07 (0.03–0.13) 0.01 (0–0.03) population 6 0.05 (0.02–0.13) 0.01 (0–0.04) population 7 0.07 (0.04–0.15) 0.008 (0–0.03) population 8 0.09 (0.04–0.19) 0.04 (0–0.08) population 9 0.04 (0.006–0.09) 0.02 (0–0.08)

N. ridei Independent population 1 0.03 (0.008–0.04) 0.96 (0.93–0.98)* 0.003 (0.0002–0.005)* 0.12 (0.1–0.14)* population 2 0.02 (0.002–0.03) population 3 0.007 (−0.007–0.02) population 4 0.006 (−0.009–0.02) population 5 0.02 (0.004–0.04) population 6 0.01 (−0.003–0.03) population 7 0.01 (−0.005–0.03) population 8 0.005 (−0.009–0.02) population 9 0.007 (−0.01–0.03) *One parameter estimated for the entire model.

110

Chapter 4: Mammal population dynamics

Table 4.3: The results of covariate MARSS models for captures of small mammals (100 trap nights, log +1 transformed), Simpson Desert, central Australia. Covariates were entered into models as their own process, and the coefficients from the off-diagonals of the B matrix measure the strength of their interaction with small mammal population growth rates. Covariates were z-scored so that direct comparisons can be made. Density dependence occurs if the diagonal element Bii is less than one. There is only one B element, as density dependence was assumed to be equal for each sub-population. Each covariate was considered significant if the 95% confidence intervals (CI) did not cross zero.

Model covariates Estimate Lower CI Upper CI P. hermannsburgensis Density dependence (Bii) 0.53 0.32 0.73 12 months cumulative rainfall 0.23 0.05 0.42 Spinifex cover −0.27 −0.44 −0.1 Spinifex seed 0.53 0.28 0.77 Mulgara 0.08 −0.11 0.27

N. alexis Density dependence (Bii) 0.45 0.07 0.44 12 months cumulative rainfall 0.25 0.07 0.44 Spinifex cover −0.16 −0.29 −0.03 Spinifex seed 0.29 0.07 0.5 Mulgara 0.38 0.13 0.63

D. blythi Single population Density dependence (Bii) 0.44 0.2 0.69 Spinifex cover 0.19 0.05 0.33 Rodents 0.26 0.11 0.41

Wildfire Density dependence (Bii) 0.45 0.24 0.66 Spinifex cover 0.17 0.05 0.28 Rodents 0.27 0.14 0.4

S. youngsoni* Density dependence (Bii) 0.91† Spinifex cover −0.02 −0.02 0.05 2 month prior mean rainfall event size 1.79 0.46 3.12 Mulgara −0.097 −0.01 −0.05

N. ridei* Density dependence (Bii) 0.88 0.78 0.99 Spinifex cover 0.05 0.002 0.09 8 months cumulative rainfall −0.008 −0.06 0.04 *Process error (Q) set to be equal across all sub-populations to minimize computational difficulties that arose in estimating large numbers of parameters. †Confidence intervals could not be calculated.

111

Chapter 4: Mammal population dynamics

Figure 4.1: Predicted population size (line) and captures from each of the nine sub- populations (dots; log+1 captures/100 trap nights) from MARSS models for (a) Pseudomys hermannsburgensis, (b) Notomys alexis, (c) Dasycercus blythi single population and (d) Dasycercus blythi populations that experienced a wildfire (red line) and populations that did not experience a wildfire (blue line), Simpson Desert, central Australia. Time-series data were collected from nine sites (sub-populations) monitored 2–6 times per year for 17–22 years (1990–2012 for one site and 1995–2012 for eight sites; 130 sampling trips). Shaded areas indicate 95% confidence intervals.

112

Chapter 4: Mammal population dynamics

Dasyurids

The three species of dasyurid showed very different temporal and spatial dynamics.

Two spatial population structures were supported equally for D. blythi; one described

synchrony across all nine sites (single population hypothesis), the other described two sub-

populations in accord with the wildfire hypothesis. Both models indicated density

dependence (Table 4.1; Fig. 4.1c and d). The mean cross-correlation coefficient was 0.49 (CI:

0.34–0.65) and did not change with distance (Fig. A2.2). The process error (Q) for the best

fitting N. ridei population model approached zero, leading to poor model convergence; it was

therefore set to be equal across all nine sub-populations to minimize computational difficulties. For both this species and S. youngsoni, the best fitting model suggested that sub-

populations at all nine sites were distinct (independent hypothesis) and exhibited weak

density dependence (Table 4.1; Fig. 4.2 and 4.3). To assess the influence of missing data

from trips 1 to 30, when only the Main Camp site was sampled, we re-ran the analysis for S.

youngsoni and N. ridei by starting the time series at trip 30 (1995). The best fitting models

were for the independent hypothesis, confirming that the result was not an artefact of the

initial missing data (Table A2.1). Nonetheless, population estimates for these species prior to

1995 (trip 30) at sites where trapping was not done clearly should be viewed with caution. In

addition, the mean cross-correlation coefficients were 0.09 (−0.08–0.236) and 0.03 (−0.10–

0.17), respectively, suggesting little synchrony between sub-populations; this effect did not

change greatly over distance (Fig. A2.2).

113

Chapter 4: Mammal population dynamics

Figure 4.2: Predicted population size (line) and captures (dots; log+1 captures/100 trap nights) from a MARSS model for Sminthopsis youngsoni depicting asynchronous spatial population dynamics, Simpson Desert, central Australia. Time-series data were collected from nine sites (sub-populations) monitored 2–6 times per year for 17–22 years (1990–2012 for one site and 1995–2012 for eight sites; 130 sampling trips). Shaded areas indicate 95% confidence intervals.

114

Chapter 4: Mammal population dynamics

Figure 4.3: Predicted population size (line) and captures (dots; log+1 captures/100 trap nights) from a MARSS model for Ningaui ridei depicting asynchronous spatial population dynamics, Simpson Desert, central Australia. Time-series data were collected from nine sites (sub-populations) monitored 2–6 times per year for 17–22 years (1990–2012 for one site and 1995–2012 for eight sites; 130 sampling trips). Shaded areas indicate 95% confidence intervals.

115

Chapter 4: Mammal population dynamics

Population growth rates (u) were positive and higher, and density dependence slightly stronger, in D. blythi than in S. youngsoni and N. ridei. Growth rates for S. youngsoni were similar across all nine sub-populations, but more variable for N. ridei (Table 4.2).

Observation errors were similar across sub-populations for the three species of dasyurids, suggesting similar trappability, but process errors were lower for S. youngsoni and N. ridei than D. blythi (Table 4.2).

Dasycercus blythi was associated positively with spinifex cover and rodent numbers in both the wildfire and single population models (Table 4.3). For S. youngsoni, we set the process error (Q) and each covariate to be equal across all sub-populations to facilitate computation, and found the population to be associated positively with mean rainfall event size two months prior, and negatively with mulgara captures; no relationship was found with spinifex cover (Table 4.3). Spinifex cover positively affected populations of N. ridei, but there was no influence of prior rainfall (Table 4.3). The density dependence term (B diagonal) was stronger for D. blythi with covariates included than without, but there was little difference for either S. youngsoni or N. ridei (Tables 4.2 and 4.3).

Discussion

Our results indicate that the single population hypothesis, with density dependence, was supported for the two species of rodents, but reveal very different, and quite disparate, spatial dynamics for the three species of dasyurid marsupials. Dispersal between sites is an unlikely mechanism underpinning the single population hypothesis, as the distance between our sites was greater than the maximum dispersal distance recorded for the study species and

116

Chapter 4: Mammal population dynamics

because the distances moved by rodents decrease markedly after large rainfall events

(Dickman et al. 1995, 2010). Thus, as predicted from Moran’s theorem, species with synchronous spatial population dynamics were influenced by landscape-scale drivers.

Different covariates also were important for these two groups, and provide focus for our

discussion below.

Rodents in central Australia undergo ‘booms’ in population size that are stimulated by

increases in productivity from extreme (> 400 mm) rainfall events (Letnic and Dickman

2006, Chapter 3, Greenville et al. 2013), and then collapse again during ‘bust’ phases when conditions dry. These events typically cover >10 000 km2 (Greenville et al. 2012, Chapter 2), and most likely drove the synchronous dynamics across the sites in our study. Indeed, we found both rodent species to be associated positively with cumulative rainfall over the previous 12 months and with spinifex seed production on the prior trip. Density dependence was found in both rodents, suggesting that both intrinsic and extrinsic factors are important in regulating their population dynamics.

Despite the positive effects of spinifex seed production, both rodents responded negatively to spinifex cover. Notomys alexis is more active and is caught in open sites more readily than P. hermannsburgensis (Murray and Dickman 1994), so the response of this species to spinifex cover is not unexpected. However, P. hermannsburgensis relies on spinifex for food and shelter from predation (Murray and Dickman 1994), so a threshold level of cover is presumably needed for this species to persist. It is possible that, after an extreme rainfall event that triggers a rodent irruption, animals can reduce their use of spinifex by exploiting other grasses and cover that become available. If spinifex is used by predators such as the mulgara, snakes or varanid lizards that also increase after rain (Letnic and Dickman

2006), avoidance of spinifex would be advantageous. Alternatively, wildfires often occur in the wake of heavy rains when plants dry and die back (Letnic and Dickman 2006, Greenville

117

Chapter 4: Mammal population dynamics

et al. 2009). As fires reduce spinifex cover when rodent populations are still high, this

process may contribute to the negative relationship between rodents and cover of this

dominant vegetation.

There was no relationship between P. hermannsburgensis and D. blythi, suggesting

that predation from the carnivorous marsupial does not affect P. hermannsburgensis

populations. In contrast, there was a positive relationship between populations of N. alexis

and D. blythi. Two explanations seem plausible. Firstly, increases in rodent numbers are

likely to facilitate increases in the D. blythi population by providing more prey to hunt. The

positive relationship between D. blythi and N. alexis also may arise simply because both species share the dune crests and mid-dune areas, whereas P. hermannsburgensis is active primarily near the dune bases (Murray and Dickman 1994). Alternatively, as D. blythi is a

keystone predator in spinifex habitats (Dickman 2003), it may indirectly increase populations

of N. alexis by suppressing potential competitors such as the large (150 g) long-haired rat R. villosissimus that disperse into the study region after extreme rainfall events (Greenville et al.

2013, Chapter 3). Species manipulation experiments are needed to further explore these

possibilities.

The three species of dasyurids exhibited contrasting population dynamics across our sites. The single population and wildfire models, with density dependence, best explained the spatial dynamics of D. blythi. For both models, spinifex cover and rodent populations were associated positively with D. blythi captures, illustrating the general importance of cover and

food availability for this species. Previous research has found more D. blythi in areas with

high spinifex cover than in recently-burnt areas with low cover (Masters 1993). However,

reduced cover due to spinifex harvesting has mixed effects on this species (Masters et al.

2003, McCarthy and Masters 2005), and home ranges are little affected by wildfire when

spinifex cover is diminished (Körtner et al. 2007). Predation is higher in open areas compared

118

Chapter 4: Mammal population dynamics with mature spinifex, suggesting that spinifex provides some refuge (Körtner et al. 2007).

The interaction between wildfire and predation may lead to differing population states across space, with a strong influence from the bottom-up effects of prey.

The independent hypothesis, with weak density dependence, was the best fitting model for S. youngsoni and N. ridei. This suggests that at the nine sites surveyed, each sub- population was on its own asynchronous trajectory, influenced by local intrinsic and extrinsic factors. There was a positive effect of antecedent mean rainfall event size on S. youngsoni, but no relationship with N. ridei. Rainfall has been identified as an important predictor of S. youngsoni captures (Masters 1993), presumably as it influences food availability. In contrast to findings in previous studies, spinifex cover had no influence on S. youngsoni (Masters

1993, Dickman et al. 2001). Spinifex cover was associated positively with N. ridei, possibly because it provides shelter and foraging sites for invertebrate prey. Lastly, captures of D. blythi had a negative effect on those of S. youngsoni. Dasycercus blythi preys, in part, upon invertebrates (Chen et al. 1998) and thus may compete with other dasyurid species for food.

In addition, D. blythi is a predator of small mammals, including smaller dasyurids (Chen et al. 1998, Dickman 2003), and hence may limit S. youngsoni directly.

Density dependence commonly regulates animal populations (Brook and Bradshaw

2006, Knape and de Valpine 2012), including populations of rodents in desert habitats

(Shenbrot et al. 2010). However, perhaps because of the marked fluctuations in population size and occasional tendency for populations in local sites to crash to zero (Fig. 4.1), it has not previously been demonstrated in arid Australia. Both species of rodents and one species of dasyurid, D. blythi, showed moderate to strong density dependence after extrinsic factors were incorporated in the population models, compared to weak density dependence in models without. Several factors can affect estimates of density dependence, including the length of the time series, the stage of the population cycle when a species is surveyed (e.g. growth

119

Chapter 4: Mammal population dynamics

phase), whether the species is at carrying capacity (Brook and Bradshaw 2006), and

observation error (Knape and de Valpine 2012). Estimates of density dependence may also arise as statistical artefacts in time series as growth rates often decrease when populations are large and increase when they are small, i.e. they show a statistical return tendency to some point (or long-term mean) and thus may not be tied to an underlying biological mechanism

(Wolda and Dennis 1993). However, in this study we used multiple long-term time series

(17–22 years, totalling 130 sampling trips) for each species, which encompassed all stages of their population cycles and, by using the MARSS statistical framework, we incorporated observation error. We suggest that changes in our density dependence estimates were most likely due to including extrinsic factors, as the addition of covariates often increases the power of tests for density dependence (Rothery et al. 1997). For example, the positive effects of rainfall and spinifex seed productivity may explain irruptions in both species of rodents but, as their populations peaked and then started to decline, density dependence could strengthen as individuals compete for diminishing food resources.

In contrast to the three species above, estimates of density dependence remained similar in both covariate and non-covariate models for S. youngsoni and N. ridei populations.

Extrinsic factors, such as rainfall and potential predation from D. blythi, appeared to have

most influence on populations of S. youngsoni. Spinifex cover also was more important than

density dependence in regulating N. ridei. These contrasting patterns between species provide

further confirmation that our results are not statistical artefacts, and highlight the importance

of including extrinsic factors into models when estimating density dependence.

The spatial population dynamics of small mammals in our study system differed between and within families, illustrating varied life-history strategies employed by desert-

dwelling mammals. Sub-populations fluctuated synchronously or exhibited two structures

(single population and wildfire hypotheses) if driven by a large-scale events (e.g. rainfall,

120

Chapter 4: Mammal population dynamics

widespread pulses of food, or wildfire), or were asynchronous (independent hypothesis) if

driven by local events, with both extrinsic and intrinsic drivers contributing at different

levels. Density dependence was detected in all species but was weakest for the insectivorous

dasyurids, suggesting that environmental stochasticity and interactions with other species on

a local scale are more important in driving their population dynamics than intrinsic factors. In

contrast, rodents and the carnivorous mulgara were driven by both extrinsic and intrinsic

factors that operate at the landscape scale, as predicted by Moran’s theorem.

Acknowledgements

We thank Bush Heritage Australia, H. Jukes, G. McDonald, D. Smith and G. Woods for allowing access to the properties in the study region; B. Tamayo, C.-L. Beh and other

members of the Desert Ecology Research Group, and many volunteers for valuable assistance

in the field. Funding was provided by the Australian Research Council, an Australian

Postgraduate Award (to ACG) and the Australian Government’s Terrestrial Ecosystems

Research Network (www.tern.gov.au), an Australian research infrastructure facility

established under the National Collaborative Research Infrastructure Strategy and Education

Infrastructure Fund - Super Science Initiative through the Department of Industry,

Innovation, Science, Research and Tertiary Education.

121

Chapter 4: Mammal population dynamics

References:

Bateman, A. W., A. Ozgul, T. Coulson, and T. H. Clutton-Brock. 2011. Density dependence in group dynamics of a highly social mongoose, Suricata suricatta. Journal of Animal Ecology: 628–639. Bjornstad, O. N. 2013. ncf: spatial nonparametric covariance functions. R package version 1.1-5. Bjørnstad, O. N., R. A. Ims, and X. Lambin. 1999. Spatial population dynamics: analyzing patterns and processes of population synchrony. Trends in Ecology & Evolution 14: 427-432. Brodie, J., H. Johnson, M. Mitchell, P. Zager, K. Proffitt, M. Hebblewhite, M. Kauffman, B. Johnson, J. Bissonette, C. Bishop, J. Gude, J. Herbert, K. Hersey, M. Hurley, P. M. Lukacs, S. McCorquodale, E. McIntire, J. Nowak, H. Sawyer, D. Smith, and P. J. White. 2013. Relative influence of human harvest, carnivores, and weather on adult female elk survival across western North America. Journal of Applied Ecology 50: 295-305. Brook, B. W. and C. J. A. Bradshaw. 2006. Strength of evidence for density depenence in abundance time series of 1198 species. Ecology 87: 1445-1451. Bureau of Meteorology. 2012. Climate data online, accessed 20 June 2012, http://www.bom.gov.au/climate/data/ Burnham, K. P. and D. R. Anderson. 2002. Model selection and multimodel inference : a practical information-theoretic approach. 2nd edition. Springer, New York. Chen, X., C. R. Dickman, and M. B. Thompson. 1998. Diet of the mulgara, Dasycercus cristicauda (Marsupialia: Dasyuridae), in the Simpson Desert, central Australia. Wildlife Research 25: 233-242. Clark, J. S. and O. N. Bjørnstad. 2004. Population time series: process variability, observation errors, missing values, lags and hidden states. Ecology 85: 3140-3150. Dickman, C. R. 2003. Species interactions: indirect effects, Pages 158-170 in P. Attiwill and B. Wilson, editors. Ecology: an Australian perspective. Oxford University Press, Melbourne, Vic. Dickman, C. R., A. C. Greenville, C.-L. Beh, B. Tamayo, and G. M. Wardle. 2010. Social organization and movements of desert rodents during population “booms” and “busts” in central Australia. Journal of Mammalogy 91: 798-810.

122

Chapter 4: Mammal population dynamics

Dickman, C. R., A. C. Greenville, B. Tamayo, and G. M. Wardle. 2011. Spatial dynamics of small mammals in central Australian desert habitats: the role of drought refugia. Journal of Mammalogy 92: 1193-1209. Dickman, C. R., A. S. Haythornthwaite, G. H. McNaught, P. S. Mahon, B. Tamayo, and M. Letnic. 2001. Population dynamics of three species of dasyurid marsupials in arid central Australia: a 10-year study. Wildlife Research 28: 493-506. Dickman, C. R., P. S. Mahon, P. Masters, and D. F. Gibson. 1999. Long-term dynamics of rodent populations in arid Australia: the influence of rainfall. Wildlife Research 26: 389-403. Dickman, C. R., M. Predavec, and F. J. Downey. 1995. Long-range movements of small mammals in arid Australia: implications for land management. Journal of Arid Environments 31: 441-452. Evans, M. E. K., K. E. Holsinger, and E. S. Menges. 2010. Fire, vital rates, and population viability: a hierarchical Bayesian analysis of the endangered Florida scrub mint. Ecological Monographs 80: 627-649. Friend, G. R., G. T. Smith, D. S. Mitchell, and C. R. Dickman. 1989. Influence of pitfall and drift fence design on capture rates of small vertebrates in semi-arid habitats of Western Australia. Australian Wildlife Research 16: 1-10. Greenville, A. C., C. R. Dickman, G. M. Wardle, and M. Letnic. 2009. The fire history of an arid grassland: the influence of antecedent rainfall and ENSO. International Journal of Wildland Fire 18: 631-639. Greenville, A. C., G. M. Wardle, and C. R. Dickman. 2012. Extreme climatic events drive mammal irruptions: regression analysis of 100-year trends in desert rainfall and temperature. Ecology and Evolution 2: 2645-2658. Greenville, A. C., G. M. Wardle, and C. R. Dickman. 2013. Extreme rainfall events predict irruptions of rat plagues in central Australia. Austral Ecology 38: 754-764. Harris, D. B. and D. W. Macdonald. 2007. Interference competition between introduced black rats and endemic Galapagos rice rats. Ecology 88: 2330-2344. Hinrichsen, R. A. and E. E. Holmes. 2009. Using multivariate state-space models to study spatial structure and dynamics, in R. S. Cantrell, C. Cosner, and S. Ruan, editors. Spatial Ecology. CRC/Chapman Hall., London. Holmes, E., E. Ward, and K. Wills. 2013. MARSS: multivariate autoregressive state-space modeling. R package version 3.4.

123

Chapter 4: Mammal population dynamics

Holmes, E. E., E. J. Ward, and M. D. Scheuerell. 2012a. Analysis of multivariate time-series using the MARSS package. NOAA Fisheries, Seattle, USA. Holmes, E. E., E. J. Ward, and K. Wills. 2012b. MARSS: multivariate autoregressive state- space models for analyzing time-series data. R Journal 4: 11-19. Ims, R. A. and H. P. Andreassen. 2000. Spatial synchronization of vole population dynamics by predatory birds. Nature 408: 194-196. Knape, J., P. Besbeas, and P. de Valpine. 2013. Using uncertainty estimates in analyses of population time series. Ecology 94: 2097–2107. Knape, J. and P. de Valpine. 2012. Are patterns of density dependence in the Global Population Dynamics Database driven by uncertainty about population abundance? Ecology Letters 15: 17-23. Kok, O. B. and J. A. J. Nel. 1996. The Kuiseb river as a linear oasis in the desert. African Journal of Ecology 34: 39-47. Körtner, G., C. R. Pavey, and F. Geiser. 2007. Spatial ecology of the mulgara in arid Australia: impact of fire history on home range size and burrow use. Journal of Zoology 273: 350-357. Letnic, M. and C. R. Dickman. 2006. Boom means bust: interactions between the El Niño/Southern Oscillation (ENSO), rainfall and the processes threatening mammal species in arid Australia. Biodiversity and Conservation 15: 3847-3880. Letnic, M., C. R. Dickman, M. K. Tischler, B. Tamayo, and C. L. Beh. 2004. The responses of small mammals and lizards to post-fire succession and rainfall in arid Australia. Journal of Arid Environments 59: 85-114. Masters, P. 1993. The effects of fire-driven succession and rainfall on small mammals in spinifex grassland at Uluru National Park, Northern Territory. Wildlife Research 20: 803-813. Masters, P. 1998. The mulgara Dasycercus cristicauda (Marsupialia: Dasyuridae) at Uluru National Park, Northern Territory. Australian Mammalogy 20: 403-407. Masters, P., C. R. Dickman, and M. Crowther. 2003. Effects of cover reduction on mulgara Dasycercus cristicauda (Marsupialia: Dasyuridae), rodent and invertebrate populations in central Australia: implications for land management. Austral Ecology 28: 658-665. McCarthy, M. A. and P. Masters. 2005. Profiting from prior information in Bayesian analyses of ecological data. Journal of Applied Ecology 42: 1012-1019.

124

Chapter 4: Mammal population dynamics

Moran, P. 1953. The statistical analysis of the Canadian lynx cycle. Australian Journal of Zoology 1: 291-298. Murray, B. R. and C. R. Dickman. 1994. Granivory and microhabitat use in Australian desert rodents: are seeds important? Oecologia 99: 216-225. NAFI. 2013. North Australian Fire Information, accessed 12th December 2012, http://www.firenorth.org.au/nafi2/ Nuno, A., N. Bunnefeld, and E. J. Milner-Gulland. 2013. Matching observations and reality: using simulation models to improve monitoring under uncertainty in the Serengeti. Journal of Applied Ecology 50: 488–498. Predavec, M. 1994. Population dynamics and environmental changes during natural irruptions of Australian desert rodents. Wildlife Research 21: 569-581. Predavec, M. 2000. Food limitation in Australian desert rodents: experiments using supplementary feeding. Oikos 91: 512-522. Previtali, M. A., M. Lima, P. L. Meserve, D. A. Kelt, and J. R. Gutiérrez. 2009. Population dynamics of two sympatric rodents in a variable environment: rainfall, resource availability, and predation. Ecology 90: 1996-2006. Purdie, J. L. 1984. Land systems of the Simpson Desert region. Natural resources series no. 2. CSIRO Division of Water and Land Resources, Melbourne, Australia. R Core Team. 2014. R: a language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria. Ranta, E., V. Kaitala, J. Lindstrom, and H. Linden. 1995. Synchrony in population dynamics. Proceedings of the Royal Society of London. Series B: Biological Sciences 262: 113- 118. Ranta, E., V. Kaitala, and P. Lundberg. 2006. Ecology of populations. Cambridge University Press, Cambridge. Rothery, P., I. Newton, L. Dale, and T. Wesolowski. 1997. Testing for density dependence allowing for weather effects. Oecologia 112: 518-523. Shenbrot, G., B. Krasnov, and S. Burdelov. 2010. Long-term study of population dynamics and habitat selection of rodents in the Negev Desert. Journal of Mammalogy 91: 776- 786. Shephard, M. 1992. The Simpson Desert: natural history and human endeavour. Royal Geographical Society of Australasia, Adelaide, Australia. Ward, E. J., H. Chirakkal, M. González-Suárez, D. Aurioles-Gamboa, E. E. Holmes, and L. Gerber. 2010. Inferring spatial structure from time-series data: using multivariate 125

Chapter 4: Mammal population dynamics

state-space models to detect metapopulation structure of California sea lions in the Gulf of California, Mexico. Journal of Applied Ecology 47: 47-56. Wolda, H. and B. Dennis. 1993. Density dependence tests, are they? Oecologia 95: 581-591. Woodman, J. D., J. E. Ash, and D. M. Rowell. 2006. Population structure in a saproxylic funnelweb spider (Hexathelidae: Hadronyche) along a forested rainfall gradient. Journal of Zoology 268: 325-333.

126

Chapter 5: Reptile population dynamics

Chapter 5: Spatial and temporal synchrony in reptile

population dynamics in variable environments

A central netted dragon (Ctenophorus nuchalis) surveys its environment. Photograph by Aaron Greenville.

127

Chapter 5: Reptile population dynamics

Abstract

Resources are seldom distributed equally across space, but many species nonetheless

exhibit spatially synchronous population dynamics. Such synchrony suggests the operation of

large scale external drivers, known as the Moran effect. However, testing the generality of

this effect is not easy, especially in variable environments, as long-term, spatially replicated datasets are required. Here, we use multivariate state-space models to investigate the spatial population structure of six species of reptiles using 17–22 years of intensive live-trapping data from nine spatially distinct sites in an arid region of central Australia that has marked climatic variability. We then use the best fitting models to incorporate drivers that potentially regulate these populations. Sub-populations of the six species fluctuated synchronously or exhibited two distinct structures depending on whether they were located in burnt versus unburnt sites or open desert versus oasis sites; density dependence was detected in two species but not in the other four. Antecedent rainfall and cover of the dominant vegetation, spinifex grass, were found to further influence populations of three of the study species but had no effect on the other three. The results suggest that regional-scale processes are more important than intrinsic factors in driving the dynamics of these species, and thus show that the Moran effect can prevail even in highly variable environments.

128

Chapter 5: Reptile population dynamics

Introduction

Determining the factors that influence the spatial dynamics of species’ populations

remains a key goal in ecology and is an imperative for managing species that are in decline.

Sub-populations across species’ ranges seldom share the same level of resources, and this

may lead to different densities and growth rates among them (Ranta et al. 2006). Dispersal of individuals between sub-populations can dampen the independence of these units, but this effect decreases with distance (Ranta et al. 1995). However, local populations can still behave synchronously across large (>1000 km2) areas (Ranta et al. 2006), suggesting that large-scale drivers are operating. Drivers may include predation (Ims and Andreassen 2000), the Moran effect, which states that sub-populations with a common density dependent structure can be synchronized by a spatially correlated density independent factor, such as climate (Moran 1953, Bjørnstad et al. 1999), or a combination of the Moran effect and dispersal (Kendall et al. 2000, Ranta et al. 2006).

The Moran effect provides a theoretical basis for maintaining population synchrony across large areas. Synchrony occurs in many taxa, including terrestrial mammals, birds, fish, insects and plants (Moran 1953, Hanski and Woiwod 1993, Myers 1998, Koenig and Knops

2000, Post and Forchhammer 2002, Cattanéo et al. 2003, Cattadori et al. 2005). In arid regions, for example, synchrony may be imposed by regional-scale sequences of droughts and flooding rains that affect all sub-populations similarly; the synchronous ‘boom’ and

‘bust’ dynamics of desert plants and rodents exemplify this pattern (Dickman et al. 1999b,

Greenville et al. 2013, Chapter 3). However, sub-populations may also experience varied conditions in different parts of species’ ranges. For example, sub-populations in different rainfall zones will likely experience different levels of productivity (Ahumada et al. 2004,

Woodman et al. 2006), while sub-populations that experience a disturbance event, such as

129

Chapter 5: Reptile population dynamics

wildfire (Letnic et al. 2004, Evans et al. 2010, Masters and Dickman 2012), may have different trajectories to those that have not been disturbed. Species that exploit patchy resources, such as ephemeral waters in arid regions, may also exhibit different dynamics

(Kok and Nel 1996), with sub-populations acting as spatially independent units (Dickman et

al. 2011). Moran’s effect thus may be weak in highly variable environments, especially in

taxa with low dispersal ability.

Here, we test the Moran effect using Australian desert lizards. Although climatic

variables often show higher levels of spatial correlation over larger distances in Australia

than elsewhere (Koenig 2002), rainfall in arid Australia is highly variable in time and space

(Van Etten 2009). Rainfall influences the population size, breeding and body condition of

many species of lizards, presumably through increases in food resources (Dickman et al.

1999a, Greenville and Dickman 2005, Schlesinger et al. 2011) and provision of foraging and

shelter sites (Daly et al. 2007, 2008). Rainfall-stimulated productivity also generates ground fuel for subsequent wildfires. Fires remove vegetation cover (Greenville et al. 2009), with different species of lizards following vegetation regrowth over time (Letnic et al. 2004,

Pianka and Goodyear 2012, Smith et al. 2013). However, rainfall has little effect on lizard dispersal; animals remain sedentary for long periods (Dickman et al. 1999a). The use of lizards in highly variable desert conditions therefore provides a unique opportunity to test the generality of the Moran effect.

The degree of synchrony across population time series is measured traditionally using correlations, with positive values indicating that populations are in synchrony (Ranta et al.

2006). However, obtaining accurate estimates of species’ population size is often challenging

(Clark and Bjørnstad 2004). Variation in estimates can arise from process noise, uncertainty in population growth rates due to environmental or demographic stochasticity, and observation error (Ward et al. 2010, Knape et al. 2013). In addition, long-term monitoring of

130

Chapter 5: Reptile population dynamics

populations is often difficult due to short-term funding cycles or inconsistent access to study sites, leading to gaps in time series. Long-term monitoring usually also involves multiple observers, potentially inflating observation error over time (Nuno et al. 2013). By partitioning process and observation errors and incorporating these into models, accurate estimates of population parameters can be made (Clark and Bjørnstad 2004, Ward et al.

2010). A framework that includes both types of error and handles missing data is the state- space model (Ward et al. 2010, Holmes et al. 2012b). This framework can also incorporate multiple sub-populations, and thus allow tests of hypotheses about spatial structure and processes that may influence species’ population dynamics (Hinrichsen and Holmes 2009,

Ward et al. 2010).

In this study, we test for the presence of five potential spatial population structures, with and without density dependence, of six species of desert-dwelling lizards. Using multivariate state-space models, we first explore whether sub-populations are asynchronous

(independent hypothesis), form two sub-populations at either ephemeral water sources or open desert sites (oasis hypothesis) or at burnt versus unburnt sites (wildfire hypothesis), form three sub-populations organized by shared rainfall (productivity hypothesis), or whether all sub-populations are synchronous (synchronous population hypothesis). Secondly, we use the best fitting models to identify potential drivers, such as local rainfall and vegetation

cover, that may influence the lizard populations. We predict that species with synchronous spatial dynamics will be driven by factors that operate at the landscape scale, such as resource-pulses from large-scale rainfall events or wildfires. In contrast, species with populations that fluctuate asynchronously will be largely influenced by factors at a local scale.

131

Chapter 5: Reptile population dynamics

Material and Methods

Study region

The study was carried out in the Simpson Desert, central Australia (Fig. A2.1). This

desert occupies 170 000 km2; dune fields comprise 73% of the region, with smaller areas

consisting of clay pans, rocky outcrops and gibber flats (Shephard 1992). The sand dunes run

parallel in a north-south direction aligned with the prevailing southerly wind, are up to 10 m

high and spaced 0.6–1 km apart (Purdie 1984). Vegetation in the interdune swales and on

dune sides is predominantly spinifex (Triodia basedowii) grassland with small stands of

gidgee trees (Acacia georginae), other woody Acacia shrubs and mallee eucalypts; low-lying clay pans fill with water temporarily after heavy rain.

During summer, daily temperatures usually exceed 40º C and minima in winter often

fall below 5º C (Purdie 1984). Highest rainfall occurs in summer, but heavy rains can fall

locally or regionally throughout the year. Long-term weather stations in the study region are

at Glenormiston (1890–2011), Boulia (1888–2011) and Birdsville (1954–2011), and have

median annual rainfalls of 186 mm (n = 122 years), 216.2 mm (n = 124 years), and 153.1 mm

(n = 57 years), respectively (Bureau of Meteorology 2012). Highest rainfall occurs in the north of the region, with a weak gradient from east to west (Bureau of Meteorology 2012).

Reptiles

Live-trapping was carried out at nine sites across Carlo Station, Tobermorey Station,

Cravens Peak and Ethabuka Reserves, an 8000 km2 desert region in south-western

Queensland (Fig. A2.1). The nine sites were spaced > 20 km apart to nullify any influence of

dispersal for the small (< 125 g) study species. Reptiles were live-trapped using pitfall traps

132

Chapter 5: Reptile population dynamics

(16 cm diameter, 60 cm deep), each equipped with a 5 m drift fence of aluminium flywire to

increase trap efficiency (Friend et al. 1989). Pitfalls were arranged on grids comprising six

lines of six traps spaced 20 m apart to cover 1 ha (i.e. 36 pitfall traps per grid). The top line of

traps was positioned on a dune crest and the bottom line 100 m distant in the swale so that

each grid sampled the topography of the dune field. Sites contained 2–12 grids, and grids

within sites were set 0.5–2 km apart in randomly chosen positions.

Traps were opened from 1990–2012 at one site (Main Camp), 2–6 times a year and

from 1995–2012 at eight more sites (Shitty Site, Field River South, Field River North, South

Site, Kunnamuka Swamp East, Carlo, Tobermorey East and Tobermorey West). Each

trapping grid was opened for 2–6 nights. Data from 1990–1994 for one site (Main Camp)

were included in the models because a large (>400 mm) rainfall event occurred in 1991,

which likely influenced the population dynamics of the reptile species (Dickman et al.

1999a). The other eight sites were set up five years later as the project expanded. Recaptures

were removed from live-trapping records here to avoid double counting, but the recapture

rate was too low (2% of total captures) to estimate detection probabilities (see modelling

approach below). To account for unequal trapping effort, live-capture counts were

standardized per 100 trap nights (TN: trap nights = traps × nights opened) and averaged per year. Not all sites were opened every year, but were included in the time-series as missing values. Live-trapping data were log-transformed (log+1) as the population models used below are in log-space.

Long-term (17–22 years) live-trapping (205 524 trap nights) yielded 58 species of reptiles. Of these, only the skinks Ctenotus pantherinus, C. ariadnae, C. dux, Lerista labialis, and the dragons Ctenophorus isolepis and C. nuchalis had sufficient data in terms of length of time-series (>10 years) and captures across sites (>4 sites) to run the population models below (Table A3.1); these species therefore form the basis of our analyses.

133

Chapter 5: Reptile population dynamics

Wildfire

Large-scale wildfires (>1000 km2) have occurred three times in the study region since

1972, but the mean return interval for wildfire at local (1 ha) sites is 26 years (Greenville et al. 2009, NAFI 2013). Wildfires can remove the dominant vegetation and lead to shifts in reptile assemblages (Letnic et al. 2004). Some species, such as C. isolepis, can show dramatic drops in numbers and only recover when vegetation returns, whereas other species, such as C. nuchalis can take advantage of recently burnt sites (Letnic et al. 2004). To investigate whether wildfire affected the population dynamics of the study species, the wildfire history at the site of each sub-population was classified according to whether it had experienced a wildfire (six sites, see Appendix A2.1) or remained unburnt (three sites, see Appendix A2.1) across the 22-year study period.

Rainfall

Daily rainfall data from automatic weather stations (Environdata, Warwick,

Queensland) located at each site were used to calculate annual total rainfall (Fig. 1.2), the number of rain days, mean rainfall per day, mean and maximum event size, and also rainfall lags for the prior year. The weather stations were active at each site from 1995–2012. Annual rainfall across these stations showed a mean cross-correlation (regional synchrony) of 0.77 that was not influenced by the distances separating the stations (Fig. A3.1). Central

Australian lizard species have varied life-histories, even within genera, and thus may respond differently to rainfall events (Dickman et al. 1999a). To avoid a “one size fits all” approach, we selected rainfall metrics for covariates with the highest Pearson correlation with trapping

134

Chapter 5: Reptile population dynamics

rates. Thus, we selected annual rainfall from the prior year for C. ariadnae and C. isolepis (r

= 0.18 and 0.47 respectively), number of rain days from the prior year for C. dux, C. pantherinus, and C. nuchalis (r = 0.17, 0.21 and −0.33, respectively), and annual rainfall for

L. labialis (r = 0.15).

Spinifex cover

To measure cover of the dominant vegetation, spinifex (T. basedowii), we scored ground cover visually as a percentage in 2.5 m radius plots around six randomly selected traps on each trapping grid used to sample reptiles. The cover estimates were pooled for each site, per year over the 17–22 year study period for each of the nine sites to yield a mean spinifex cover and mean spinifex cover in the prior year. As above, we selected the spinifex cover metrics for covariates that showed the highest Pearson correlation with trapping rates.

Thus, we selected spinifex cover from the same year of trapping for C. dux, C. ariadnae, C. pantherinus, and C. nuchalis (r = 0.22, 0.27, 0.42 and −0.54, respectively), and spinifex cover from the prior year for C. isolepis and L. labialis (r = −0.13 and −0.14, respectively).

Multivariate autoregressive state-space models

We used multivariate autoregressive state-space (MARSS) models to analyse the 17–

22 year live-trapping datasets from our nine sites. MARSS models are based on the Gompertz growth model and assume that population growth rate varies exponentially with population size and that populations are closed to immigration and emigration (Hinrichsen and Holmes

2009). The MARSS framework is hierarchical and allows modelling of different spatial population structures and parameters, such as density dependence, while including both

135

Chapter 5: Reptile population dynamics

process (state) and observation variability (Ward et al. 2010). Process variability represents temporal variability in population growth rates due to environmental stochasticity (Ward et al. 2010). Observation variability includes sampling error (e.g. temporal changes in detectability or error resulting in only a sub-sample of the population being counted)

(Hinrichsen and Holmes 2009, Ward et al. 2010). These models can incorporate missing values and estimate them based on model parameters (Holmes et al. 2012b).

The process component is a multivariate first-order autoregressive process and is written in log-space (see Holmes et al. 2012a, Holmes et al. 2012b) as:

= + + ; ~ (0, ) (1)

퐗푡 퐁퐗푡−1 퐮 퐰푡 퐰푡 푀푀푀 퐐 where Xt represents a column vector of all m sub-populations (up to nine sub-populations) at time t (year), and u is a column vector of length m. B and Q are m × m matrixes that denote

process parameters. The B diagonal elements represent the coefficients of autoregression in

sub-populations through time and represent the strength of density dependence (Bii = 1 represents density independence, Bii < 1 = density dependence). The off-diagonal elements in the B matrix allow for interactions between processes, such as between sub-populations or

covariates (Holmes et al. 2012a). The parameter u describes the mean growth rate of the sub-

population. We allowed u to vary, as we assumed that growth rates may differ across sites. wt

denotes process errors which we assumed to be independent and to follow a multivariate

normal distribution with a mean of zero and variance-covariance matrix Q (i.e. Q represents how the sub-populations are correlated to one another). We allowed Q to have different variances across sites (each diagonal element of the Q matrix was allowed to vary), as process variation was assumed to differ across space.

136

Chapter 5: Reptile population dynamics

The observation component, written in log-space (see Holmes et al. 2012a, 2012b), is:

= + + ; ~ (0, ) (2)

퐘푡 퐙퐗푡 퐚 퐯푡 퐯푡 푀푀푀 퐑 where Yt represents a column vector of observations all n survey sites at time t, a denotes the

mean bias between sites, and Z is an n × m matrix of 0s and 1s that assigns observations to a

sub-population structure. Observation errors, vt, are assumed to be uncorrelated and to follow

a multivariate normal distribution with a mean of zero and a variance-covariance matrix R.

We set R to have equal variance across sites, as the same trapping methods were employed throughout the study and thus we did not expect a difference in trappability. Equations (1) and (2) comprise the MARSS model.

The modelling approach was undertaken in two steps to reduce computational difficulties. Firstly, for each species, we tested five possible spatial sub-population structures based on our five hypotheses: independent, oasis, productivity, wildfire and synchronous population (each hypothesis was modelled by assigning a matrix of 1s and 0s via the Z- matrix. See Appendices A2.1 and A2.2 for site selection and model formulation). Each hypothesis was tested with and without density dependence (B in equation 1), and thus ten models were compared. The best-fitting model was chosen using Akaike’s Information

Criterion (AICc), adjusted for small sample size (Burnham and Anderson 2002). Models were

run with the Expectation-Maximization (EM) algorithm, and if they did not converge models

were simplified based on the parameters that failed to converge (e.g. Q was set to be equal

across all sites in the event of convergence failure, or if the initial state estimates, x0, failed to

converge the models were constrained to estimate states from the first time period, x1). If the simplified models still failed to converge, the EM parameters were used as starting points for the quasi-Newton (BFGS) algorithm to find the best estimates (Holmes et al. 2012a).

137

Chapter 5: Reptile population dynamics

Secondly, the best fitting model selected above was expanded to include covariates

that might influence lizard population dynamics. All variables were standardized by

subtracting the mean and dividing by the standard deviation (z-scored) to allow direct comparisons between covariates to be made. Covariates were modelled as separate processes to account for missing values, and their interaction with each species’ population was assessed by examining the off-diagonals of the B matrix (Holmes et al. 2012a). The diagonal

elements of the B matrix were assessed for the presence of density dependence (see above).

In all models that included covariates, we set u = 0 as numeric difficulties can arise when

estimating both u and B within the same model and each dataset has been de-meaned (i.e. via

z-score transformation so u should equal 0) (Holmes et al. 2012a).

We calculated 95% confidence intervals for each parameter using the Hessian

matrix (the Hessian CIs are based on the asymptotic normality of maximum-likelihood

estimation parameters under a large-sample approximation) and, if they did not cross zero,

they were determined to be significant. Inspection of diagnostic plots indicated that all

models met statistical assumptions. All analyses were conducted in R 2.15.3 (R Core Team

2014), using the MARSS 3.4 package (Holmes et al. 2013).

Measures of synchrony

To further assess the degree of regional synchrony and compare results with those from the MARSS models, we calculated zero-lag cross-correlations of the abundance (log) of the nine sub-populations for each species. The mean cross-correlation coefficients were calculated with 1000 resamples using the ncf 1.1-5 package (Bjornstad 2013), in R 3.02 (R

Core Team 2014).

138

Chapter 5: Reptile population dynamics

Results

In total, 17 681 captures of individual reptiles were made over the period of study, including 12 570 of the six study species (Table A3.1), yielding an overall trap success rate of

8.6% in relation to the total trapping effort.

Skinks

The four species of skinks showed varied temporal and spatial dynamics. Both C. pantherinus and C. dux exhibited similar spatial population structure. The best fitting model predicted that sub-populations of these skinks showed density dependence and were structured by wildfire (Table 5.1; Fig. 5.1a and b). Both species had positive population growth rates (u) and similar process and observation errors (Table 5.2), with relatively larger observation errors that indicated similarity in their population dynamics and trappability.

Sub-populations of C. pantherinus were affected positively by spinifex cover and by number of rain days in the prior year; these factors had no effect on sub-populations of C. dux (Table

5.3). The density dependence term (Bii) was stronger with covariates included for C. pantherinus than without (Tables 5.2 and 5.3). The synchronous population hypothesis, with density dependence, was the best fitting model for C. ariadnae (Table 5.1), indicative of spatial synchrony. However, the confidence intervals for the density dependence term (Bii) overlapped zero, suggesting that density dependence is not important in this species (Table

5.2). This species had a positive growth rate and observation error that was larger than the process error, suggesting that error in detecting populations of this skink is greater than that from biotic or abiotic variation that may influence the species’ populations (Table 5.2).

Spinifex cover had no effect on sub-population estimates for C. ariadnae, but rain in the year prior was associated with fewer captures (Table 5.3).

139

Chapter 5: Reptile population dynamics

The wildfire and synchronous population hypotheses, with density dependence, both

provided good fits to the data for L. labialis (Table 5.1; Fig. 5.2a and b). However,

confidence intervals for estimates of the density dependence terms overlapped zero,

suggesting that density dependence is not important in this species (Table 5.2). Growth rates

were strongly positive in both models, with observation errors higher than process errors,

suggesting that the error in detecting populations of this skink is greater than that from biotic

or abiotic variation that influences the species’ populations (Table 5.2). No covariates were

significant for L. labialis (Table 5.3). Mean cross-correlation coefficients were 0.13 (CI:

−0.02–0.33), 0.18 (CI: 0.03–0.36), 0.16 (CI: −0.05–0.39) and 0.19 (CI: −0.002–0.41), for C.

pantherinus, C. ariadnae, C. dux and L. labialis, respectively, indicating little influence of distance on abundance (Fig. A3.2).

140

Chapter 5: Reptile population dynamics

Table 5.1: AICc values for MARSS models describing five possible spatial sub-population structures, with and without density dependence, for six species of reptiles in the Simpson Desert, central Australia. Data are based on 17–22 years of live-trapping from nine sites. Smallest AICc, highlighted in bold, indicates the best-fitting model.

Model Without density dependence With density dependence Ctenotus pantherinus Independent 135.03 147.94 Oasis 140.33 140.44 Productivity 130.51 128.57 Wildfire 129.02 119.18 Synchronous population 129.53 128.44

Ctenotus ariadnae Independent 178.79 198.60 Oasis 181.11 183.84 Productivity 181.51 183.37 Wildfire 183.31 183.83 Synchronous population 182.06 175.00

Ctenotus dux Independent 130.11 136.31 Oasis 140.21 140.50 Productivity 134.79 130.84 Wildfire 136.79 128.93 Synchronous population 137.07 136.32

Lerista labialis Independent 325.68 297.2 Oasis 296.67 292.06 Productivity 297.87 305.06 Wildfire 295.26 282.61 Synchronous population 288.27 282.21

Ctenophorus isolepis Independent 220.32 226.41 Oasis 206.27 192.1 Productivity 218.57 208.85 Wildfire 213.5 200.24 Synchronous population 202.44 194.66

Ctenophorus nuchalis Independent 261.94 239.3 Oasis 172.27 167.98 Productivity 204.45 197.87 Wildfire 161.7 156.34 Synchronous population 145.13 143.86

141

Chapter 5: Reptile population dynamics

Table 5.2: Model estimates (95% confidence intervals) for best fitting MARSS population models (lowest AICc) for captures of reptiles (100 trap nights, log +1 transformed), Simpson Desert, central Australia. Density dependence occurs if Bii < 1.

Model Growth rate (u) Density Process error (Qii) Observation error dependence (Bii) (Rii) Ctenotus pantherinus Wildfire: burnt 0.35 (0.17–0.53) 0.62* (0.44–0.80) 0.03 (0.001–0.05) 0.09* (0.07–0.11) unburnt 0.35 (0.19–0.51) 0.004 (−0.01–0.02)

Ctenotus ariadnae Synchronous 0.78 (0.39–1.18) 0.03 (−0.41–0.46) 0.03 (0.00001–0.06) 0.14 (0.1–0.17) population

Ctenotus dux Wildfire: burnt 0.39 (0.22–0.57) 0.53* (0.34–0.72) 0.01* (−0.0009–0.03) 0.1* (0.08–0.13) unburnt 0.46 (0.27–0.66)

Lerista labialis Synchronous 1.17 (0.33–2.02) 0.28 (−0.28–0.77) 0.07 (0.005–0.14) 0.28 (0.21–0.35) population Wildfire: burnt 1.16 (0.5–1.8) 0.25* (−0.16–0.66) 0.08 (−0.002–0.15) 0.24* (0.18–0.31) unburnt 1.14 (0.44–1.83) 0.19 (0.003–0.38)

Ctenophorus isolepis Oasis: open desert 0.8 (0.38–1.22) 0.2* (−0.2–0.61) 0.05 (0.002–0.1) 0.13* (0.09–0.16) oasis 0.73 (0.31–1.15) 0.1 (0.01–0.19)

Ctenophorus nuchalis Synchronous 0.2 (−0.07–0.48) 0.2 (−0.07–0.48) 0.08 (0.02–0.14) 0.1 (0.07–0.12) population *One parameter estimated for entire model.

142

Chapter 5: Reptile population dynamics

Table 5.3: The results of the covariate MARSS models for captures of reptiles (100 trap nights, log +1 transformed), Simpson Desert, central Australia. Covariates were entered into models as their own process, and the coefficients from the off-diagonals of the B matrix measure the strength of their interaction with reptile population growth rates. The diagonals of the B matrix indicate the strength of density dependence, which occurs when Bii < 1. Covariates were z-scored so that direct comparisons can be made. Each covariate was considered significant if the 95% confidence intervals (CI) did not cross zero (in bold).

Model covariates Estimate Lower CI Upper CI Ctenotus pantherinus Density dependence (Bii) 0.0078† 0.0078 0.0078 Spinifex cover 0.32 0.04 0.6 Rain days year prior 0.48 0.11 0.85

Ctenotus ariadnae Density dependence (Bii) 0.16 −0.49 0.81 Spinifex cover 0.05 −0.24 0.33 Annual rain year prior −0.36 −0.68 −0.04

Ctenotus dux Density dependence (Bii) 0.41 0.23 0.58 Spinifex cover 0.18 −0.07 0.43 Rain days year prior −0.23 −0.54 0.07

Lerista labialis Single population: Density dependence (Bii) 0.16 −0.4 0.72 Spinifex cover year prior −0.05 −0.39 0.29 Annual rainfall −0.14 −0.4 0.11 Wildfire: Density dependence (Bii) 0.23* Spinifex cover year prior −0.01 −0.35 0.33 Annual rainfall −0.18 −0.58 0.22

Ctenophorus isolepis Density dependence (Bii) 0.32 −0.052 0.69 Spinifex cover year prior −0.068 −0.42 0.29 Annual rainfall year prior −0.15 −0.47 0.17

Ctenophorus nuchalis Density dependence (Bii) 0.15 −0.37 0.68 Spinifex cover −0.29 −0.76 0.18 Rain days year prior −0.85 −1.42 −0.28 *Confidence intervals could not be calculated. † Estimates and CI calculated to a maximum of four decimal places (SE = 4.24 ×10−9).

143

Chapter 5: Reptile population dynamics

Figure 5.1: Predicted population size (line) and captures (dots; log+1 captures/100 trap nights) estimated from MARSS models for (a) Ctenotus pantherinus (red line: burnt, blue line unburnt), (b) Ctenotus dux (red line burnt, blue line: unburnt) and (c) Ctenotus ariadnae, Simpson Desert, central Australia. Time-series data were collected from nine sites (sub- populations) monitored 2–6 times per year for 17–22 years (1990–2012 for one site and 1995–2012 for eight sites; pooled per year). Shaded areas indicate 95% confidence intervals.

144

Chapter 5: Reptile population dynamics

Figure 5.2: Predicted population size (line) and captures (dots; log+1 captures/100 trap nights) estimated from MARSS models for (a) Lerista labialis synchronous population and (b) L. labialis populations that had experienced a wildfire (red line: burnt, blue line unburnt), Simpson Desert, central Australia. Time-series data were collected from nine sites (sub- populations) monitored 2–6 times per year for 17–22 years (1990–2012 for one site and 1995–2012 for eight sites; pooled per year). Shaded areas indicate 95% confidence intervals.

Dragons

The two species of dragons showed very different temporal and spatial dynamics. The synchronous population hypothesis, describing synchrony across all nine sites, was the best

145

Chapter 5: Reptile population dynamics

fitting model for C. nuchalis and the oasis hypothesis was the best fit for C. isolepis. Both models indicated density dependence (Table 5.1; Fig. 5.3a and b). However, confidence

Figure 5.3: Predicted population size (line) and captures (dots; log+1 captures/100 trap nights) estimated from MARSS models for (a) Ctenophorus nuchalis and (b) Ctenophorus isolepis populations at either oases (blue line) or open desert sites (black line), Simpson Desert, central Australia. Time-series data were collected from nine sites (sub-populations) monitored 2–6 times per year for 17–22 years (1990–2012 for one site and 1995–2012 for eight sites; pooled per year). Shaded areas indicate 95% confidence intervals.

intervals for estimates of the density dependence terms overlapped zero, suggesting that density dependence is not important in these species (Table 5.2). Growth rates for both species were positive; however, for C. nuchalis the confidence intervals for growth rate overlapped with zero. Process and observation errors were similar for both species, with 146

Chapter 5: Reptile population dynamics

observation errors being higher (Table 5.2). Neither spinifex cover nor rainfall in the prior year had any effect on C. isolepis, although sub-populations were high in oasis sites during a dry period from 2004–2007 (Table 5.3, Fig. 5.3b, Fig. 1.2). Spinifex cover had no effect on sub-populations of C. nuchalis, but number of rain days in the prior year had a negative effect on capture rate (Table 5.3). Mean cross-correlation coefficients for C. isolepis and C. nuchalis were 0.31 (CI: 0.16–0.46) and 0.72 (CI: 0.64–0.81), respectively, and were not influenced by distance (Fig. A3.3).

Discussion

Our results support the synchronous population hypothesis for C. ariadnae and C. nuchalis, the oasis hypothesis for C. isolepis, and the wildfire hypothesis for C. pantherinus and C. dux. For one species, L. labialis, the synchronous population hypothesis and wildfire hypothesis were equally supported. Support for the synchronous population hypothesis in both skinks and dragons indicates that spatial synchrony among the nine sub-populations was not restricted to just one family of lizards. In addition, half of the six species we investigated showed synchronous population dynamics, suggesting that such synchrony is relatively common. The wildfire hypothesis was equally important for structuring reptile populations, followed by the oasis hypothesis for one species. In contrast, mean cross-correlations for all nine sub-populations were low for all species except C. nuchalis, and distance had no influence, suggesting that population size was different at each time period between each sub- population and that there was little influence of dispersal between sites. Taking these results together, populations of our six study species are generally structured in space into one or two independent populations, of differing sizes, suggesting that landscape-scale population drivers are particularly important. Thus, there is support for the Moran effect in our highly

147

Chapter 5: Reptile population dynamics

variable desert environment, even though population drivers were not identified for some

species.

Antecedent rainfall was associated positively with population estimates for C.

pantherinus and negatively with those for C. ariadnae and C. nuchalis, but had little evident

influence on the other species that we investigated. Previous studies have shown relationships

between rainfall and reptile captures, but results have been mixed. For example, Letnic et al.

(2004) found a negative correlation between rainfall and numbers of C. pantherinus and no

relationship between rainfall and C. nuchalis. However, Dickman et al. (1999a) reported a

positive relationship between autumn rainfall and captures of C. nuchalis. Arthropod

abundance often increases after rainfall (Letnic et al. 2004) so that positive correlations

between lizards and rainfall could be driven by increases in their invertebrate prey. However,

lizards often also have flexible diets and shift their prey preferences between wet and dry

years (James 1991), which may dampen the effects of rainfall. Working in the same study

region, Pastro et al. (2013) also found little effect of rainfall on lizard assemblages, but

suggested that both food availability and predation might be important.

Spinifex cover was associated positively with numbers of C. pantherinus, but no

relationships were found for other species. This is consistent with past studies on skinks

(Letnic et al. 2004, Greenville and Dickman 2009) but not with those on the two species of

dragons; these often show inverse associations with vegetation (Dickman et al. 1999a). For example, C. nuchalis is often active in open areas and becomes scarce after rainfall-induced increases in spinifex cover (Dickman et al. 1999a). The negative association between C. nuchalis and rainfall in our study suggests that the population changes we observed were driven by rainfall-induced changes in factors other than vegetation cover, such as increased activity of mammalian or avian predators (Daly et al. 2008, Masters and Dickman 2012).

Wildfire also was important in structuring the spatial dynamics of C. pantherinus, and a

148

Chapter 5: Reptile population dynamics

possible mechanism for this is via reduction in spinifex cover. However, populations of C.

dux and L. labialis were also structured by wildfire, but showed no relationship with cover.

While these observations accord with past studies in showing that these species increase in

abundance with age since fire (Letnic et al. 2004), they suggest that vegetation cover may not drive the changes. For example, L. labialis is a fossorial skink that eats termites, and is not

associated with spinifex cover (Greenville and Dickman 2005, 2009). Wildfire may change

termite abundance (Letnic et al. 2004), but it also removes vegetation and in turn allows

movement of loose sand. Loose sand allows easier movement for L. labialis and greater

access to foraging areas (Greenville and Dickman 2009).

Density dependence was evident in only two species, C. pantherinus and C. dux. For

both species the density dependence estimate was strong, suggesting that intrinsic factors are

important in regulating populations of these skinks. Both species are relatively large (often

>5.0 g) and were the most commonly captured members of their genus (Table 5.1), traits that

perhaps are conducive to density dependent regulation. There was, by contrast, no evidence

of density dependence in our four other study species. Density dependence is not as strong in

reptile populations as in many other taxa (Brook and Bradshaw 2006) and, in arid regions

where many reptile species persist at very low numbers (Pianka 2013), extrinsic factors may

usually be more important in population regulation.

The spatial population dynamics of lizards in our study system were not consistent

within or across families, indicating that desert-dwelling reptiles employ varied life-history

strategies. Sub-populations fluctuated synchronously or exhibited two structures

(synchronous population, oasis- or wildfire-dependent), and thus are likely to have been

influenced by external environmental processes such as rainfall and wildfire rather than by

intrinsic drivers. These observations suggest that—provided driving processes occur at

regional scales—the Moran effect can operate even in highly variable environments. Despite

149

Chapter 5: Reptile population dynamics

evidence for the Moran effect here, however, we identified some of the extrinsic processes

driving populations of only three of our study species. The processes shaping the spatial

dynamics of the other species remain elusive and serve as foci for future work.

Acknowledgements

We thank Bush Heritage Australia, H. Jukes, G. McDonald, D. Smith and G. Woods

for allowing access to the properties in the study region, members of the Desert Ecology

Research Group, especially B. Tamayo and C.-L. Beh, and many volunteers for valuable assistance in the field. Funding was provided by the Australian Research Council and the

Australian Government’s Terrestrial Ecosystems Research Network (www.tern.gov.au), an

Australian research infrastructure facility established under the National Collaborative

Research Infrastructure Strategy and Education Infrastructure Fund - Super Science Initiative

through the Department of Industry, Innovation, Science, Research and Tertiary Education.

ACG was supported by an Australian Postgraduate Award, and CRD by an Australian

Research Council Fellowship.

150

Chapter 5: Reptile population dynamics

References:

Ahumada, J. A., D. Lapointe, and M. D. Samuel. 2004. Modeling the population dynamics of Culex quinquefasciatus (Diptera: Culicidae), along an elevational gradient in Hawaii. Journal of Medical Entomology 41: 1157-1170. Bjornstad, O. N. 2013. ncf: spatial nonparametric covariance functions. R package version 1.1-5. Bjørnstad, O. N., R. A. Ims, and X. Lambin. 1999. Spatial population dynamics: analyzing patterns and processes of population synchrony. Trends in Ecology & Evolution 14: 427-432. Brook, B. W. and C. J. A. Bradshaw. 2006. Strength of evidence for density depenence in abundance time series of 1198 species. Ecology 87: 1445-1451. Bureau of Meteorology. 2012. Climate data online, accessed 20 June 2012, http://www.bom.gov.au/climate/data/ Burnham, K. P. and D. R. Anderson. 2002. Model selection and multimodel inference : a practical information-theoretic approach. 2nd edition. Springer, New York. Cattadori, I. M., D. T. Haydon, and P. J. Hudson. 2005. Parasites and climate synchronize red grouse populations. Nature 433: 737-741. Cattanéo, F., B. Hugueny, and N. Lamouroux. 2003. Synchrony in brown trout, Salmo trutta, population dynamics: a ‘Moran effect’ on early-life stages. Oikos 100: 43-54. Clark, J. S. and O. N. Bjørnstad. 2004. Population time series: process variability, observation errors, missing values, lags and hidden states. Ecology 85: 3140-3150. Daly, B. G., C. R. Dickman, and M. S. Crowther. 2007. Selection of habitat components by two species of agamid lizards in sandridge desert, central Australia. Austral Ecology 32: 825-833. Daly, B. G., C. R. Dickman, and M. S. Crowther. 2008. Causes of habitat divergence in two species of agamid lizards in arid central Australia. Ecology 89: 65-76. Dickman, C. R., A. C. Greenville, B. Tamayo, and G. M. Wardle. 2011. Spatial dynamics of small mammals in central Australian desert habitats: the role of drought refugia. Journal of Mammalogy 92: 1193-1209. Dickman, C. R., M. Letnic, and P. S. Mahon. 1999a. Population dynamics of two species of dragon lizards in arid Australia: the effects of rainfall. Oecologia 119: 357-366.

151

Chapter 5: Reptile population dynamics

Dickman, C. R., P. S. Mahon, P. Masters, and D. F. Gibson. 1999b. Long-term dynamics of rodent populations in arid Australia: the influence of rainfall. Wildlife Research 26: 389-403. Evans, M. E. K., K. E. Holsinger, and E. S. Menges. 2010. Fire, vital rates, and population viability: a hierarchical Bayesian analysis of the endangered Florida scrub mint. Ecological Monographs 80: 627-649. Friend, G. R., G. T. Smith, D. S. Mitchell, and C. R. Dickman. 1989. Influence of pitfall and drift fence design on capture rates of small vertebrates in semi-arid habitats of Western Australia. Australian Wildlife Research 16: 1-10. Greenville, A. C. and C. R. Dickman. 2005. The ecology of Lerista labialis (Scincidae) in the Simpson Desert: reproduction and diet. Journal of Arid Environments 60: 611-625. Greenville, A. C. and C. R. Dickman. 2009. Factors affecting habitat selection in a specialist fossorial skink. Biological Journal of the Linnean Society 97: 531-544. Greenville, A. C., C. R. Dickman, G. M. Wardle, and M. Letnic. 2009. The fire history of an arid grassland: the influence of antecedent rainfall and ENSO. International Journal of Wildland Fire 18: 631-639. Greenville, A. C., G. M. Wardle, and C. R. Dickman. 2013. Extreme rainfall events predict irruptions of rat plagues in central Australia. Austral Ecology 38: 754-764. Hanski, I. and I. P. Woiwod. 1993. Spatial synchrony in the dynamics of moth and aphid populations. Journal of Animal Ecology 62: 656-668. Hinrichsen, R. A. and E. E. Holmes. 2009. Using multivariate state-space models to study spatial structure and dynamics, in R. S. Cantrell, C. Cosner, and S. Ruan, editors. Spatial Ecology. CRC/Chapman Hall., London. Holmes, E., E. Ward, and K. Wills. 2013. MARSS: multivariate autoregressive state-space modeling. R package version 3.4. Holmes, E. E., E. J. Ward, and M. D. Scheuerell. 2012a. Analysis of multivariate time-series using the MARSS package. NOAA Fisheries, Seattle, USA. Holmes, E. E., E. J. Ward, and K. Wills. 2012b. MARSS: multivariate autoregressive state- space models for analyzing time-series data. R Journal 4: 11-19. Ims, R. A. and H. P. Andreassen. 2000. Spatial synchronization of vole population dynamics by predatory birds. Nature 408: 194-196. James, C. D. 1991. Temporal variation in diets and trophic partitioning by coexisting lizards (Ctenotus) in central Australia. Oecologia 85: 553−561.

152

Chapter 5: Reptile population dynamics

Kendall, B. E., O. N. Bjørnstad, J. Bascompte, T. H. Keitt, and W. F. Fagan. 2000. Dispersal,environmental correlation, and spatial synchrony in population dynamics. The American Naturalist 155: 628-636. Knape, J., P. Besbeas, and P. de Valpine. 2013. Using uncertainty estimates in analyses of population time series. Ecology 94: 2097–2107. Koenig, W. D. 2002. Global patterns of environmental synchrony and the Moran effect. Ecography 25: 283-288. Koenig, W. D. and J. M. H. Knops. 2000. Patterns of annual seed production by northern hemisphere trees: a global perspective. The American Naturalist 155: 59-69. Kok, O. B. and J. A. J. Nel. 1996. The Kuiseb river as a linear oasis in the Namib desert. African Journal of Ecology 34: 39-47. Letnic, M., C. R. Dickman, M. K. Tischler, B. Tamayo, and C. L. Beh. 2004. The responses of small mammals and lizards to post-fire succession and rainfall in arid Australia. Journal of Arid Environments 59: 85-114. Masters, P. and C. R. Dickman. 2012. Population dynamics of Dasycercus blythi (Marsupialia: Dasyuridae) in central Australia: how does the mulgara persist? Wildlife Research 39: 419-428. Moran, P. 1953. The statistical analysis of the Canadian lynx cycle. Australian Journal of Zoology 1: 291-298. Myers, J. H. 1998. Synchrony in outbreaks of forest Lepidoptera: a possible example of the Moran effect. Ecology 79: 1111-1117. NAFI. 2013. North Australian Fire Information, accessed 12th December 2012, http://www.firenorth.org.au/nafi2/ Nuno, A., N. Bunnefeld, and E. J. Milner-Gulland. 2013. Matching observations and reality: using simulation models to improve monitoring under uncertainty in the Serengeti. Journal of Applied Ecology 50: 488–498. Pastro, L., C. Dickman, and M. Letnic. 2013. Effects of wildfire, rainfall and region on desert lizard assemblages: the importance of multi-scale processes. Oecologia 173: 603-614. Pianka, E. R. 2013. Rarity in Australian desert lizards. Austral Ecology 39: 214–224. Pianka, E. R. and S. E. Goodyear. 2012. Lizard responses to wildfire in arid interior Australia: long-term experimental data and commonalities with other studies. Austral Ecology 37: 1-11. Post, E. and M. C. Forchhammer. 2002. Synchronization of animal population dynamics by large-scale climate. Nature 420: 168-171. 153

Chapter 5: Reptile population dynamics

Purdie, J. L. 1984. Land systems of the Simpson Desert region. Natural resources series no. 2. CSIRO Division of Water and Land Resources, Melbourne, Australia. R Core Team. 2014. R: a language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria. Ranta, E., V. Kaitala, J. Lindstrom, and H. Linden. 1995. Synchrony in population dynamics. Proceedings of the Royal Society of London. Series B: Biological Sciences 262: 113- 118. Ranta, E., V. Kaitala, and P. Lundberg. 2006. Ecology of populations. Cambridge University Press, Cambridge. Schlesinger, C. A., K. A. Christian, C. D. James, and S. R. Morton. 2011. Seven lizard species and a blind snake: activity, body condition and growth of desert herpetofauna in relation to rainfall. Australian Journal of Zoology 58: 273-283. Shephard, M. 1992. The Simpson Desert: natural history and human endeavour. Royal Geographical Society of Australasia, Adelaide, Australia. Smith, A. L., C. Michael Bull, and D. A. Driscoll. 2013. Successional specialization in a reptile community cautions against widespread planned burning and complete fire suppression. Journal of Applied Ecology 50: 1178–1186. Van Etten, E. J. B. 2009. Inter-annual rainfall variability of arid Australia: greater than elsewhere? Australian Geographer 40: 109-120. Ward, E. J., H. Chirakkal, M. González-Suárez, D. Aurioles-Gamboa, E. E. Holmes, and L. Gerber. 2010. Inferring spatial structure from time-series data: using multivariate state-space models to detect metapopulation structure of California sea lions in the Gulf of California, Mexico. Journal of Applied Ecology 47: 47-56. Woodman, J. D., J. E. Ash, and D. M. Rowell. 2006. Population structure in a saproxylic funnelweb spider (Hexathelidae: Hadronyche) along a forested rainfall gradient. Journal of Zoology 268: 325-333.

154

Chapter 6: Intraguild predator interactions

Chapter 6: Bottom up and top down processes interact to

modify intraguild interactions in resource-pulse

environments

Top dog: the dingo (Canis dingo) is central Australia’s apex predator. Photograph by Bobby Tamayo.

A version of this chapter is published as Greenville A.C., Wardle G.M., Tamayo B., Dickman C.R. 2014. Bottom-up and top-down processes interact to modify intraguild interactions in resource-pulse environments. Oecologia 175: 1349–1358. My contribution was very substantial, and included conceptualisation of ideas, engagement in fieldwork, completion of data manipulations and analyses, and writing and editing all drafts, in consultation with my co-authors. This chapter was also a popular science article for Australasian Science, helped inform a policy statement for the Ecological Society of Australia Hot Topics and opinion piece in The Conversation (Appendix 6).

155

Chapter 6: Intraguild predator interactions

Abstract

Top-predators are declining globally, in turn allowing populations of smaller

predators, or mesopredators, to increase and potentially have negative effects on biodiversity.

However, detection of interactions among sympatric predators can be complicated by fluctuations in the background availability of resources in the environment, which may modify both the numbers of predators and the strengths of their interactions. Here, we first present a conceptual framework that predicts how top-down and bottom-up interactions may regulate sympatric predator populations in environments that experience resource pulses. We then test it using two years of remote camera trapping data to uncover spatial and temporal interactions between a top-predator, the dingo Canis dingo, and the mesopredatory European red fox Vulpes vulpes and feral cat Felis catus, during population booms, declines and busts in numbers of their prey in a model desert system. We found that dingoes predictably suppress abundances of the mesopredators and that the effects are strongest during declines and busts in prey numbers. Given that resource pulses are usually driven by large yet infrequent rains, we conclude that top-predators like the dingo provide net benefits to prey populations by suppressing mesopredators during prolonged bust periods when prey populations are low and potentially vulnerable.

156

Chapter 6: Intraguild predator interactions

Introduction

Dominant, or top-predators, are important in structuring many ecological communities

and maintaining biodiversity in lower trophic levels (Johnson et al. 2007). One extensively studied process that drives this pattern is the depredation of herbivore populations; this in turn reduces impacts on primary producers (Ripple and Beschta 2012). Another, more recently studied process, occurs when top-predators regulate populations of smaller predators, or mesopredators, and in turn reduce their impacts on small prey (Ritchie et al. 2012). As top- predator populations have declined around the world from habitat destruction and persecution from humans, populations of mesopredators have increased, often with negative effects on diversity (Estes et al. 2011, Ripple et al. 2014). The mesopredator release hypothesis predicts that, if present, dominant predators will suppress their subordinate counterparts but, if removed, the mesopredators will be “released” from competition or direct interference; increases in their numbers then may lead to increased impacts on small prey (Prugh et al.

2009).

Despite this simple theoretical framework, confirmation of the mechanisms that underpin top- and mesopredator relationships is often difficult to obtain. This is because predator densities are often low and interactions are consequently difficult to observe, and because interactions between top- and subordinate predators may not be constant. In particular, bottom-up processes such as the productivity of prey populations often vary over time, and will potentially influence the timing and strength of predator relationships (Elmhagen and

Rushton 2007). Understanding how bottom-up processes affect interactions among predators is a central question in ecology, as limited empirical evidence is sometimes used to argue both for and against reintroductions of top-predators or even to avoid preserving them

157

Chapter 6: Intraguild predator interactions

(Fleming et al. 2012, Johnson and Ritchie 2013). The effects of fluctuating productivity may be most profound and pervasive in environments that experience large yet short-lived resource pulses, as these pulses can allow prey populations temporarily to escape predator regulation and also decouple interactions between the predators (Letnic and Dickman 2010).

Resource-pulse environments are globally widespread (Yang et al. 2010), but are most conspicuous in deserts and rangelands. These environments also are frequently arenas for conflict between shepherds and their livestock, and top-predators.

Environments that experience resource pulses from rapid increases in primary productivity, such as from flood rains, show dramatic ‘booms’ and ‘busts’ in populations of primary consumer organisms (Caughley 1987, Dickman et al. 1999, 2010, Greenville et al.

2013, Chapter 3). Even though resource pulses may be infrequent, many consumers efficiently exploit them; for example, populations of some mammals increase 10–60 fold within six months via elevated reproduction and immigration (Dickman et al. 1999, 2010).

This produces different phases in their population cycles, with long busts, or periods of low numbers, and short-lived booms followed by rapid declines (Fig. 6.1). The decline phase is of particular interest for prey populations as it can be associated with high per capita predation because of a lag in the numerical response of the predators (Spencer et al. 2014), which often peak during the decline phase (e.g. Fig. A4.1). By contrast, the increase phase of small mammals occurs quickly and inevitably when predator numbers are low; per capita predation then is likely to be low and to have little or no effect on prey populations. Arid environments, covering 40% of the global land area (Ward 2009), often exhibit such dynamics in consumer populations due to rainfall unreliability (Van Etten 2009, Dickman et al. 2010).

158

Chapter 6: Intraguild predator interactions

Here, we firstly outline a conceptual framework that predicts how predator interactions vary in resource-pulse environments, and then present an empirical test of the predictions using predators and prey in central Australia as a model system. For simplicity, we depict relationships between predators as being linear, but note that true relationships could take other forms. We assume that dominant predators exert their effects via interference competition or intraguild killing (Moseby et al. 2012), or by driving partitioning of space and time, and use changes in prey populations as a measure of bottom-up processes.

Resource pulse

Bust Boom

Abundance

Decline Species

Time

Figure 6.1: Conceptual model depicting the population dynamics of prey in resource- pulse environments. ‘Booms’, or population irruptions, arise from large resource pulses that are stimulated by events such as flooding rainfall. Declines represent subsequent, sometimes brief, periods of negative increase, while ‘busts’ are often long periods when resources and prey numbers are relatively low.

159

Chapter 6: Intraguild predator interactions

Conceptual models

Dominant–subordinate predator interactions can be influenced by top-down or bottom-

up processes, or both processes together, in resource-pulse environments. If top-down processes are most influential, then the top-predator will have a negative effect on subordinate predators, regardless of the prey population phase (boom, bust or decline) (Fig.

6.2a). For example, wolves (Canis lupus) generally suppress coyote (C. latrans) numbers across different habitat types due to interference competition (Fedriani et al. 2000, Merkle et al. 2009). A scenario based on only bottom-up processes will result in both top- and subordinate predators increasing (or decreasing) together, as the prey base ‘booms’ or declines (Fig. 6.2b). Finally, predator populations may be structured by both top-down and bottom-up processes (Elmhagen and Rushton 2007). As resource pulses drive an irruption in the prey base (Fig. 6.1), populations of both dominant and subordinate predators will increase; however, in the bust and decline phases of prey populations when prey is scarce, the dominant predator will likely have negative effects on subordinate predators owing to increased levels of competition or direct intraguild killing (Fig. 6.2c).

Our conceptual models aim to capture the major effects of dominant on subordinate predators in resource-pulse environments, but do not preclude the possibility that other interactions might occur. For example, interactions are likely between predators and other species in most ecological systems, and interaction strengths may vary depending on which species co-occur (Billick and Case 1994). In addition, the influence of bottom-up effects may change over time, with decreases in prey availability arising due to behavioural changes (e.g. increased predator avoidance), habitat changes (e.g. increases in cover due to plant growth after a resource pulse), or changes in resources other than prey (Power 1992). However, while such effects may modify the slopes or shapes of our conceptual models, we assume that

160

Chapter 6: Intraguild predator interactions their form is robust owing to the strong interference relationships that characterize most interactions between predator species (Palomares and Caro 1999).

(a) Top-down only

Boom

Decline Bust Subordinate predator predator Subordinate

(b) Bottom-up only Boom

Decline predator predator Bust Subordinate Subordinate

(c) Interaction between bottom-up and top-down Boom

Decline

Subordinate predator predator Subordinate Bust Dominant predator Figure 6.2: Conceptual models of dominant and subordinate predator interactions in resource-pulse environments. Three processes are possible; a) top-down only, b) bottom- up only and c) both top-down and bottom-up. Linear relationships are shown for simplicity, but non-linear relationships may be possible. Scale represents abundance.

161

Chapter 6: Intraguild predator interactions

Case study

Central Australia has experienced widespread extinctions of its mammal fauna (Smith

and Quin 1996) and now the most abundant mammalian predators are the dingo (Canis dingo

(Crowther et al. 2014), body mass 15–25 kg), invasive red fox (Vulpes vulpes, 3.5–7.5 kg) and feral cat (Felis catus, 3.8–4.4 kg). The dingo is the continent’s mammalian top-predator

and thought to be important in structuring populations of both mesopredators; strong

suppressive effects have been documented on populations of the fox, and weaker effects on

the feral cat (Letnic and Koch 2010, Letnic et al. 2011, Brook et al. 2012, Letnic et al. 2012).

All three predators have high dietary overlap in central Australia, with rodents often

representing 70–95% by occurrence in their diets, and there is also some evidence of direct

killing by dingoes of red foxes and feral cats (Mahon 1999, Pavey et al. 2008, Cupples et al.

2011, Moseby et al. 2012). Interactions between the red fox and feral cat are largely

unknown. However, high dietary overlap and evidence of predation on feral cats by the red

fox suggest that interference competition and perhaps intraguild predation may take place

(Molsher 1999, Paltridge 2002).

Flooding rains in 2010 stimulated a resource pulse across central Australia and resulted in

dramatic irruptions of small rodents in 2011 (Greenville et al. 2012, 2013, Chapters 2 to 4).

Predator populations also increased in the wake of this resource pulse; their populations

usually respond within 12 months of flood rains (Letnic and Dickman 2006, Pavey et al.

2008). Here, we track the activity of dingoes, foxes and cats from the pre-irruption bust

phase, through the rodent prey pulse, the subsequent decline and into the next bust period to

test our conceptual models (Fig. 6.2a,b,c). Because of the cryptic nature of the predators, we

use two years of data obtained from remote cameras to track temporal changes in their

populations (Vine et al. 2009). If top-down forces predominate, we expect that dingoes will

162

Chapter 6: Intraguild predator interactions

exert suppressive effects at all times (Fig. 6.2a). If bottom-up forces predominate, predators

should respond only to the abundance of their prey (Fig. 6.2b). If top-down and bottom-up

processes are both important in shaping the activity or demographic performance of the

sympatric mesopredators (Fig. 6.2c), we predict that:

1. Top-down forcing of dingoes on mesopredators (red fox and feral cat) will be

manifest only in the bust and decline phases of the predators’ shared prey, and

2. Populations of the mesopredators will increase from bottom-up processes, with the

red fox increasing more than the feral cat if it is the dominant of the two

mesopredators.

In addition to these population-level interactions, we predict that the red fox and feral cat

will have different daily activity times to the dominant dingo to avoid potential encounters,

and that the daily activity of the dingo will coincide most closely with that of its rodent prey.

Materials and methods

Study region

The study region occupies 8000 km2 in the north-eastern Simpson Desert in central

Australia, and covers three properties: Ethabuka Reserve, Carlo Station and Cravens Peak

Reserve. Dune-fields comprise 73% of the region, with smaller areas consisting of clay pans, rocky outcrops and gibber flats. Sand dunes run parallel in a north-south direction aligned with the prevailing southerly wind. Dunes typically reach 10 m in height and are 0.6–1 km apart (Dickman et al. 2010). Vegetation in the interdune swales and on dune sides is predominantly spinifex grassland (Triodia basedowii) with small stands of gidgee trees

163

Chapter 6: Intraguild predator interactions

(Acacia georginae), woody Acacia shrubs or mallee eucalypts; low-lying clay pans fill with

water temporarily after heavy rain.

During summer, daily temperatures usually exceed 40º C and minima in winter fall

below 5º C (Dickman et al. 2010). Highest rainfall occurs in summer, but heavy rains can fall

locally or regionally at any time in the year. Long-term weather stations in the study area are

at Glenormiston (1890–2011), Boulia (1888–2011) and Birdsville (1954–2011), and have

median annual rainfalls of 186 mm (n = 122 years), 216.2 mm (n = 124 years), and 153.1 mm

(n = 57 years), respectively (Greenville et al. 2012, Chapter 2).

Remote cameras

To survey the predators and their rodent prey and estimate their relative abundances, we

placed 25 remote cameras (24 Moultrie i40 and one Reconyx RapidFire) 1–10 km apart next

to access tracks in spinifex habitat in the interdune swales, as animals—especially the

predators—frequently use these tracks (Mahon et al. 1998). Two years of continuous monitoring were required to capture any lag (up to 12 months, see above) in changes in predator numbers following irruptions of prey (Fig. A4.1). Cameras were mounted atop 1.5 m metal stakes, and angled at ~10° so the field of view covered the track; their locations were assumed to be independent (spatial autocorrelation, Moran’s I: dingo: I = 0.15, P = 0.22, red fox: I = 0.31, P = 0.14 and feral cat: I = 0.33, P = 0.06). Cameras were active from April

2010 to April 2012 and downloaded 3–4 times a year. Each photograph was tagged with the site name, camera identification number, download trip, moon phase, species and number of individuals recorded, and the tags written to the exif data of each file (jpeg) using EXIFPro

2.0 (Kowalski and Kowalski 2012). EXIFPro 2.0 was used to database the photographs and export the exif data as a text file for analysis. To ensure independence, a delay of one minute

164

Chapter 6: Intraguild predator interactions was programmed on-camera between each trigger, and multiple photographs of the same presumptive individual (photographs taken < 2 minutes apart) were removed prior to analysis. This resulted in a total of at least three minutes between photographs. Histograms were inspected for each species to confirm that is was an appropriate breakpoint (Fig. A4.2)

The major prey of the three predator species, rodents, irrupted across all sites in the study region after flood rains in 2010 (Greenville et al. 2012, 2013, Chapters 2 and 3). Population phases were defined as bust (April 2010–December 2010), boom (January 2011–September

2011) and decline (October 2011–April 2012) from inspection of camera trapping records

(Fig. A4.3) and also from a concurrent live-trapping study (see Greenville et al. 2012,

Chapters 2 and 4; Fig. A4.3), as small mammals may have lower detection rates compared to the larger predators. The numbers of photographs for each species (dingo, red fox, feral cat and rodents—all species combined) were pooled for each of these phases.

Data analyses

To test our predictions, we used negative binomial generalized linear models (GLMs) to compare numbers of photographs of the three predators in the bust, boom and decline phases, and the predator × phase interaction. The negative binomial distribution was chosen instead of the Poisson as the count data showed signs of over-dispersion (Zuur 2009). Due to camera failures, wildfire or memory cards filling up before downloads, cameras sometimes had different sampling effort. To calculate sampling effort per camera, we recorded every day each camera was active over the two-year period. Summing these records yielded a total of

10 260 active camera days and nights. An offset for number of days each camera was active was used in our models to account for unequal camera effort. Analysis of Deviance was used

165

Chapter 6: Intraguild predator interactions

to test the effects and followed a chi-squared distribution (Zuur 2009). Analyses used the

MASS 7.3-16 package (Venables and Ripley 2002), in R 2.14.1 (R Core Team 2014).

To detect differences in activity times of the dingo, the mesopredators and prey

(rodents), the time stamps for photographs of each predator and for rodents were pooled for each phase. Activity times followed a circular distribution over 24 hours. To test for differences in activity times of predators within and between phases, and for any coincidence of activity times between the predators and prey, we used ANOVAs with a likelihood ratio test that followed the chi-squared distribution (Cordeiro et al. 1994). To estimate confidence intervals for mean activity times, bootstrapping with 1000 replicates was used. Analysis used the Circular 0.4-3 package (Agostinelli and Lund 2011) in R 2.14.1 (R Core Team 2014).

Results

In total, 1247 independent photographs of the three predator species were captured over the period of study, and comprised 626 images of feral cats, 238 of dingoes, 383 of red foxes, and 483 of rodents. Four species of rodents were identified (long-haired rat Rattus villosissimus, sandy inland mouse Pseudomys hermannsburgensis, spinifex hopping mouse

Notomys alexis and house mouse Mus musculus) and their images pooled. In general, photographs of rodents showed the presence of multiple individuals and photographs of predators had only one individual, but each photograph was considered an event. There was a significant interaction between the dingo and population phase of prey with the number of photographic records of the two species of mesopredators (Table 6.1, Table A4.1). In the boom phase, records of the dingo and red fox were positively associated, but in the bust and decline phases records of the fox decreased with increasing dingo activity (Fig. 6.3a). In contrast, as dingoes increased, there was a decrease in the number of feral cat records in all

166

Chapter 6: Intraguild predator interactions

phases (Fig. 6.3b). The red fox and feral cat increased together in all phases, with cats

increasing at a progressively faster rate during bust conditions (Fig. 6.3c). To assess the effect of the outliers in Figs 6.3a and b, they were removed and the tests re-run. Both analyses yielded significant results, thus indicating that the outliers were not unduly important in driving the observed relationships (Table A4.2).

Table 6.1: GLM results for number of detections of animals in photographs from remote camera trapping of dingoes, red fox and feral cat in the Simpson Desert, southwestern Queensland. Cameras were active for two years and photographs were pooled for each phase each year. Phase is the population phase of prey (bust, boom or decline).

Interaction df Deviance Residual df Residual P deviance

Dingo-fox interaction

Dingo 1 93.78 71 15569 <0.001 Phase 2 1758.56 69 13811 <0.001 Dingo × Phase 2 2222.00 67 11589 <0.001

Dingo-cat interaction

Dingo 1 195.78 71 12497 <0.001

Phase 2 895.57 69 11601 <0.001 Dingo × Phase 2 130.1 67 11471 <0.001

Fox-cat interaction Fox 1 2186.36 71 12551 <0.001 Phase 2 600.07 69 11951 <0.001 Fox × Phase 2 479.67 67 11471 <0.001

167

Chapter 6: Intraguild predator interactions

Figure 6.3: Interactions between prey population phase (bust, boom and decline) and the effects of dominant predators on subordinate predators: a) dingo-red fox, b) dingo-feral cat and c) red fox-feral cat, represented by numbers of photographic images from two years of continuous remote camera trapping in the Simpson Desert, central Australia. Log numbers are shown on the y-axis. Solid and dashed lines are predicted values from GLMs.

168

Chapter 6: Intraguild predator interactions

Mean daily activity times (95% confidence intervals) for dingoes, red foxes and feral cats in the bust phase were 2:12 h (0:32–3:50 h), 23:23 h (22:31–0:17 h) and 23:59 h (23:16–

0:44 h), respectively (Fig. 6.4a). For the boom phase, respective mean activity times were

2:03 h (0:58–3:11 h), 23:42 h (22:49–0:38 h) and 0:18 h (23:46–0:52 h), (Fig. 6.4b) and for the decline phase 2:47 h (1:56–3.38 h), 0:40 h (23:47–1:29 h) and 0:52 h (0:11–1:36 h) (Fig.

6.4c). Daily activity times differed between the three predator species in the bust (χ2 = 12.95,

P = 0.002, df = 2), boom (χ2 = 12.22, P = 0.002, df = 2) and decline phases (χ2 = 14.06, P =

0.001, df = 2), but there was no difference in each predator’s activity time between phases:

dingo (χ2 = 1.12, P = 0.57, df = 2), red fox (χ2 = 4.09, P = 0.13, df = 2) and feral cat (χ2 =

3.71, P = 0.16, df = 2). The red fox showed some activity by day in the boom and decline phases compared to the bust, the dingo showed some daytime activity in all phases, but feral cats were entirely nocturnal (Fig. 6.4a,b,c).

Rodents were active only at night, with mean activity times (95% confidence

intervals) for bust, boom and decline population phases being 23:36 h (22:09–0:59 h), 23:31

h (23:07–23:58 h) and 0:46 h (0:11–1:17 h), respectively. Activity times differed between

phases (χ2 = 9.209, P = 0.01, df = 2), shifting to early morning in the decline phase (Fig.

6.4a,b,c). There was no difference between the mean activity times of the red fox and rodents

for any population phase (bust: χ2 = 0.08, P = 0.8, df = 1; boom: χ2 = 0.17, P = 0.7, df = 1;

decline: χ2 = 0.04, P = 0.8, df = 1). The activity of feral cats coincided with that for rodents

during the bust (χ2 = 0.27, P = 0.3, df = 1) and decline phases (χ2 =0.05, P = 0.8, df = 1), but

not in the boom phase (χ2 = 5.8, P = 0.02, df = 1). Dingoes differed in mean activity time

from rodents in all population phases (Bust: χ2 = 5.6, P = 0.02, df = 1; Boom: χ2 = 22.6,

P<0.001, df = 1; Decline: χ2 = 14.7, P< 0.001, df = 1).

169

Chapter 6: Intraguild predator interactions

Figure 6.4: Daily activity patterns of the dingo, red fox, feral cat and rodents (prey) during a) bust, b) boom and c) decline phases of the prey population, Simpson Desert, central Australia. Numbers represent 24-hour clock times, arrows represent mean activity times and grey boxes frequency (square-root) of observations recorded each hour from two years of continuous remote camera trapping.

Discussion

The results broadly support our initial predictions, and show that interactions between dominant and subordinate predators can shift markedly during the course of a bottom-up resource pulse. Bottom-up processes are important in structuring populations of sympatric predators in other systems (Jaksic et al. 1997, Elmhagen and Rushton 2007, White 2008), suggesting that such effects may be general. In the present study, top-predator suppression

170

Chapter 6: Intraguild predator interactions was most obvious in the consistently negative relationship between the dingo and feral cat, but was evident also in the negative relationship between dingoes and foxes in the bust and decline phases of the rodent populations. As predicted for this latter pair, foxes evidently escaped suppression by dingoes only during boom conditions when rodents were very abundant. Although the relationship between fox and cat populations was positive in each of the three population phases of their prey, the relatively greater increase in cat numbers during bust conditions when foxes were scarce is consistent with the interpretation that the cat is the subordinate predator in the desert system. We interpret shifts in pairwise interactions between the predators below.

Top-down forcing of foxes by dingoes was most obvious during periods of rodent decline and bust. Competition and intraguild killing of subordinate by dominant predators are most likely when prey is scarce (Donadio and Buskirk 2006), and suggest that top-down processes may generally dominate at these times. In an informative recent study, however,

Moseby et al. (2012) showed that dingoes kill but do not eat foxes upon encounter with them, suggesting that the interaction is one of extreme interference competition. If this is correct, it is relevant to ask why dingoes do not maintain top-down forcing of foxes during prey irruptions.

In the first instance, while dingoes and foxes often show large overlaps in diet, dingoes hunt and kill larger prey if these are available (Cupples et al. 2011, Spencer et al. 2014). If larger prey, such as kangaroos, disperse after rain and become available to dingoes, this may reduce competition for shared small prey between the two canid species and allow foxes to escape suppression. While possible, this explanation seems unlikely. Kangaroos and other large potential prey usually show limited responses even to large pulses of primary productivity in spinifex grasslands and remain present but at consistently low density (Letnic and Dickman 2006); the dominant vertebrates are small mammals, reptiles and birds.

171

Chapter 6: Intraguild predator interactions

Populations of some reptile species respond after rainfall, although breeding takes place only

during the austral spring and summer, and sexual maturity occurs after 11–12 months (James

1991, Greenville and Dickman 2005). Thus reptile populations respond slowly to rainfall events compared to rodent populations. Many species of birds exhibit population irruptions after large rainfall events (Tischler et al. 2013), but this prey type represents only 6–12% of the diets of the three predators (Pavey et al. 2008, Cupples et al. 2011).

Secondly, the negative correlation between dingoes and the smaller predators may be due to the cat and red fox responding more quickly to increases and decreases in prey compared to the dingo, while the dingo switches to larger alternative prey as small mammals become scarce. However, the red fox is usually scarce in the study region and moves into the sand dune environment from more mesic areas after large rainfall events (Letnic and

Dickman 2006). Thus the arrival and subsequent breeding of red foxes corresponds to the boom and decline phases of the rodents. Cats are persistent in the region but, as kangaroos and larger prey species show limited population responses to rainfall in the study region, prey switching by dingoes does not seem likely.

Thirdly, rain-stimulated increases in primary productivity leads to increased cover of ground-level vegetation in the study system (Dickman et al. 1999), and this in turn may provide foxes with sufficient shelter to escape detection by dingoes. Fourthly, and perhaps most plausibly, bottom-up pulses of productivity may allow such large increases in rodent populations that the benefits of ready access to abundant food for dingoes exceed the costs of interference. In arid Australia, rapid 10–60-fold increases in rodent populations occur after rain (Dickman et al. 1999, 2010) and these alone may be enough to reduce top-down effects, at least between the dominant and subordinate canid predators. Elsewhere in arid Australia,

Lundie-Jenkins et al. (1993) found that competition between dingoes and mesopredators was alleviated during irruptions of rodents and locusts.

172

Chapter 6: Intraguild predator interactions

In contrast to the temporal shifts in interactions between the canid predators, dingoes

suppressed populations of feral cats in all phases of the prey boom and bust cycle, suggesting

that top-down effects generally predominate. As the smallest predators in the study system,

cats could be expected to be subordinate to both the dingo and the fox. Dingoes kill cats

quickly if they encounter them (Moseby et al. 2012), and dietary evidence confirms that foxes also may kill and eat cats, albeit infrequently (Molsher 1999, Paltridge 2002). Despite this, the significant phase interaction in the comparison of cat and dingo abundance suggests that cats may experience some alleviation in top-down forcing during boom compared to bust periods (Fig. 6.3b). If cats share rodent prey with both canids and also must avoid encounters with foxes in dingo-free space, this could explain why cats may alleviate top-down forcing during boom periods but not escape it altogether. The increase in cat abundance relative to that of the fox during bust periods, when this canid predator is often very scarce (Fig. 6.3c), suggests further that cats may escape suppression by foxes but not by dingoes during these times.

We also expected the predators to alter their times of daily activity in relation to each other and to their rodent prey, and our results here provide more insight into the interactions we observed. In the first instance, irrespective of the population phase of the rodents, the red fox and feral cat showed peak activity around the middle of the night some 2 hours before peak activity of the dingo. Although all three predator species showed some activity during all the hours of darkness, this staggering in activity time between the mesopredators and the top-predator would potentially reduce the likelihood of encounters (Brook et al. 2012). The lack of difference in activity times of the fox and cat is harder to reconcile if a dominant- subordinate relationship exists between them. During the boom phase, however, coinciding activity with the peak activity times of prey would likely yield energetic benefits that offset the risks of mutual encounter. A similar explanation holds for the lack of temporal

173

Chapter 6: Intraguild predator interactions

partitioning among the large carnivores of Africa (Cozzi et al. 2012). During the decline and bust phases, when prey populations are low, it is possible that the two mesopredators were active at the same times but in different places, but insufficient observations were obtained across all the camera stations to test this.

We had expected also that the daily activity of the dingo would coincide most closely with that of its rodent prey, but this expectation was not met. Instead, fox, and then cat, activity showed the strongest coincidence in timing with the activity of rodents, with mean dingo activity peaking at least 2 hours after the equivalent peak for rodents. For foxes and cats the benefits of focussing their activity when prey is most active may, as noted, exceed the potential risks of occasional mutual encounter. For the dingo, however, the relative delay in activity may suggest that, while it spends some time foraging for rodents, it has exclusive access to larger albeit rarer prey, such as kangaroos, that are easier to hunt at or just after dawn. Two pieces of evidence support this interpretation. Firstly, the dingo was more active during the early hours of daylight than the smaller predators (Fig. 6.4). Secondly, large mammals, especially red kangaroos, comprise 23–44% of the diet of the dingo by frequency of occurrence in the study area, compared to just 3–4% of the diet of the fox (Cupples et al.

2011, Spencer et al. 2014). By contrast, the respective representation of small mammals in the diets of these predators is 54% and 86% (Cupples et al. 2011).

Our observations are broadly consistent with the idea that top-down forcing constrains populations of mesopredators in our study system, and that this is most obvious during periods of low prey activity. However, alternative possibilities need to be addressed before any further implications can be drawn. Most importantly, as this is an observational study, we cannot be certain that we are uncovering causal relationships between the predators. For example, changes in plant cover between boom and bust phases of prey may provide the mesopredators with different food or cover resources at these times. An increase in plant

174

Chapter 6: Intraguild predator interactions

cover during the boom phase is the most obvious of these changes but, as noted above, does

not readily explain why only foxes may exploit it to escape top-down suppression. There are,

alternatively, other predators in the study system; these potentially could introduce indirect

effects that mask the relationships between the dingo and the two smaller mammalian

predators (Billick and Case 1994). For example, the goannas Varanus giganteus and V.

gouldii and wedge-tailed eagle Aquila audax hunt rodents and larger prey (Aumann 2001,

Pianka et al. 2004). However, all these other predators are diurnal (Aumann 2001, Pianka et

al. 2004) and probably do not interact in any direct way with the three largely nocturnal

mammalian predators. Habitat shifts by the dingo across the prey phases may lead to negative

associations between dingoes and the mesopredators. For example, in the rodent bust phase, populations of the red fox and cat decline, but dingoes may be able to shift their foraging to alternative habitats with larger prey. Further research is required to test this possibility. In

view of the high dietary overlap of the mammalian predators in our study system (Mahon

1999, Cupples et al. 2011), the extreme interference effects of dingoes on the two

mesopredators (Moseby et al. 2012), and evidence of top-down suppression by dingoes in

other habitats (Johnson and VanDerWal 2009, Letnic and Koch 2010, Letnic et al. 2011,

Colman et al. 2014), it seems reasonable to conclude that our camera-based observations

reflect real shifts in top-predator-mesopredator interactions.

In summary, our results suggest that, in the resource-pulse conditions that characterize

much of the Australian environment, mesopredator suppression by dingoes is not constant but

is most marked in the bust and decline phases of prey populations (Fig. 6.5). The presence of

the top-predator may provide particularly strong net benefits to small native prey species at

these times when they are at low numbers and perhaps at their most vulnerable to additive

predator impacts and stochastic extinction risk (Sinclair et al. 1998). Given that resource

pulses are often infrequent and ephemeral, as they are when driven by flood rains in arid and

175

Chapter 6: Intraguild predator interactions semi-arid environments, we suggest that top-predators like the dingo will generally benefit small prey by suppressing mesopredators during the prolonged bust periods when prey populations are low.

Figure 6.5: Complex dominant-subordinate predator interactions in a) the boom phase and b) the bust or decline phases of their prey. Line weights represent relative interaction strengths. Dashed lines represent indirect interactions. This model predicts that, in times when the prey population is most vulnerable to predation from the mesopredators (bust and decline phases), the top predator has a net positive effect on prey. Artwork by A. Foster.

176

Chapter 6: Intraguild predator interactions

Acknowledgements

We thank Bush Heritage Australia and G. Woods for allowing access to the properties in the study region; members of the Desert Ecology Research Group, N. Hills, D. Nelson and

G. Madani for valuable assistance in the field, and V. Nguyen for statistical advice. Special thanks to A. Foster for providing the artwork. ACG was supported by an Australian

Postgraduate Award and Paddy Pallin Grant, Royal Zoological Society of NSW. Funding support for GW and CD was provided by the Australian Research Council and by the

Australian Government’s Terrestrial Ecosystems Research Network (www.tern.gov.au), an

Australian research infrastructure facility established under the National Collaborative

Research Infrastructure Strategy and Education Infrastructure Fund – Super Science Initiative

through the Department of Industry, Innovation, Science, Research and Tertiary Education.

177

Chapter 6: Intraguild predator interactions

References:

Agostinelli, C. and U. Lund. 2011. R package 'circular': circular statistics (version 0.4-3). Aumann, T. 2001. Habitat use, temporal activity patterns and foraging behaviour of raptors in the south-west of the Northern Territory, Australia. Wildlife Research 28: 365-378. Billick, I. and T. J. Case. 1994. Higher order interactions in ecological communities: what are they and how can they be detected? Ecology 75: 1529-1543. Brook, L. A., C. N. Johnson, and E. G. Ritchie. 2012. Effects of predator control on behaviour of an apex predator and indirect consequences for mesopredator suppression. Journal of Applied Ecology 49: 1278-1286. Caughley, G. 1987. Ecological relationships, Pages 159-187 in G. Caughley, N. Shepherd, and J. Short, editors. Kangaroos: their ecology and management in the sheep rangelands of Australia. Cambridge University Press, New York. Colman, N. J., C. E. Gordon, M. S. Crowther, and M. Letnic. 2014. Lethal control of an apex predator has unintended cascading effects on forest mammal assemblages. Proceedings of the Royal Society B: Biological Sciences 281: 20133094. Cordeiro, G., G. Paula, and D. Botter. 1994. Improved likelihood ratio tests for dispersion models. International Statistical Review 62: 257-274. Cozzi, G., F. Broekhuis, J. W. McNutt, L. A. Turnbull, D. W. Macdonald, and B. Schmid. 2012. Fear of the dark or dinner by moonlight? Reduced temporal partitioning among Africa's large carnivores. Ecology 93: 2590-2599. Crowther, M. S., M. Fillios, N. Colman, and M. Letnic. 2014. An updated description of the Australian dingo (Canis dingo Meyer, 1793). Journal of Zoology 293: 192-203. Cupples, J. B., M. S. Crowther, G. Story, and M. Letnic. 2011. Dietary overlap and prey selectivity among sympatric carnivores: could dingoes suppress foxes through competition for prey? Journal of Mammalogy 92: 590-600. Dickman, C. R., A. C. Greenville, C.-L. Beh, B. Tamayo, and G. M. Wardle. 2010. Social organization and movements of desert rodents during population “booms” and “busts” in central Australia. Journal of Mammalogy 91: 798-810. Dickman, C. R., P. S. Mahon, P. Masters, and D. F. Gibson. 1999. Long-term dynamics of rodent populations in arid Australia: the influence of rainfall. Wildlife Research 26: 389-403. Donadio, E. and S. W. Buskirk. 2006. Diet, morphology, and interspecific killing in Carnivora. The American Naturalist 167: 524-536.

178

Chapter 6: Intraguild predator interactions

Elmhagen, B. and S. P. Rushton. 2007. Trophic control of mesopredators in terrestrial ecosystems: top-down or bottom-up? Ecology Letters 10: 197-206. Estes, J. A., J. Terborgh, J. S. Brashares, M. E. Power, J. Berger, W. J. Bond, S. R. Carpenter, T. E. Essington, R. D. Holt, J. B. Jackson, R. J. Marquis, L. Oksanen, T. Oksanen, R. T. Paine, E. K. Pikitch, W. J. Ripple, S. A. Sandin, M. Scheffer, T. W. Schoener, J. B. Shurin, A. R. Sinclair, M. E. Soule, R. Virtanen, and D. A. Wardle. 2011. Trophic downgrading of planet Earth. Science 333: 301-306. Fedriani, J. M., T. K. Fuller, R. M. Sauvajot, and E. C. York. 2000. Competition and intraguild predation among three sympatric carnivores. Oecologia 125: 258-270. Fleming, P. J. S., B. L. Allen, and G.-A. Ballard. 2012. Seven considerations about dingoes as biodiversity engineers: the socioecological niches of dogs in Australia. Australian Mammalogy 34: 119-131. Greenville, A. C. and C. R. Dickman. 2005. The ecology of Lerista labialis (Scincidae) in the Simpson Desert: reproduction and diet. Journal of Arid Environments 60: 611-625. Greenville, A. C., G. M. Wardle, and C. R. Dickman. 2012. Extreme climatic events drive mammal irruptions: regression analysis of 100-year trends in desert rainfall and temperature. Ecology and Evolution 2: 2645-2658. Greenville, A. C., G. M. Wardle, and C. R. Dickman. 2013. Extreme rainfall events predict irruptions of rat plagues in central Australia. Austral Ecology 38: 754-764. Jaksic, F. M., S. I. Silva, P. L. Meserve, and J. R. Gutierrez. 1997. A long-term study of vertebrate predator responses to an El Nino (ENSO) disturbance in western South America. Oikos 78: 341-354. James, C. D. 1991. Annual variation in reproductive cycles of scincid lizards (Ctenotus) in central Australia. Copeia 1991: 744-760. Johnson, C. N., J. L. Isaac, and D. O. Fisher. 2007. Rarity of a top predator triggers continent- wide collapse of mammal prey: dingoes and marsupials in Australia. Proceedings of the Royal Society B: Biological Sciences 274: 341-346. Johnson, C. N. and E. G. Ritchie. 2013. The dingo and biodiversity conservation: response to Fleming et al. (2012). Australian Mammalogy 35: 8-14. Johnson, C. N. and J. VanDerWal. 2009. Evidence that dingoes limit abundance of a mesopredator in eastern Australian forests. Journal of Applied Ecology 46: 641-646. Kowalski, M. and M. Kowalski. 2012. EXIFPro 2.0, California.

179

Chapter 6: Intraguild predator interactions

Letnic, M. and C. R. Dickman. 2006. Boom means bust: interactions between the El Niño/Southern Oscillation (ENSO), rainfall and the processes threatening mammal species in arid Australia. Biodiversity and Conservation 15: 3847-3880. Letnic, M. and C. R. Dickman. 2010. Resource pulses and mammalian dynamics: conceptual models for hummock grasslands and other Australian desert habitats. Biological Reviews 85: 501-521. Letnic, M., A. Greenville, E. Denny, C. R. Dickman, M. Tischler, C. Gordon, and F. Koch. 2011. Does a top predator suppress the abundance of an invasive mesopredator at a continental scale? Global Ecology and Biogeography 20: 343-353. Letnic, M. and F. Koch. 2010. Are dingoes a trophic regulator in arid Australia? A comparison of mammal communities on either side of the dingo fence. Austral Ecology 35: 167-175. Letnic, M., E. G. Ritchie, and C. R. Dickman. 2012. Top predators as biodiversity regulators: the dingo Canis lupus dingo as a case study. Biological Reviews 87: 390-413. Lundie-Jenkins, G., L. Corbett, and C. Phillips. 1993. Ecology of the rufous hare-wallaby, Lagorchestes hirsutus Gould (Marsupialia : Macropodidae) in the , Northern Territory. III Interactions with introduced mammal species. Wildlife Research 20: 495-511. Mahon, P. S. 1999. Predation by feral cats and red foxes and the dynamics of small mammal populations in arid Australia. PhD Thesis. University of Sydney, Sydney. Mahon, P. S., P. B. Banks, and C. R. Dickman. 1998. Population indices for wild carnivores: a critical study in sand-dune habitat, south-western Queensland. Wildlife Research 25: 11-22. Merkle, J. A., D. R. Stahler, and D. W. Smith. 2009. Interference competition between gray wolves and coyotes in Yellowstone National Park. Canadian Journal of Zoology 87: 56-63. Molsher, R. L. 1999. The ecology of feral cats, Felis catus, in open forest in New South Wales: interactions with food resources and foxes. PhD Thesis. University of Sydney, Sydney. Moseby, K. E., H. Neilly, J. L. Read, and H. A. Crisp. 2012. Interactions between a top order predator and exotic mesopredators in the Australian rangelands. International Journal of Ecology Article 250352. Palomares, F. and T. M. Caro. 1999. Interspecific killing among mammalian carnivores. The American Naturalist 153: 492-508. 180

Chapter 6: Intraguild predator interactions

Paltridge, R. 2002. The diets of cats, foxes and dingoes in relation to prey availability in the Tanami Desert, Northern Territory. Wildlife Research 29: 389-403. Pavey, C. R., S. R. Eldridge, and M. Heywood. 2008. Population dynamics and prey selection of native and introduced predators during a rodent outbreak in arid Australia. Journal of Mammalogy 89: 674-683. Pianka, E. R., D. R. King, and R. A. King. 2004. Varanoid lizards of the world. Indiana University Press, IN. Power, M. E. 1992. Top-down and bottom-up forces in food webs: do plants have primacy? Ecology 73: 733-746. Prugh, L. R., C. J. Stoner, C. W. Epps, W. T. Bean, W. J. Ripple, A. S. Laliberte, and J. S. Brashares. 2009. The rise of the mesopredator. Bioscience 59: 779-791. R Core Team. 2014. R: a language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria. Ripple, W. J. and R. L. Beschta. 2012. Trophic cascades in Yellowstone: the first 15 years after wolf reintroduction. Biological Conservation 145: 205-213. Ripple, W. J., J. A. Estes, R. L. Beschta, C. C. Wilmers, E. G. Ritchie, M. Hebblewhite, J. Berger, B. Elmhagen, M. Letnic, M. P. Nelson, O. J. Schmitz, D. W. Smith, A. D. Wallach, and A. J. Wirsing. 2014. Status and ecological effects of the world's largest carnivores. Science 343: 1241484. Ritchie, E. G., B. Elmhagen, A. S. Glen, M. Letnic, G. Ludwig, and R. A. McDonald. 2012. Ecosystem restoration with teeth: what role for predators? Trends in Ecology & Evolution 27: 265-271. Sinclair, A. R. E., R. P. Pech, C. R. Dickman, D. Hik, P. Mahon, and A. E. Newsome. 1998. Predicting effects of predation on conservation of endangered prey. Conservation Biology 12: 564-575. Smith, A. P. and D. G. Quin. 1996. Patterns and causes of extinction and decline in Australian conilurine rodents. Biological Conservation 77: 243-267. Spencer, E., M. Crowther, and C. Dickman. 2014. Diet and prey selectivity of three species of sympatric mammalian predators in central Australia. Journal of Mammalogy in press. Tischler, M., C. R. Dickman, and G. M. Wardle. 2013. Avian functional group responses to rainfall across four vegetation types in the Simpson Desert, central Australia. Austral Ecology 38: 809-819.

181

Chapter 6: Intraguild predator interactions

Van Etten, E. J. B. 2009. Inter-annual rainfall variability of arid Australia: greater than elsewhere? Australian Geographer 40: 109-120. Venables, W. N. and B. D. Ripley. 2002. Modern Applied Statistics with S. Fourth Edition edition. Springer, New York. Vine, S. J., M. S. Crowther, S. J. Lapidge, C. R. Dickman, N. Mooney, M. P. Piggott, and A. W. English. 2009. Comparison of methods to detect rare and cryptic species: a case study using the red fox (Vulpes vulpes). Wildlife Research 36: 436-446. Ward, D. 2009. The biology of deserts. Oxford University Press, New York. White, T. C. R. 2008. The role of food, weather and climate in limiting the abundance of animals. Biological Reviews 83: 227-248. Yang, L. H., K. F. Edwards, J. E. Byrnes, J. L. Bastow, A. N. Wright, and K. O. Spence. 2010. A meta-analysis of resource pulse–consumer interactions. Ecological Monographs 80: 125-151. Zuur, A. F. 2009. Mixed effects models and extensions in ecology with R. Springer, New York ; London.

182

Chapter 7: Predicting species interactions

Chapter 7: Predicting species interactions under a

changing climate

A thunderstorm roles across the Simpson Desert. Lightning strikes are the primary ignition source for wildfires in our study region. Photograph by Aaron Greenville.

183

Chapter 7: Predicting species interactions

Abstract

Species interact as they seek food, shelter and breeding sites and these interactions are

in turn, influenced by the environment. Together, this web of biotic and abiotic interactions

drives the composition and functioning of ecological assemblages and in doing so supports the on-going diversity of the system. Climate as a key environmental driver, acts either directly or indirectly on species interactions and therefore, climate change is of central interest for ecological models of species persistence. Previous studies usually have focused on non-interactive effects of global climate change across trophic levels and used single or independent multiple population drivers. However, to derive more reliable and accurate

predictions, simultaneous analyses of multiple drivers are needed to investigate the

interactive effect of changes in population drivers due to climate change. Here we use

structural equation modelling to integrate results from remote camera trapping and live-

trapping of vertebrates with long term (17–22 years) vegetation data from the Simpson Desert

to investigate interactions between the biota and two key external drivers, rainfall and fire.

We then use these models to predict how changes in rainfall and wildfire events, in-line with

future climate scenarios, will permeate up the trophic levels and interact with top-down effects from mammalian carnivores over the next 100 years. Our results show that simultaneous analysis of external population drivers is important for structuring interactions between species and between trophic groups. Population responses were influenced by both bottom-up and top-down regulation, highlighting the need for models to incorporate both external population drivers and species interactions when predicting species’ responses to climate change.

184

Chapter 7: Predicting species interactions

Introduction

Species interactions, both within and between trophic groups, are important for

ecosystem functioning and maintaining diversity (Bruno et al. 2003, Baskett and Salomon

2010) and may occur either within or between trophic groups. For example, the removal of

apex predators can cause cascades that extend to lower trophic groups (Heithaus et al. 2008,

Ripple et al. 2014) or competition between granivorous species may limit each species’ population growth (Davidson et al. 1985). However, species interactions, such as those that facilitate nutrient cycles, plant-pollinator networks or predator-prey relationships can be modified due to a changing climate (Schweiger et al. 2010, Traill et al. 2010). Most studies have investigated non-interactive effects of global climate change across trophic levels and focused on single or independent multiple drivers (Tylianakis et al. 2008). Other studies have focused on the ecology of individual species or species groups using habitat or niche modelling and dispersal capacities, while ignoring biotic interactions and evolutionary considerations (Lavergne et al. 2010). Biotic interactions can have profound effects when predicting responses in species composition due to climate change (e.g. Suttle et al. 2007), leading to the current need for ecologists to investigate the interactive effect of changes in population drivers from climate change (Elmhagen et al. 2015, Ross et al. 2015).

Climate change has the potential to increase or decrease the strength of interactions and may act differently if the interactions are positive or negative, leading to unpredictable changes in the composition of assemblages. Identifying the importance of particular biotic interactions within broader networks of interactions is difficult, particularly when multiple trophic levels are involved, but they can be useful for detecting the first evidence of an altered environment (Popic and Wardle 2012). For example, small shifts in plant phenology may have large effects on mutualisms, such as plant-pollination networks (Memmott et al.

185

Chapter 7: Predicting species interactions

2007). Extreme weather events, such as droughts, can disrupt flowering phenology, leading to

declines or even local extinctions of pollinators, which in turn can have cascading effects on

frugivores and seed dispersal (Harrison 2000, 2001).

Antagonistic interactions may also intensify under a changing climate (Tylianakis et

al. 2008, Traill et al. 2010). For example, a review of 688 studies found consistently that

competitive interactions between plants were shifting due to climate change (Tylianakis et al.

2008). However, predator-prey interactions could be negatively or positively affected

(Tylianakis et al. 2008), highlighting the difficulties of reliably predicting species’ population

changes and the need for studies into individual systems. In addition, the timing in the

importance of bottom-up and top-down control can change due to climate (Letnic et al.

2011b). For example, in dry years, which predominate in central Australia, top-predators may

limit introduced mesopredators and provide some relief from predation pressure for small

prey species (Greenville et al. 2014, Chapter 6). However, after irruptions in small prey species due to rain-induced resource pulses, this relationship weakens and mesopredators numbers increase (Greenville et al. 2014, Chapter 6). In central Australia, predation pressure on small prey species can be higher from introduced predators compared to native predators

(Pavey et al. 2008) and therefore any increases in the frequency and intensity of rainfall events, which favour higher introduced predator numbers, could have negative consequences

for many smaller prey species.

Climate plays an important role in driving other abiotic processes important for

influencing species populations. For example, global phases of the El Nino-Southern

Oscillation cycle can drive patterns of wildfires across the Americas and Oceania (Le Page et

al. 2008). In central Australia, risk from wildfires increases after large, yet rare, rainfall

events (Letnic and Dickman 2006, Greenville et al. 2009, Nano et al. 2012). Large rainfall

events are increasing in frequency and magnitude in central Australia and thus may promote

186

Chapter 7: Predicting species interactions

higher fuel loads (Greenville et al. 2012, Chapter 2). In addition, temperatures have increased

over the last 100 years, which also increases evaporation rates, and thus more frequent and

intense wildfires are predicted (Cary et al. 2012, Greenville et al. 2012, Chapter 2).

Here we use a trophic interaction model to integrate past research from this thesis and

long-term (17–22 years) studies in central Australia, and propose a conceptual model that describes the interactions between abiotic drivers and effects on multiple trophic levels in the

Simpson Desert (see Fig. 7.1 for modelling steps). To identify significant interactions, we applied structural equation modelling to long-term datasets on rainfall, wildfire history, vegetation surveys, captures of small mammals and reptiles, and a shorter two-year survey of mammalian predators. We then use the results from the structural equation model to make the following projections, in-line with future climate predictions from the study region

(Greenville et al. 2012):

1. Increases in mean rainfall will compensate for increases in wildfire frequency and

lead to greater magnitude of seeding events in the dominant vegetation (spinifex),

which, in turn, will lead to increases in rodent numbers.

2. Removal of introduced mammalian predators will lead to an increase in rodent

captures due to alleviation of strong top-down suppression; in contrast, removal of

both introduced and native mammalian predators will not lead to as great an increase

in rodent captures due to facilitation effects from the top predator.

187

Chapter 7: Predicting species interactions

Develop trophic interaction Past studies model

Thesis chapters

Time series Test with datasets SEM

Knowledge of Refined trophic biotic and interaction abiotic model processes

Simulated rainfall change

Simulations

Simulated wildfire change Species Species population responses to predictions climate change

Datasets Modelling steps Expert knowledge

Figure 7.1: Overview of the modelling process used to integrate multiple datasets into a trophic interactions model, informed by the results of this and past studies. Once significant pathways (biotic and abiotic interactions) are found using structural equation modelling (SEM), the model is refined and used to predict species’ population changes due to changes in rainfall and wildfire, in line with climate change scenarios.

188

Chapter 7: Predicting species interactions

Methods and materials

Study region

The study was carried out in the Simpson Desert, central Australia. This region

occupies 170 000 km2; dune fields comprise 73% of the region, with smaller areas consisting

of clay pans, rocky outcrops and gibber flats (Shephard 1992). The sand dunes run parallel in

a north-south direction aligned with the prevailing southerly wind. The dunes are up to 10 m

high and spaced 0.6–1 km apart (Purdie 1984). Vegetation in the interdune swales and on

dune sides is predominantly spinifex grassland (Triodia basedowii) with patchy cover of

small stands of gidgee trees (Acacia georginae) and other woody Acacia shrubs or mallee

eucalypts; low-lying clay pans fill with water temporarily after heavy rain.

During summer, daily temperatures usually exceed 40º C and minima in winter often

fall below 5º C (Purdie 1984). Highest rainfall occurs in summer, but heavy rains can fall

locally or regionally throughout the year. Long-term weather stations in the study area are at

Glenormiston (1890–2011), Boulia (1888–2011) and Birdsville (1954–2011), and have

median annual rainfalls of 186 mm (n = 122 years), 216.2 mm (n = 124 years), and 153.1 mm

(n = 58 years), respectively (Bureau of Meteorology 2012).

Small mammals and reptiles

Live-trapping was carried out at nine sites across Carlo Station, Tobermorey Station,

Cravens Peak and Ethabuka Reserves, an 8000 km2 region in the Simpson Desert, south-

western Queensland, Australia. Small mammals were live-trapped using pitfalls (16 cm diameter, 60 cm deep) each equipped with a 5 m drift fence of aluminium flywire to increase trap efficiency (Friend et al. 1989). Pitfalls were arranged on grids comprising six lines of six

189

Chapter 7: Predicting species interactions

pitfall traps spaced 20 m apart to cover 1 ha (i.e. 36 pitfall traps per grid). Sites contained 2–

12 grids and were spaced at least 10 km apart. Grids within sites were set 0.5–2 km apart in randomly chosen positions along access tracks. Traps were opened 2–6 times a year from

1990–2012 for 2–6 nights, at one site and from 1995–2012 at eight more sites (Shitty Site,

Field River South, Field River North, South Site, Kunnamuka Swamp East, Carlo,

Tobermorey East and Tobermorey West, see Fig. A2.1). Long-term (17–22 years; 130 sampling trips) live-trapping (205 524 trap nights) data yielded 11 species of small mammals and 58 species of reptiles (Chapter 4 and Table A5.1).

Predator monitoring

Dingoes (Canis dingo), feral cats (Felis catus) and introduced red foxes (Vulpes vulpes) were monitored as described in Chapter 6 and Greenville et al. (2014). Twenty-five remote camera traps were continuously in operation from April 2010 to April 2012 and the numbers of photographs of each predator species from these cameras were pooled per month for each site (see Chapter 6 for details).

Spinifex cover and seed

To measure the cover of the dominant vegetation (spinifex, T. basedowii) we scored the percentage cover in a 2.5m radius around six traps at each trapping grid that was sampled for small mammals. In addition, a seed index (0–5, where 0 represents no seeding present and

5 represents every plant with maximum seed) was used to score the amount of spinifex seeding. Spinifex seed was chosen due to its dominance in the landscape and because it is an important component of the diets of both species of study rodents, representing up to 52% of

190

Chapter 7: Predicting species interactions

their diet by occurrence (Murray and Dickman 1994). Cover and seed estimates were

averaged per trap and then for each trapping grid, per trip over 17–22 years for each of the

nine sites as in Chapter 4.

Rainfall and wildfire

Daily rainfall data from automatic weather stations (Environdata, Warwick,

Queensland) located at each site were used to calculate rainfall variables. Rainfall variables

that were identified as important predictors for mammals and reptiles are number of rain days in a month and mean event size lagged two-trips prior (Chapters 4 and 5). Rainfall generally falls from October to April, but is highly unpredictable (Letnic and Dickman 2006). For spinifex cover and seeding, 8 months cumulative rainfall was used to represent this wet period and the variability around it.

Large scale wildfires (>1000 km2) have occurred three times in the study region since

1972 and the mean wildfire return interval for the study region is approximately 26 years

(Greenville et al. 2009, NAFI 2013). To investigate if wildfire had an effect on vertebrate

captures, the time (years) since last wildfire was calculated for each trapping grid based on

remotely sensed mapping of fire scars (see Greenville et al. 2009, NAFI 2013).

Structural equation modelling

To investigate the direct and indirect interactions in our system, we used a graph-

theoretic approach to structural equation modelling (SEM). Each species or species group

was entered as a node and the direction of the arrow represented a hypothesised causal

pathway or direct interaction (Fig. 7.2a). This approach is not based on a global covariance

191

Chapter 7: Predicting species interactions

matrix, but each node is modelled separately, allowing different statistical techniques to be used for local estimation (Grace et al. 2012). Thus nine separate models were derived (see

Table 7.1 and Table A5.2). To investigate the effect of predictor variables on small mammal and reptile captures, a Poisson generalised linear mixed model (GLMM) was used, with an offset for the number of trapping nights used to account for un-equal trapping effort. Because the same trapping grid or remote camera was sampled each trip, it was added as the random factor to account for observations repeated over time (Shipley 2000). The effects of the predictor variables on spinifex cover and seed were modelled as proportional odds and a binomial GLMM was used. Lastly, predator data from remote camera traps were modelled using a Poisson GLMM, with the number of nights each camera was operational used as an offset to account for camera malfunctions. We calculated standardised coefficients for each pathway, by computing the change in the response variable from its predicted minimum value to its predicted maximum value, adjusted for the response’s maximum range (Grace and

Bollen 2005). All analyses was performed in R 2.15.3 (R Core Team 2014), using the package lm4 (Bates et al. 2012).

192

Chapter 7: Predicting species interactions

Table 7.1: Individual models for each node in the structural equation model. Trip was entered as a random factor to account for the repeated measures. Mean rainfall event size2 = Mean rainfall event size two trips prior.

Model Insectivorous dasyurids ~ Poisson Log(dasy) Site + Mean rainfall event size2 + mulgara + (~1|Trip) + offset(Trap nights) + ɛ

Spinifex cover ~ Binomial Logit(Spinifex cover) Years since wildfire + 8 month cumulative rainfall + (~1|Trip)+ ɛ

Spinifex seed ~ Binomial Logit(spinifex seed) Spinifex cover +8 month cumulative rainfall +(~1|Trip) +ɛ

Rodents ~ Poisson Log(rodents) Spinifex seed + Feral cat + Red fox + Dingo + (~1|Trip) +offset(Trap nights) +ɛ

Mulgara~ Poisson Log(Mulgara) Rodents + Spinifex cover + Years since wildfire + (~1|Trip) + offset(Trap nights) +ɛ

Feral cat ~ Poisson Log(Feral cat) Rodents + Dingo + (~1|Trip) +offset(Camera nights) +ɛ

Red fox~ Poisson Log(Red fox) Rodents + Dingo + (~1|Trip) +offset(Camera nights) + ɛ

Dingo~ Poisson Log(Dingo) Rodents + (~1|Trip) +offset(Camera nights) + ɛ

Reptiles~ Poisson Log(Reptiles) Years since wildfire + Spinifex cover + Rain days + (~1| Trip) +offset(Trap nights) + ɛ

Model justification

The proposed causal pathways between variables in the conceptual model (Fig. 7.2a) were determined by a priori knowledge. Both bottom-up and top-down effects can operate in

193

Chapter 7: Predicting species interactions

this system (Dickman et al. 1999b, Letnic and Dickman 2010, Greenville et al. 2014, Chapter

6) and thus are incorporated into the model. Rainfall, wildfire and spatial driver variables were chosen based on Chapters 4 and 5. Captures across sites were pooled for rodents, mulgara and reptiles as population synchrony is exhibited for these groups, but not for smaller, insectivorous dasyurids (Chapters 4 and 5). It was assumed that changes in spinifex cover could influence spinifex seed production, as larger plants (greater cover) could compete more successfully for resources and produce more successful seeding events. Rodents consume large amounts of spinifex seed in, proportion to their total diet, and seeding events are positively associated with population increases (Murray and Dickman 1994, Ricci 2003,

Chapter 4). Increases in rodent populations can lead to increases in predator populations, but predation pressure on rodents may differ between native and non-native predators (Pavey et al. 2008). For example, in Chapter 4 I found no predatory effects of the native mulgara on rodent populations, but feral cats and the introduced red fox can force populations to extinction (Dickman 1996). Australia’s top mammalian predator, the dingo, can suppress populations of the mesopredators, the feral cat and red fox, conferring some advantage to small mammal populations by reducing predation on them by meso-predators (Letnic et al.

2009, Letnic et al. 2011a, Greenville et al. 2014, Chapter 6). Reptile populations can be influenced by spinifex cover, wildfire and number of rain days (Dickman et al. 1999a, Letnic et al. 2004, Chapter 5). Small dasyurid populations differ across space and can be influenced by both bottom-up effects from rainfall and top-down effects from the predatory mulgara

(Dickman 2003, Haythornthwaite and Dickman 2006, Chapter 4). Chapter 4 showed that wildfire, spinifex cover and rodent captures are important drivers for mulgara populations.

194

Chapter 7: Predicting species interactions

Changes in rainfall and wildfire from climate change

To determine how the two drivers, rainfall and wildfire, may influence interactions

between trophic groups under a changing climate, we simulated the projected change in years

since wildfire and cumulative 8 month rainfall over the next 100 years. We generated a

current rainfall scenario for 2000 samples from the negative binomial distribution, using

parameters estimated from the actual 8-month cumulative rainfall data (µ = 152, size = 1.57, n = 2000). Extreme rainfall events (>95th quantile) have increased in frequency and

magnitude over the last ~100 years for central Australia, leading to an approximate increase

of 4 mm per year (Greenville et al. 2012, Chapter 2). Thus we generated 8-month cumulative rainfall datasets using a negative binomial distribution to project an increase in rainfall at 12 time steps over the next 100 years (100 years: µ = 419, n = 2000; see Table A4.3). The negative binomial distribution was chosen over the Poisson as rainfall in arid environments is over-dispersed due to large numbers of nil or small rainfall events. Histograms were used to confirm that simulated datasets had similar distributions to the actual rainfall datasets (Fig.

A5.1).

Wildfire return intervals are predicted to decrease due to changes in rainfall patterns and evaporation rates for central Australia (Cary et al. 2012, Greenville et al. 2012, Chapter

2). To simulate a reduction in the years since wildfire across the next 100 years, we used a negative binomial distribution. As above, we generated a current wildfire scenario based on parameter estimates from the actual wildfire data (µ= 21, size =3.34, n = 2000). We assumed that in 100 years the mean time between wildfires would change from 21 years to 10 years

(µ= 10, size = 3.34, n = 2000) and generated 12 time steps over the next 100 years (see Table

A4.3). All mean years since wildfire parameters used are within reported fire return intervals for central Australia (Allan and Southgate 2002, Nano et al. 2012). As for rainfall, histograms

195

Chapter 7: Predicting species interactions

were used to confirm that simulated datasets had similar distributions to the actual wildfire

datasets (Fig. A5.2).

The simulated datasets were entered into the SEM (Fig. 7.1) and predicted values for

each trophic level were re-entered at each node from spinifex cover, spinifex seed and finally

to rodents. In addition, to test the influence of top-down effects from introduced predators,

models were run with all mammalian predators, without the introduced red fox and feral cat

and without all mammalian predators. Models were run 2000 times, except for the rodent

models whose datasets was shorter due to predator monitoring using remote camera traps

only running for two years (n = 342).

Results

Structural equation modelling

In contrast to our a priori SEM model (Fig. 7.2a), three pathways were not

significant; spinifex cover was not associated with reptile captures, mulgara captures were

not associated with rodent numbers and the spinifex seed index was not associated with 8-

months cumulative rainfall (Fig. 7.2b). In accordance with our a priori model, spinifex cover

was positively associated with increases in 8 months cumulative rainfall and the spinifex seed

index increased with increasing cover (Fig. 7.2b). Rodent captures increased with increases in spinifex seed index and were negatively associated with the introduced red fox and feral cat, but were positively associated with the native dingo (Fig. 7.2b). Insectivorous dasyurids were

driven by both spatial and bottom-up effects from mean rainfall event size 2 months prior, but

also limited by top-down effects from the predatory mulgara (Fig. 7.2b). As predicted from

our a prior model, the mulgara was associated positively with spinifex cover. Reptile

196

Chapter 7: Predicting species interactions

captures were associated positively with the number of rain days and associated negatively with years since wildfire (Fig. 7.2b).

Figure 7.2: (a) Conceptual model describing the proposed interactions between drivers and trophic levels in the Simpson Desert, Australia. (b) Results from structural equation model. Values above drawings are percentage deviance explained. Standardised path coefficients are shown for each arrow. Arrow thickness proportional to effect size and represent significant path coefficients (P<0.05). Artwork by A. Foster.

197

Chapter 7: Predicting species interactions

Climate change

There were decreases in projected mean spinifex cover (Fig. 7.3a) and seed index

(Fig. 7.3b) over the next 100 years, but not for projected rodent captures (Fig. 7.3c). When feral predators were removed from the models there was an increase in projected rodent captures by approximately 9%. However, when both native and non-native mammalian predators were removed rodent captures increased by approximately 3% (Fig. 7.3c).

Figure 7.3: Projected results (±SE) for changes in (a) spinifex cover (b) spinifex seed index, and (c) rodent captures with all mammalian predators present, introduced predators (cats and foxes) removed and all mammalian predators removed, under a changing rainfall and wildfire scenario.

198

Chapter 7: Predicting species interactions

Discussion

As predicted from our a priori trophic interaction model, higher rainfall significantly increased captures of small mammal and reptiles, either directly or indirectly through changes in food resources. However, these flow-on effects from higher rainfall differed between

rodents and dasyurids. Rodent captures appeared to be limited by top-down effects from

introduced predators in contrast to the dasyurid populations which exhibited a response to the

mulgara, a native dasyurid predator. The dominant vegetation cover, spinifex, increased with

years since wildfire and rainfall, which in turn, seed production. Seed production increased

with greater spinifex cover, but increases in spinifex cover can provide fuel for wildfires

(Nano et al. 2012). Thus wildfire can indirectly change rodent populations through changes

in food availability. As a group, reptile numbers were influenced by wildfire and number of

rain days, but surprisingly not by spinifex cover. Wildfires were shown to have a direct effect

on reptiles, but not an indirect effect via spinifex cover. The grouping of reptile above the

species level may be masking more subtle processes and interspecific differences in

responses such as those found in Chapter 5.

Prey (rodent) populations were important in driving the populations all mammalian predators, except the mulgara. A positive pathway from dingoes to rodents suggests a facilitation interaction. As a top-predator in Australia, dingoes can suppress the introduced mesopredators, the feral cat and red fox and thereby reduce predation pressure on small prey species (Ritchie and Johnson 2009, Letnic et al. 2011a, Colman et al. 2014, Greenville et al.

2014, Chapter 6). However, we predicted that there would be a negative pathway from the dingo to the smaller introduced predators, but in contrast we found positive pathways from dingo to the mesopredators. Fleming et al. (2012) argued that the evidence for the dingoes suppressing mesopredators and thus having a positive effect of prey species was inconclusive

199

Chapter 7: Predicting species interactions

(cf. Johnson and Ritchie 2013). Alternatively, and more likely, is that all predators increase

together with increases in prey during rodent irruptions (Letnic et al. 2011b, Greenville et al.

2014, Chapter 6). During rodent irruptions the suppression of cats and foxes by dingoes

weakens compared with dry times, allowing increases in the populations of all mammalian

predators in central Australia (Greenville et al. 2014, Chapter 6).

Climate change

Simultaneous analyses of multiple drivers highlighted the interactive effect of changes

in population drivers from climate change. Contrary to our prediction, an increase in

vegetation cover from an increase in rainfall predicted for central Australia was offset by

increases in wildfire frequency, leading to a mean decrease in projected spinifex cover over

the next 100 years. A decrease in spinifex cover suggests a shift to smaller and younger

plants and, in turn, a reduction in reproductive output. Abrupt climate change can lead to

shifts in the vegetation community. For example, climate change records from charcoal and

pollen samples have shown that rapid changes in climate can increase fire activity and shifts

in the vegetation community (Lynch et al. 2007, Marlon et al. 2009). Using simulations to

predict vegetation change from anthropogenic climate change, Mouillot et al. (2002) found that a changing climate decreased the time between wildfires and a shift in vegetation to shrub-dominated landscapes. After wildfire in spinifex grasslands there is an increase in cover of annual and perennial vegetation (Nano et al. 2012). Given that spinifex grasslands occupy one quarter of the area of continental Australia (Allan and Southgate 2002), changes in rainfall and fire frequency may lead to a large-scale shift in the dominant vegetation across central and northern arid regions of Australia. For example, Wardle et al. (2015) projected

200

Chapter 7: Predicting species interactions

that changes in climate over the next 50 years will see a decline of suitable habitat for Acacia

georginae woodlands across central Australia.

There was no change in the projected mean number of rodent captures with decreases

in spinifex seeding under our climate change scenario. There are three possible explanations.

Firstly, rodents may be able to supplement their food with seed from other plant species, or

invertebrates. For example, Murray and Dickman (1994) found some dietary switching,

especially in the Austral autumn when invertebrates could make up to 60% of the diets of

desert rodents. Secondly, during resource pulses the increase in spinifex seed may be in

excess of what is required for the rodent populations to irrupt. For example, rodent

populations in central Australia respond once a rainfall threshold is reached and once over

this threshold there could be an excess of food (Greenville et al. 2012, 2013, Chapters 2 and

3). However, given their fast reproductive rates and rapid population responses to rainfall events, rodents could be expected to maximise their response to the available food supply. In addition, moderate to strong density dependence and rapid population declines after resource pulses operate in this system, suggesting that competition for food may be operating

(Dickman et al. 1999b, Chapter 4). Lastly, rodent populations may be limited by predation from introduced predators. Introduced predators can depredate rodents and other native mammals more heavily than native predators (Salo et al. 2007) and are one of the drivers of extinction of Australian mammals (Johnson 2006, Pavey et al. 2008). Furthermore and in line with our prediction, when mammalian predators were removed from our model there was an increase in predicted captures of rodents and this was greatest when only introduced predators were removed.

The SEM framework was useful for predicting direct and indirect interactions with existing species (Grace et al. 2012, Pasanen-Mortensen et al. 2013, Gordon et al. 2015), but this framework does have some limitations. Firstly, the SEM framework does not account for

201

Chapter 7: Predicting species interactions

novel interactions. For example, invasive species such as European rabbit (Oryctolagus cuniculus) and house mouse (Mus musculus) are largely absent in our study region, but increases in the magnitude and frequency of large (>90th quantile) rainfall events may allow these species to establish (Letnic and Dickman 2006, Greenville et al. 2012, Chapter 2). The influence of these introduced species on vegetation and predator populations should be the focus of future research. Secondly, weak interactions may not be well represented in this model, but they can have important effects for the overall stability of food webs (McCann et al. 1998, Tylianakis et al. 2008). Lastly, little is known about other important components of the community, such as nutrient cycling, invertebrates, birds and interactions between reptile, dasyurid and eutherian predators (e.g. Tischler et al. 2013, Travers and Eldridge 2013, Read and Scoleri 2014). We suggest that these areas be foci for further research.

Conclusion

Our results show that simultaneous analysis of multiple external population drivers, was important to uncover and fully appreciate how interactions are structured between species and between trophic groups. The trophic interaction modelling framework used here was useful for bringing together the results from multiple studies. We found that population responses were influenced by both bottom-up and top-down regulation, highlighting the need for models to incorporate both external population drivers and biotic interactions when predicting changes in species’ populations.

202

Chapter 7: Predicting species interactions

Acknowledgements

We thank Bush Heritage Australia, H. Jukes, G. McDonald, D. Smith and G. Woods

for allowing access to the properties in the study region, members of the Desert Ecology

Research Group, especially B. Tamayo, C.-L. Beh and D. Nelson, and many volunteers for

valuable assistance in the field. Special thanks to A. Kwok for discussions on early themes of

this study and to A. Foster for the artwork. Funding was provided by the Australian Research

Council and the Australian Government’s Terrestrial Ecosystems Research Network

(www.tern.gov.au), an Australian research infrastructure facility established under the

National Collaborative Research Infrastructure Strategy and Education Infrastructure Fund -

Super Science Initiative through the Department of Industry, Innovation, Science, Research

and Tertiary Education. ACG was supported by an Australian Postgraduate Award, and CRD

by an Australian Research Council Fellowship.

203

Chapter 7: Predicting species interactions

References:

Allan, G. and R. Southgate. 2002. Fire regimes in the spinifex landscapes of Australia, Pages 145-176 in R. A. Bradstock, J. E. Williams, and A. M. Gill, editors. Flammable Australia: the fire regimes and biodiversity of a continent. Cambridge University Press, Cambridge. Baskett, M. L. and A. K. Salomon. 2010. Recruitment facilitation can drive alternative states on temperate reefs. Ecology 91: 1763-1773. Bates, D., M. Maechler, and B. Bolker. 2012. lme4: linear mixed-effects models using S4 classes. R package version 0.999999-0. http://CRAN.R-project.org/package=lme4. Bruno, J. F., J. J. Stachowicz, and M. D. Bertness. 2003. Inclusion of facilitation into ecological theory. Trends in Ecology & Evolution 18: 119-125. Bureau of Meteorology. 2012. Climate data online, accessed 20 June 2012, http://www.bom.gov.au/climate/data/ Cary, G. J., R. A. Bradstock, A. M. Gill, and R. J. Williams. 2012. Global change and fire regimes in Australia, Pages 149-169 in R. A. Bradstock, A. M. Gill, and R. J. Williams, editors. Flammable Australia: fire regimes, biodiversity and ecosystems in a changing world. CSIRO Publishing, Vic. Colman, N. J., C. E. Gordon, M. S. Crowther, and M. Letnic. 2014. Lethal control of an apex predator has unintended cascading effects on forest mammal assemblages. Proceedings of the Royal Society B: Biological Sciences 281: 20133094. Davidson, D. W., D. A. Samson, and R. S. Inouye. 1985. Granivory in the Chihuahuan Desert: interactions within and between trophic levels. Ecology 66: 486-502. Dickman, C. R. 1996. Impact of exotic generalist predators on the native fauna of Australia Wildlife Biology 2: 185-195. Dickman, C. R. 2003. Species interactions: indirect effects, Pages 158-170 in P. Attiwill and B. Wilson, editors. Ecology: an Australian perspective. Oxford University Press, Melbourne, Vic. Dickman, C. R., M. Letnic, and P. S. Mahon. 1999a. Population dynamics of two species of dragon lizards in arid Australia: the effects of rainfall. Oecologia 119: 357-366. Dickman, C. R., P. S. Mahon, P. Masters, and D. F. Gibson. 1999b. Long-term dynamics of rodent populations in arid Australia: the influence of rainfall. Wildlife Research 26: 389-403.

204

Chapter 7: Predicting species interactions

Elmhagen, B., G. Destouni, A. Angerbjörn, S. Borgström, E. Boyd, S. A. O. Cousins, L. Dalén, J. Ehrlén, M. Ermold, P. A. Hambäck, J. Hedlund, K. Hylander, F. Jaramillo, V. K. Lagerholm, S. W. Lyon, H. Moor, B. Nykvist, M. Pasanen-Mortensen, J. Plue, C. Prieto, Y. van der Velde, and R. Lindborg. 2015. Interacting effects of change in climate, human population, land use, and water use on biodiversity and ecosystem services. Ecology and Society 20: 23. Fleming, P. J. S., B. L. Allen, and G.-A. Ballard. 2012. Seven considerations about dingoes as biodiversity engineers: the socioecological niches of dogs in Australia. Australian Mammalogy 34: 119-131. Friend, G. R., G. T. Smith, D. S. Mitchell, and C. R. Dickman. 1989. Influence of pitfall and drift fence design on capture rates of small vertebrates in semi-arid habitats of Western Australia. Australian Wildlife Research 16: 1-10. Gordon, C. E., A. Feit, J. Gruber, and M. Letnic. 2015. Mesopredator suppression by an apex predator alleviates the risk of predation perceived by small prey. Proceedings of the Royal Society B: Biological Sciences 282: 20142870. Grace, J. B. and K. A. Bollen. 2005. Interpreting the results from multiple regression and structural equation models. Bulletin of the Ecological Society of America: 283-295. Grace, J. B., D. R. Schoolmaster, G. R. Guntenspergen, A. M. Little, B. R. Mitchell, K. M. Miller, and E. W. Schweiger. 2012. Guidelines for a graph-theoretic implementation of structural equation modeling. Ecosphere 3: art73. Greenville, A. C., C. R. Dickman, G. M. Wardle, and M. Letnic. 2009. The fire history of an arid grassland: the influence of antecedent rainfall and ENSO. International Journal of Wildland Fire 18: 631-639. Greenville, A. C., G. M. Wardle, and C. R. Dickman. 2012. Extreme climatic events drive mammal irruptions: regression analysis of 100-year trends in desert rainfall and temperature. Ecology and Evolution 2: 2645-2658. Greenville, A. C., G. M. Wardle, and C. R. Dickman. 2013. Extreme rainfall events predict irruptions of rat plagues in central Australia. Austral Ecology 38: 754-764. Greenville, A. C., G. M. Wardle, B. Tamayo, and C. R. Dickman. 2014. Bottom-up and top- down processes interact to modify intraguild interactions in resource-pulse environments. Oecologia 175: 1349-1358. Harrison, R. D. 2000. Repercussions of El Niño: drought causes extinction and the breakdown of mutualism in Borneo. Proceedings of the Royal Society of London. Series B: Biological Sciences 267: 911-915. 205

Chapter 7: Predicting species interactions

Harrison, R. D. 2001. Drought and the consequences of El Niño in Borneo: a case study of figs. Population Ecology 43: 63-75. Haythornthwaite, A. S. and C. R. Dickman. 2006. Distribution, abundance, and individual strategies: a multi-scale analysis of dasyurid marsupials in arid central Australia. Ecography 29: 285-300. Heithaus, M. R., A. Frid, A. J. Wirsing, and B. Worm. 2008. Predicting ecological consequences of marine top predator declines. Trends in Ecology & Evolution 23: 202-210. Johnson, C. 2006. Australia's mammal extinctions: a 50,000 year history. Cambridge University Press, Port Melbourne, Vic. Johnson, C. N. and E. G. Ritchie. 2013. The dingo and biodiversity conservation: response to Fleming et al. (2012). Australian Mammalogy 35: 8-14. Lavergne, S., N. Mouquet, W. Thuiller, and O. Ronce. 2010. Biodiversity and climate change: integrating evolutionary and ecological responses of species and communities. Annual Review of Ecology, Evolution, and Systematics 41: 321-350. Le Page, Y., J. M. C. Pereira, R. Trigo, C. da Camara, D. Oom, and B. Mota. 2008. Global fire activity patterns (1996-2006) and climatic influence: an analysis using the World Fire Atlas. Atmospheric Chemistry Physics 8: 1911-1924. Letnic, M., M. S. Crowther, and F. Koch. 2009. Does a top-predator provide an endangered rodent with refuge from an invasive mesopredator? Animal Conservation 12: 302- 312. Letnic, M. and C. R. Dickman. 2006. Boom means bust: interactions between the El Niño/Southern Oscillation (ENSO), rainfall and the processes threatening mammal species in arid Australia. Biodiversity and Conservation 15: 3847-3880. Letnic, M. and C. R. Dickman. 2010. Resource pulses and mammalian dynamics: conceptual models for hummock grasslands and other Australian desert habitats. Biological Reviews 85: 501-521. Letnic, M., C. R. Dickman, M. K. Tischler, B. Tamayo, and C. L. Beh. 2004. The responses of small mammals and lizards to post-fire succession and rainfall in arid Australia. Journal of Arid Environments 59: 85-114. Letnic, M., A. Greenville, E. Denny, C. R. Dickman, M. Tischler, C. Gordon, and F. Koch. 2011a. Does a top predator suppress the abundance of an invasive mesopredator at a continental scale? Global Ecology and Biogeography 20: 343-353.

206

Chapter 7: Predicting species interactions

Letnic, M., P. Story, G. Story, J. Field, O. Brown, and C. R. Dickman. 2011b. Resource pulses, switching trophic control, and the dynamics of small mammal assemblages in arid Australia. Journal of Mammalogy 92: 1210-1222. Lynch, A. H., J. Beringer, P. Kershaw, A. Marshall, S. Mooney, N. Tapper, C. Turney, and S. Van Der Kaars. 2007. Using the paleorecord to evaluate climate and fire interactions in Australia. Annual Review of Earth and Planetary Sciences 35: 215-239. Marlon, J. R., P. J. Bartlein, M. K. Walsh, S. P. Harrison, K. J. Brown, M. E. Edwards, P. E. Higuera, M. J. Power, R. S. Anderson, C. Briles, A. Brunelle, C. Carcaillet, M. Daniels, F. S. Hu, M. Lavoie, C. Long, T. Minckley, P. J. H. Richard, A. C. Scott, D. S. Shafer, W. Tinner, C. E. Umbanhowar, and C. Whitlock. 2009. Wildfire responses to abrupt climate change in North America. Proceedings of the National Academy of Sciences 106: 2519-2524. McCann, K., A. Hastings, and G. R. Huxel. 1998. Weak trophic interactions and the balance of nature. Nature 395: 794-798. Memmott, J., P. G. Craze, N. M. Waser, and M. V. Price. 2007. Global warming and the disruption of plant–pollinator interactions. Ecology Letters 10: 710-717. Mouillot, F., S. Rambal, and R. Joffre. 2002. Simulating climate change impacts on fire frequency and vegetation dynamics in a Mediterranean-type ecosystem. Global Change Biology 8: 423-437. Murray, B. R. and C. R. Dickman. 1994. Granivory and microhabitat use in Australian desert rodents: are seeds important? Oecologia 99: 216-225. NAFI. 2013. North Australian Fire Information, accessed 12th December 2012, http://www.firenorth.org.au/nafi2/ Nano, C. E., P. Clarke, and C. R. Pavey. 2012. Fire regimes in arid hummock grasslands and Acacia shrublands, Pages 195-214 in R. A. Bradstock, A. M. Gill, and R. J. Williams, editors. Flammable Australia: fire regimes, biodiversity and ecosystems in a changing world. CSIRO Publishing, Vic. Pasanen-Mortensen, M., M. Pyykönen, and B. Elmhagen. 2013. Where lynx prevail, foxes will fail – limitation of a mesopredator in Eurasia. Global Ecology and Biogeography 22: 868-877. Pavey, C. R., S. R. Eldridge, and M. Heywood. 2008. Population dynamics and prey selection of native and introduced predators during a rodent outbreak in arid Australia. Journal of Mammalogy 89: 674-683.

207

Chapter 7: Predicting species interactions

Popic, T. J. and G. M. Wardle. 2012. Extremes: understanding flower – visitor interactions in a changing climate., Pages 99-106 in D. Lunney and P. Hutchings, editors. Wildlife and climate change: towards robust conservation strategies for Australian fauna. Royal Zoological Society of New South Wales, Mosman, NSW, Australia. Purdie, J. L. 1984. Land systems of the Simpson Desert region. Natural resources series no. 2. CSIRO Division of Water and Land Resources, Melbourne, Australia. R Core Team. 2014. R: a language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria. Read, J. L. and V. Scoleri. 2014. Ecological implications of reptile mesopredator release in arid South Australia. Journal of Herpetology 49: 64-69. Ricci, S. 2003. Population dynamics and trophic ecology of two species of Australian desert rodents. PhD Thesis. University of Sydney, Sydney. Ripple, W. J., J. A. Estes, R. L. Beschta, C. C. Wilmers, E. G. Ritchie, M. Hebblewhite, J. Berger, B. Elmhagen, M. Letnic, M. P. Nelson, O. J. Schmitz, D. W. Smith, A. D. Wallach, and A. J. Wirsing. 2014. Status and ecological effects of the world's largest carnivores. Science 343: 1241484. Ritchie, E. G. and C. N. Johnson. 2009. Predator interactions, mesopredator release and biodiversity conservation. Ecology Letters 12: 982-998. Ross, B. E., M. B. Hooten, J.-M. DeVink, and D. N. Koons. 2015. Combined effects of climate, predation, and density dependence on scaup population dynamics. Ecological Applications. Salo, P., E. Korpimäki, P. B. Banks, M. Nordström, and C. R. Dickman. 2007. Alien predators are more dangerous than native predators to prey populations. Proceedings of the Royal Society B: Biological Sciences 274: 1237-1243. Schweiger, O., J. C. Biesmeijer, R. Bommarco, T. Hickler, P. E. Hulme, S. Klotz, I. Kühn, M. Moora, A. Nielsen, R. Ohlemüller, T. Petanidou, S. G. Potts, P. Pyšek, J. C. Stout, M. T. Sykes, T. Tscheulin, M. Vilà, G.-R. Walther, C. Westphal, M. Winter, M. Zobel, and J. Settele. 2010. Multiple stressors on biotic interactions: how climate change and alien species interact to affect pollination. Biological Reviews 85: 777- 795. Shephard, M. 1992. The Simpson Desert: natural history and human endeavour. Royal Geographical Society of Australasia, Adelaide, Australia.

208

Chapter 7: Predicting species interactions

Shipley, B. 2000. Cause and correlation in biology : a user's guide to path analysis, structural equations and causal inference. Cambridge University Press, Cambridge, U.K. ; New York. Suttle, K. B., M. A. Thomsen, and M. E. Power. 2007. Species Interactions Reverse Grassland Responses to Changing Climate. Science 315: 640-642. Tischler, M., C. R. Dickman, and G. M. Wardle. 2013. Avian functional group responses to rainfall across four vegetation types in the Simpson Desert, central Australia. Austral Ecology 38: 809-819. Traill, L. W., M. L. M. Lim, N. S. Sodhi, and C. J. A. Bradshaw. 2010. Mechanisms driving change: altered species interactions and ecosystem function through global warming. Journal of Animal Ecology 79: 937-947. Travers, S. K. and D. J. Eldridge. 2013. Increased rainfall frequency triggers an increase in litter fall rates of reproductive structures in an arid eucalypt woodland. Austral Ecology 38: 820–830. Tylianakis, J. M., R. K. Didham, J. Bascompte, and D. A. Wardle. 2008. Global change and species interactions in terrestrial ecosystems. Ecology Letters 11: 1351-1363. Wardle, G. M., A. C. Greenville, A. S. K. Frank, M. Tischer, N. J. Emery, and C. R. Dickman. 2015. Ecosystem risk assessment of Georgina gidgee woodlands in central Australia. Austral Ecology 40: 444-459.

209

210

Chapter 8: General discussion and conclusions

Chapter 8: General Discussion and Conclusions

The Milky Way stretches over a small group of trees that is camp at our long-term research site in the Simpson Desert, Australia. Photograph by Aaron Greenville.

211

Chapter 8: General discussion and conclusions

Overview

Species persist ultimately because individuals cope with environmental demands and

pressures from other species placed upon them as they try to survive and reproduce. As long

as their populations are stable over the long-term, this variability is part and parcel of the

dynamic of life. The understanding that advances our knowledge is in realising that not all

environments are easy, or predictable, and not all species cope in the same way. Stemming

from thinking in the first half of the twentieth century, debates in the 1950s and 1960s were

an important stage in shaping our current understanding of the role of biotic and abiotic

factors in driving the dynamics of species’ populations (e.g. Andrewartha and Birch 1954,

Nicholson 1954, MacArthur 1958, Lack 1966, 1967, Chapter 1). It is now acknowledged that

both factors—in their many forms—are important in driving species’ populations, and many examples of their joint effects can be found in both terrestrial and marine systems (Kolar and

Rahel 1993, Vucetich and Peterson 2004). It is also becoming increasingly recognised that many, if not most, species interactions are themselves climate-dependent (Tylianakis et al.

2008), further emphasising the combined or interactive importance of these factors. Without taking into account biotic interactions, predictions from abiotic factors, such as climate change are likely to be imprecise (Van der Putten et al. 2010).

The overall objective of this thesis was to incorporate both biotic and abiotic interactions into one conceptual framework to make more accurate predictions of how species’ populations change across space and time (Fig. 1.1). To help progress this objective,

I built a trophic interaction model to capture the complexity of biotic and abiotic interactions in a model arid system. The model was built upon smaller components that investigated the spatial and temporal dynamics of species’ populations and their biotic and abiotic drivers. I

212

Chapter 8: General discussion and conclusions

now summarise the main achievements of this study (Table 8.1), discuss its limitations, and then suggest avenues for future research.

Table 8.1: Summary of the key findings and achievements of this thesis, and the research area they advance.

Key findings Research area Advancement Chapter Non-linear interactions Population biology Populations may not change 2 & 3 until an event reaches a threshold level

Biotic and abiotic Spatial population Demonstration of Moran’s 4 & 5 interactions influence biology Theorem population synchrony

Density dependence Population biology Density dependence can 4 & 5 operates in some species operate in resource-pulse environments, but not ubiquitously

Timing of interactions and Trophic ecology Conceptual framework 6 their strengths are context- proposed to investigate and specific explain the relative importance of both bottom-up and top-down effects

Simultaneous analysis of Population and Communities are driven by 7 population drivers is conservation biology multiple spatial and temporal required to understand biotic and abiotic interactions population dynamics

Achievements

“Ecology: it’s not rocket science, it’s far more complicated” Anon

A key achievement of this thesis is that it confirms the importance and operation of both biotic and abiotic interactions in resource-pulse environments (Table 8.1). The trophic

213

Chapter 8: General discussion and conclusions

interactions model highlights how interactions are context-specific and can change over space and time. It incorporates both bottom-up and top-down processes, and shows that the relative strength of these processes can change (Fig. 8.1). I show also that simultaneous analysis of abiotic factors is important in predicting changes in species’ populations, and that biotic interactions (e.g. predation) can limit increases in consumer populations (Chapter 7; Fig. 8.1).

The processes contributing to this model arid system are illustrated below (Fig. 8.1).

Figure 8.1: Summary of thesis findings on biotic and abiotic interactions in an arid resource- pulse environment. This trophic interaction model highlights how interactions are context- specific and can change over space and time. It incorporates both bottom-up and top-down processes, but the strength of these processes can switch. Grey arrows are possible future interactions to explore. Artwork by A. Foster.

214

Chapter 8: General discussion and conclusions

Interactions are not always linear; species’ populations may not change until a

threshold is met. For example, extreme rainfall events (>95th quantile or >400 mm) were

required for rodent populations to increase, resulting in a boom in population numbers

(Greenville et al. 2012, 2013, Chapters 2 & 3). However, not all small mammals respond to the same drivers, as dasyurid marsupial captures were not correlated with extreme rainfall events (Greenville et al. 2012, Chapter 2). Non-linear interactions are becoming increasing important to understand, as extreme weather events, such as tropical cyclones, heat waves and flooding rains, are increasing in magnitude and frequency (IPCC 2007, 2014). Increases

in extreme weather events may change species’ populations through increases in mortality

rates, decreases in reproduction, and by facilitation of invasive species or novel interactions

(Welbergen et al. 2008, Smith 2011, Chapter 2). In addition, extreme weather events, such as

flooding rains, are important not only for driving resource pulses in arid systems, but also

other abiotic events such as wildfire (Letnic and Dickman 2006, Greenville et al. 2009).

Biotic and abiotic interactions together influence population synchrony. Spatially

separated sub-populations can act in synchrony due to dispersal or predation (Bjørnstad et al.

1999, Lande et al. 1999, Ims and Andreassen 2000). However, in the absence of dispersal, a widespread abiotic factor can cause population synchrony (Moran 1953, Bjørnstad et al.

1999). Two groups of small mammals, rodents and dasyurids, showed contrasting levels of synchrony in this study (Chapter 4). Rodent populations exhibited high levels of population synchrony across widely-spaced sites (>20 km), with rare flooding rainfall events being the likely driver (Chapter 4). Reptile species showed similar patterns to the rodents and had high levels of population synchrony (Chapter 5). Shared productivity and wildfire history were among the possible drivers of reptilian population synchrony (Chapter 5), suggesting that the

Moran effect could operate for rodent and reptile populations in the desert study system.

215

Chapter 8: General discussion and conclusions

Biotic factors, such as density dependence, are difficult to detect in resource-pulse

systems. In general, populations are low during the predominately dry (bust) periods

(Dickman et al. 1999, Chapters 4 & 5), limiting interactions among individuals and also

reducing the power to detect density dependence (Sinclair and Krebs 2002). During

population booms following flooding rainfall events, competition is limited due to the

abundance of food (Letnic et al. 2011). It is only for a brief period when populations start to

decline when density dependence would be likely to be most influential. Using a novel

statistical technique, multivariate autoregressive state-space (MARSS) modelling, this study

is the first to detect density dependence in some species of small mammals and reptiles in the

Australian desert system (Chapters 4 & 5). Density dependence has been demonstrated in

small mammals in other arid or semi-arid environments (e.g. Lima and Jaksic 1999, Previtali

et al. 2009), but these are probably not as strongly resource-pulsing as those in central

Australia (Letnic and Dickman 2010, Kelt and Meserve 2014).

The timing of interactions and their strengths is context-specific. The relative strength of bottom-up (productivity increases due to rainfall events) and top-down processes

(predation) can undergo marked changes, showing that both operate and are probably of great importance in the desert study system (Greenville et al. 2014, Chapter 6). Bottom-up and top- down processes are important in other terrestrial systems and it is becoming increasingly recognised that it is hard to classify a system as either top-down or bottom-up dominated as both will often operate in different contexts (Ritchie 2000, Sinclair et al. 2000, Vucetich and

Peterson 2004). Sinclair and Krebs (2002) proposed that food supply is the primary factor determining population growth rates and therefore that bottom-up effects are always important. This proposal received support from meta-analyses of supplementary feeding experiments in small mammals, which showed that most of the species studied were food- limited (Prevedello et al. 2013). However, the strength of bottom-up control can be modified

216

Chapter 8: General discussion and conclusions

due to top-down processes, social interactions within species and stochastic disturbances

(Sinclair and Krebs 2002, Krebs 2013). By using the conceptual framework proposed in

Chapter 6 (Fig. 6.2), abiotic disturbances from flooding rainfall events can increase the growth rates of meso-predator populations to levels where they can escape top-down regulation (Letnic et al. 2011, Greenville et al. 2014, Chapter 6). As resources decline, top- down effects increase again in strength (Letnic et al. 2011, Greenville et al. 2014, Chapter 6) and can help to drive rapid declines in population size.

Limitations and future research

This thesis brings together previous findings from experimental and observational studies, uses these to identify potentially important interactions and interaction pathways, and incorporates them in a trophic interactions model (Figs 1.1 & 8.1). By combining observational and experimental data, causal mechanisms that underpin biotic and abiotic interactions can be found (Eberhardt and Thomas 1991, Sagarin and Pauchard 2009). For example, to confirm and explain the observations of top-predator–mesopredator interactions, further experimental studies are required (Greenville et al. 2014, Chapter 6). The exact mechanism for suppression of mesopredators by dingoes is unknown. Possible processes could be direct predation, extreme interference competition (e.g. Moseby et al. 2012) or perhaps even exploitation competition. Experiments, such as the dingo-reintroduction experiments proposed recently by Newsome et al. (2015), will be invaluable in uncovering and quantifying the strength and nature of intra-guild predator interactions.

This thesis also found that density dependence is not ubiquitous (Chapters 4 and 5).

For example, small dasyurid marsupial populations in this desert system are generally low in dry times, presumably due to lack of prey (Dickman et al. 2001, Chapter 4). Population

217

Chapter 8: General discussion and conclusions

numbers of dasyurid marsupials do not respond to large rainfall events and in boom times,

flooding rains could destroy burrows and limit population responses (Dickman et al. 2001,

Chapter 4). Alternatively, small dasyurid marsupial populations may be limited during

resource booms by competition with irruptive rodent populations or direct predation, or fear

of predators, by the larger mulgara (Dasycercus blythi) (Dickman 2003, Chapters 4 and 7).

Thus populations of small dasyurids may not increase to levels high enough for density dependence to operate. By using new techniques to model species interactions (e.g. Chapter

7), studies into how both bottom-up and top-down effects influence species populations will become an increasingly exciting new field in population biology and trophic ecology.

The focus of much past research has concentrated on species that have regular

breeding systems and thus any variation at the intra-annual scale has been partitioned as background environmental noise (Turchin 1995). However, for species that are not restricted to a breeding season or exhibit large population fluctuations on time-scales of less than a

year, such as the rodent species in this study (Chapters 3 and 4), this variation can contain

important information about possible population drivers. Thus the temporal scale used to

investigate species’ population changes can be important, and the common usage of averaged

annual population numbers in many studies is perhaps why abiotic factors have been

perceived as being less important than biotic factors. The quality and length of long-term population time-series are increasing and new, more powerful autoregressive techniques are becoming available (e.g. Chapters 4 and 5), together, these will lead to further increases in our understanding about the factors that drive species’ populations.

The trophic interactions model does not account for all novel interactions. For example, invasive species such as European rabbit (Oryctolagus cuniculus) and house mouse

(Mus musculus) are largely absent in the desert study region, but increases in the magnitude and frequency of large (>90th quantile) rainfall events may allow these species to establish

218

Chapter 8: General discussion and conclusions

(Letnic and Dickman 2006, Greenville et al. 2012, Chapter 2). However, the modelling

framework allows novel interactions to be added by using prior information for adjacent

systems to parameterise the model. This would be a useful avenue for future research into the

impact of invasive species.

Lastly, little is known about other important components and processes in the community, such as nutrient cycling, invertebrates and interactions between reptile, dasyurid

and eutherian predators (Fig. 8.1). I suggest that these areas be foci for further research.

Concluding remarks

Long-term studies offer a unique chance to study biotic and abiotic interactions,

including interactions between trophic groups. This is especially true when abiotic factors,

such as flooding rainfall events are rare and also when interactions can vary both temporally

and spatially (Greenville et al. 2012, Chapters 2 and 4). The responses of species to abiotic changes, such as climate change, and the effects of species interactions on their ecosystems are often complex, nonlinear, and difficult to predict on the basis short-term studies (Brown

et al. 2001). Hence, it is imperative that long-term studies are encouraged to fully elucidate

the complexity of ecological systems and provide solutions to mitigate biodiversity loss

(Lindenmayer et al. 2012). The trophic interactions model introduced in Chapter one (Fig.

1.1) and summarised in this chapter (Fig. 8.1) was a successful framework to explore species’

spatio-temporal biotic and abiotic interactions. This framework highlights the complexity of

ecological relationships and their dynamic nature over time and space.

However, the trophic interactions model is also a model that reflects the current

situation with extant species. While the study region has a relatively complete small mammal

219

Chapter 8: General discussion and conclusions

and reptile fauna, most of the native medium-sized mammals are now gone (Johnson 2006,

Letnic and Dickman 2006, Woinarski et al. 2014). This is a landscape whose unique nocturnal nightlife has been forever silenced by extinction. The burrows, diggings, foraging modes and other ecosystem services provided by medium-sized mammals (Read et al. 2008) were removed from this highly complex environment with the advent of European settlement and the introduction of invasive species. The system is still-complex, but now much simplified compared to how it may have functioned a century or more ago. Indeed,

Finlayson’s (1935) pioneering exploration and natural history works of central Australia forewarned “that our wool-clips, and beef and timber trades have been dearly bought”. Future attempts to model the arid system could attempt to include these now-extinct species, perhaps informed by their effects elsewhere or via reintroduction trials.

This thesis has attempted to take advantage of over 20 years of ecological research focused on the ecology of central Australia (Dickman et al. 2014). Of this history, I have been fortunate enough to have been part of the last 15 years of the endeavour, working in various capacities. Even though the hand of extinction has been brought down most strongly in Australia’s arid regions (Burbidge and McKenzie 1989, Smith and Quin 1996, McKenzie et al. 2007), studies such as this elucidate the complexity of the current ecology of central

Australia and the sense of wonder that this environment conveys. This wonder feeds curiosity and surely curiosity, the most noble of human traits, needs to be conserved.

Acknowledgements

Thank you to Chris Dickman, Glenda Wardle and Alan Kwok for providing constructive criticism and comments on earlier versions of this chapter. Funding was

220

Chapter 8: General discussion and conclusions

provided by the Australian Research Council, an Australian Postgraduate Award and the

Australian Government’s Terrestrial Ecosystems Research Network (www.tern.gov.au).

221

Chapter 8: General discussion and conclusions

References:

Andrewartha, H. G. and L. C. Birch. 1954. The distribution and abundance of animals. University of Chicago Press, Chicago, USA. Bjørnstad, O. N., R. A. Ims, and X. Lambin. 1999. Spatial population dynamics: analyzing patterns and processes of population synchrony. Trends in Ecology & Evolution 14: 427-432. Brown, J. H., T. G. Whitham, S. K. Morgan Ernest, and C. A. Gehring. 2001. Complex species interactions and the dynamics of ecological systems: long-term experiments. Science 293: 643-650. Burbidge, A. A. and N. L. McKenzie. 1989. Patterns in the modern decline of western Australia's vertebrate fauna: Causes and conservation implications. Biological Conservation 50: 143-198. Dickman , C., G. M. Wardle, J. Foulkes, and N. de Preu. 2014. Desert complex environments, Pages 379-438 in D. Lindenmayer, E. Burns, N. Y. Thurgate, and A. Lowe, editors. Biodiversity and environmental change: monitoring, challenges and direction. CSIRO Publishing, Vic. Dickman, C. R. 2003. Species interactions: indirect effects, Pages 158-170 in P. Attiwill and B. Wilson, editors. Ecology: an Australian perspective. Oxford University Press, Melbourne, Vic. Dickman, C. R., A. S. Haythornthwaite, G. H. McNaught, P. S. Mahon, B. Tamayo, and M. Letnic. 2001. Population dynamics of three species of dasyurid marsupials in arid central Australia: a 10-year study. Wildlife Research 28: 493-506. Dickman, C. R., P. S. Mahon, P. Masters, and D. F. Gibson. 1999. Long-term dynamics of rodent populations in arid Australia: the influence of rainfall. Wildlife Research 26: 389-403. Eberhardt, L. L. and J. M. Thomas. 1991. Designing environmental field studies. Ecological Monographs 61: 53-73. Finlayson, H. H. 1935. Red Centre: man and beast in the heart of Australia. Angus and Robertson, Sydney. Greenville, A. C., C. R. Dickman, G. M. Wardle, and M. Letnic. 2009. The fire history of an arid grassland: the influence of antecedent rainfall and ENSO. International Journal of Wildland Fire 18: 631-639.

222

Chapter 8: General discussion and conclusions

Greenville, A. C., G. M. Wardle, and C. R. Dickman. 2012. Extreme climatic events drive mammal irruptions: regression analysis of 100-year trends in desert rainfall and temperature. Ecology and Evolution 2: 2645-2658. Greenville, A. C., G. M. Wardle, and C. R. Dickman. 2013. Extreme rainfall events predict irruptions of rat plagues in central Australia. Austral Ecology 38: 754-764. Greenville, A. C., G. M. Wardle, B. Tamayo, and C. R. Dickman. 2014. Bottom-up and top- down processes interact to modify intraguild interactions in resource-pulse environments. Oecologia 175: 1349-1358. Ims, R. A. and H. P. Andreassen. 2000. Spatial synchronization of vole population dynamics by predatory birds. Nature 408: 194-196. IPCC. 2007. Climate change 2007: Synthesis report. Contribution of Working Groups I, II and III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change, Geneva. IPCC. 2014. Climate Change 2014: Synthesis Report. Contribution of Working Groups I, II and III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Core Writing Team, R.K. Pachauri and L.A. Meyer (eds.)]. IPCC, Geneva, Switzerland. Johnson, C. 2006. Australia's mammal extinctions: a 50,000 year history. Cambridge University Press, Port Melbourne, Vic. Kelt, D. A. and P. L. Meserve. 2014. Status and challenges for conservation of small mammal assemblages in South America. Biological reviews 89: 705-722. Kolar, C. and F. Rahel. 1993. Interaction of a biotic factor (predator presence) and an abiotic factor (low oxygen) as an influence on benthic invertebrate communities. Oecologia 95: 210-219. Krebs, C. J. 2013. Population fluctuations in rodents. University of Chicago Press, Chicago. Lack, D. 1966. Population studies of birds. Clarendon Press, Oxford,. Lack, D. 1967. The natural regulation of animal numbers. 1st edition. Clarendon Press, Oxford,. Lande, R., S. Engen, and B. E. Sæther. 1999. Spatial scale of population synchrony: environmental correlation versus dispersal and density regulation. The American Naturalist 154: 271-281. Letnic, M. and C. R. Dickman. 2006. Boom means bust: interactions between the El Niño/Southern Oscillation (ENSO), rainfall and the processes threatening mammal species in arid Australia. Biodiversity and Conservation 15: 3847-3880. 223

Chapter 8: General discussion and conclusions

Letnic, M. and C. R. Dickman. 2010. Resource pulses and mammalian dynamics: conceptual models for hummock grasslands and other Australian desert habitats. Biological Reviews 85: 501-521. Letnic, M., P. Story, G. Story, J. Field, O. Brown, and C. R. Dickman. 2011. Resource pulses, switching trophic control, and the dynamics of small mammal assemblages in arid Australia. Journal of Mammalogy 92: 1210-1222. Lima, M. and F. M. Jaksic. 1999. Population dynamics of three Neotropical small mammals: Time series models and the role of delayed density-dependence in population irruptions. Australian Journal of Ecology 24: 25-34. Lindenmayer, D. B., G. E. Likens, A. Andersen, D. Bowman, C. M. Bull, E. Burns, C. R. Dickman, A. A. Hoffmann, D. A. Keith, M. J. Liddell, A. J. Lowe, D. J. Metcalfe, S. R. Phinn, J. Russell-Smith, N. Thurgate, and G. M. Wardle. 2012. Value of long-term ecological studies. Austral Ecology 37: 745-757. MacArthur, R. H. 1958. Population ecology of some warblers of northeastern coniferous forests. Ecology 39: 599-619. McKenzie, N. L., A. A. Burbidge, A. Baynes, R. N. Brereton, C. R. Dickman, G. Gordon, L. A. Gibson, P. W. Menkhorst, A. C. Robinson, M. R. Williams, and J. C. Z. Woinarski. 2007. Analysis of factors implicated in the recent decline of Australia's mammal fauna. Journal of Biogeography 34: 597-611. Moran, P. 1953. The statistical analysis of the Canadian lynx cycle. Australian Journal of Zoology 1: 291-298. Moseby, K. E., H. Neilly, J. L. Read, and H. A. Crisp. 2012. Interactions between a top order predator and exotic mesopredators in the Australian rangelands. International Journal of Ecology Article 250352. Newsome, T. M., G.-A. Ballard, M. S. Crowther, J. A. Dellinger, P. J. S. Fleming, A. S. Glen, A. C. Greenville, C. N. Johnson, M. Letnic, K. E. Moseby, D. G. Nimmo, M. P. Nelson, J. L. Read, W. J. Ripple, E. G. Ritchie, C. R. Shores, A. D. Wallach, A. J. Wirsing, and C. R. Dickman. 2015. Resolving the value of the dingo in ecological restoration. Restoration Ecology 23: 201-208. Nicholson, A. 1954. An outline of the dynamics of animal populations. Australian Journal of Zoology 2: 9-65. Prevedello, J. A., C. R. Dickman, M. V. Vieira, and E. M. Vieira. 2013. Population responses of small mammals to food supply and predators: a global meta-analysis. Journal of animal Ecology 82: 927–936. 224

Chapter 8: General discussion and conclusions

Previtali, M. A., M. Lima, P. L. Meserve, D. A. Kelt, and J. R. Gutiérrez. 2009. Population dynamics of two sympatric rodents in a variable environment: rainfall, resource availability, and predation. Ecology 90: 1996-2006. Read, J. L., J. Carter, K. M. Moseby, and A. Greenville. 2008. Ecological roles of rabbit, bettong and bilby warrens in arid Australia. Journal of Arid Environments 72: 2124- 2130. Ritchie, M. E. 2000. Nitrogen limitation and trophic vs. abiotic influences on insect herbivores in a temperate grassland. Ecology 81: 1601-1612. Sagarin, R. and A. Pauchard. 2009. Observational approaches in ecology open new ground in a changing world. Frontiers in Ecology and the Environment 8: 379-386. Sinclair, A. R. and C. J. Krebs. 2002. Complex numerical responses to top-down and bottom- up processes in vertebrate populations. Philos Trans R Soc Lond B Biol Sci 357: 1221-1231. Sinclair, A. R. E., C. J. Krebs, J. M. Fryxell, R. Turkington, S. Boutin, R. Boonstra, P. Seccombe-Hett, P. Lundberg, and L. Oksanen. 2000. Testing hypotheses of trophic level interactions: a boreal forest ecosystem. Oikos 89: 313-328. Smith, A. P. and D. G. Quin. 1996. Patterns and causes of extinction and decline in Australian conilurine rodents. Biological Conservation 77: 243-267. Smith, M. D. 2011. The ecological role of climate extremes: current understanding and future prospects. Journal of Ecology 99: 651-655. Turchin, P. 1995. Population regulation: old arguments and a new synthesis, Pages 19-40 in N. Cappuccino and P. W. Price, editors. Population dynamics: new approaches and synthesis. Academic Press, San Diego, California, USA. Academic Press, California. Tylianakis, J. M., R. K. Didham, J. Bascompte, and D. A. Wardle. 2008. Global change and species interactions in terrestrial ecosystems. Ecology Letters 11: 1351-1363. Van der Putten, W. H., M. Macel, and M. E. Visser. 2010. Predicting species distribution and abundance responses to climate change: why it is essential to include biotic interactions across trophic levels. Philos Trans R Soc Lond B Biol Sci 365: 2025- 2034. Vucetich, J. A. and R. O. Peterson. 2004. The influence of top-down, bottom-up and abiotic factors on the moose (Alces alces) population of Isle Royale. Proc Biol Sci 271: 183- 189.

225

Chapter 8: General discussion and conclusions

Welbergen, J. A., S. M. Klose, N. Markus, and P. Eby. 2008. Climate change and the effects of temperature extremes on Australian flying-foxes. Proceedings of the Royal Society B: Biological Sciences 275: 419-425. Woinarski, J. C. Z., A. A. Burbidge, and P. L. Harrison. 2014. The action plan for Australian mammals 2012. CSIRO Publishing, Vic.

226

Appendices

227

Appendix 1

Appendix 1:

Figure A1.1: Annual rainfall (mm) recorded each year for a) Boulia, b) Sandringham, c) Arltunga, and d) Mt. Dare weather stations, Simpson Desert, central Australia. Grey lines show 95th, 90th, 50th, 10th and 5thquantile. 90 and 95thquantiles represent boom years, 50th is the median and 10th and 5th represent bust years.

228

Appendix 2

Appendix 2:

Figure A2.1: Location of study sites across Cravens Peak Reserve, Carlo Station, Tobermorey Station and Ethabuka Reserve, Simpson Desert, Australia. Insert shows location of study region in Australia.

229

Appendix 2

Appendix A2.1: Study sites used to test five potential spatial population structures of small mammals in central Australia. Each hypothesis is entered into the model via the Z-matrix.

The Z-matrix is an n × m matrix of 0s and 1s, where n = number of sites (9) and m = each sub-population. See figure A2.1 for study region map:

a) Independent (m = 9): Mammals at all nine sites have asynchronous dynamics (each

site has a different population structure);

b) Oasis (m = 2): Mammals at sites associated with ephemeral water sources

(Tobermorey East, Tobermorey West, Field River South, Field River North and

Kunnamuka Swamp East) will have different population processes to those at open

desert sites (Main Camp, Shitty Site, South Site and Carlo; two population structures);

c) Productivity (m = 3) : Mammals at sites within the same rainfall zone will have

similar population dynamics (northern sites: Kunnamuka Swamp East, Carlo,

Tobermorey East; western sites: Field River South, Field River North and

Tobermorey West; and south-eastern sites: Main Camp, South Site and Shitty Site;

three population structures);

d) Wildfire (m = 2): Mammals at sites that have experienced a wildfire over the 22-year

study period will have different population dynamics to those at unburnt sites (six

sites burnt: Main Camp, Shitty Site, Field River South, Field River North, South Site

and Kunnamuka Swamp East; three sites unburnt: Carlo, Tobermorey East and

Tobermorey West; two population structures);

e) Single population (m = 1): Mammals at all sites have synchronous population

dynamics (one prevailing population structure).

230

Appendix 2

Appendix A2.2: Model formulations for spatial sub-population structures.

Process model, where X is abundance at each site n, the B diagonal elements represent the coefficients of autoregression in the populations through time t, and Bii represents the strength of density dependence (diagonal element Bii = 1 represents density independence, Bii < 1 = density

dependence), u describes the mean growth rate of the sub-population. w is the error, which follows a multivariate normal distribution, with mean

0 and variance Q:

= + + ; ~ (0, )

퐗t 퐁퐗t−1 퐮 퐰푡 풘푡 푀푀푀 푸 Observation model, where Z is an n × m matrix of 0s and 1s that assigns observations to a sub-population structure, a denotes the mean bias

between sites and V is error that follows a multivariate normal distribution, with mean 0 and variance R

= + + ; ~ (0, )

퐘푡 퐙퐗t 퐚 퐯푡 퐯푡 푀푀푀 퐑

231

Appendix 2

Independent model (m = 9)

Process model:

, 0 0 0 0 0 0 0 0 , , 0 0 0 0 0 0 0 0 , 푋푀푀푀푀 푡 𝑚𝑚 푋푀푀푀푀 푡−1 , 퐵 0 0 0 0 0 0 0 0 , 푀푀푀푀 ⎡ 퐶�퐶퐶퐶ℎ푡 ⎤ 퐶�퐶퐶퐶 ℎ ⎡ 퐶�퐶퐶퐶ℎ푡−1 ⎤ 휇 ℎ 푋 ⎡ ⎤ 푋 퐶�퐶퐶퐶 , 0 퐵 0 0 0 0 0 0 0 , ⎡ ⎤ ⎢ 푆�푆� 푡ℎ ⎥ ⎢ 푆�푆� ℎ ⎥ ⎢ 푆�푆� 푡ℎ−1 ⎥ 휇 ℎ ⎢ 푋 ⎥ 0 0 퐵 0 0 0 0 0 0 ⎢ 푋 ⎥ ⎢ 휇푆�푆� ⎥ 퐹�퐹�퐹𝐹𝐹ℎ,푡 = ⎢ 퐹�퐹�퐹𝐹𝐹 ℎ ⎥ 퐹�퐹�퐹𝐹𝐹ℎ,푡−1 + ℎ ⎢푋 ⎥ ⎢ 0 0 0 퐵 0 0 0 0 0 ⎥ ⎢푋 ⎥ ⎢휇퐹�퐹�퐹𝐹𝐹 ⎥ 퐹�퐹�ℎ퐹퐹𝐹�, 푡 퐹�퐹�퐹퐹𝐹� ℎ 퐹�퐹�ℎ퐹퐹𝐹,푡 푡−1 ⎢ 퐹�퐹�ℎ퐹퐹𝐹� ⎥ ⎢푋 ⎥ ⎢ 0 0 0 0 퐵 0 0 0 0 ⎥ ⎢푋 ⎥ 휇 푆 𝑖�푖 푡, 푆 𝑖�푖 푆 𝑖�푖 푡,−1 ⎢ 푆 𝑖�푖 ⎥ ⎢ 푋 ⎥ ⎢ 퐵 ⎥ ⎢ 푋 ⎥ ⎢ 휇 ⎥ , ⎢ 0 0 0 0 0 0 0 0 ⎥ , 퐾�퐾�퐾퐾 ⎢ 퐾�퐾�퐾퐾 푡 ⎥ 퐵퐾�퐾�퐾퐾 ⎢ 퐾�퐾�퐾퐾 푡−1 ⎥ ⎢ 휇 ⎥ 푋 ⎢ 0 0 0 0 0 0 0 0 ⎥ 푋 푇�푇푇�푇� ⎢ 푇�푇푇�푇�,푡 ⎥ 푇�푇푇�푇� ⎢ 푇�푇푇�푇�,푡−1 ⎥ ⎢ 휇 ⎥ ⎢ 푋 ⎥ ⎢ 퐵 ⎥ ⎢ 푋 ⎥ 푇�푇푇�푇� 푇�푇푇�푇� 푡 푇�푇푇�푇� 푇�푇푇�푇� 푡−1 ⎣ 휇 ⎦ ⎣ 푋 ⎦ ⎣ , 퐵 ⎦ ⎣ 푋 ⎦ , 푀푀푀푀 푡 푤 , ⎡ 퐶�퐶퐶퐶ℎ푡 ⎤ 푤 , ⎢ 푆�푆� 푡ℎ ⎥ + 푤 , ; ~ ( , ) ⎢ 퐹�퐹�퐹𝐹𝐹ℎ 푡⎥ 푤 ℎ ⎢ 푤퐹�퐹�퐹퐹𝐹� 푡⎥ 푡 ⎢ ⎥ 퐰 푀푀푀 ퟎ 퐐 푤푆 𝑖�푖 ⎢ 퐾�퐾�퐾퐾 ⎥ ⎢ 푤 ⎥ ⎢ 푤푇�푇푇�푇� ⎥ ⎣ 푤푇�푇푇�푇� ⎦ 0 0 0 0 0 0 0 0 20 0 0 0 0 0 0 0 푞푀푀푀푀 ⎡ 0 20 0 0 0 0 0 0 ⎤ 푞퐶�퐶퐶퐶 ℎ ⎢ 0 0 20 0 0 0 0 0 ⎥ ⎢ 푞푆�푆� ℎ ⎥ = 0 0 0 2 0 0 0 0 0 ⎢ 푞퐹�퐹�퐹𝐹𝐹 ℎ ⎥ ⎢ 0 0 0 0 2 0 0 0 0 ⎥ 퐐 푞퐹�퐹�퐹퐹𝐹� ℎ ⎢ 0 0 0 0 0 2 0 0 0 ⎥ ⎢ 푞푆 𝑖�푖 ⎥ 0 0 0 0 0 0 2 0 0 ⎢ 푞퐾�퐾�퐾퐾 ⎥ 2 ⎢ 0 0 0 0 0 0 0 푇�푇푇0 �푇� ⎥ ⎢ 푞 2 ⎥ ⎣ 232 푞푇�푇푇�푇�⎦

Appendix 2

Observation model:

, , 1 0 0 0 0 0 0 0 0 0 , , , 푀푀푀푀 푡 0 1 0 0 0 0 0 0 0 푀푀푀푀 푡 , 푌 푋 0 푀푀푀푀 푡 , , 푣 ⎡ 퐶�퐶퐶퐶ℎ 푡 ⎤ 0 0 1 0 0 0 0 0 0 ⎡ 퐶�퐶퐶퐶ℎ푡 ⎤ 0 ℎ, 푌 ⎡ ⎤ 푋 ⎡ 퐶�퐶퐶퐶 푡 ⎤ ⎢ ℎ, ⎥ 0 0 0 1 0 0 0 0 0 ⎢ ℎ, ⎥ ⎡ ⎤ 푣 , 푆�푆� 푡 푆�푆� 푡 0 푆�푆� 푡ℎ ⎢ 푌 ⎥ ⎢ ⎥ ⎢ 푋 ⎥ ⎢ ⎥ ⎢ 푣 ⎥ 퐹�퐹�퐹𝐹𝐹ℎ,푡 = 0 0 0 0 1 0 0 0 0 퐹�퐹�퐹𝐹𝐹ℎ,푡 + 0 + ℎ, ; ~ ( , ) 푌 ⎢ ⎥ 푋 ⎢ ⎥ ⎢ 푣퐹�퐹�퐹𝐹𝐹 푡 ⎥ ⎢ , ⎥ 0 0 0 0 0 1 0 0 0 ⎢ , ⎥ 0 ℎ ⎢ 퐹�퐹�ℎ퐹퐹𝐹� 푡⎥ ⎢ ⎥ ⎢ 퐹�퐹�ℎ퐹퐹𝐹� 푡⎥ ⎢ ⎥ ⎢ 푣퐹�퐹�퐹퐹𝐹� 푡⎥ 푡 푌 ⎢0 0 0 0 0 0 1 0 0⎥ 푋 0 ⎢ ⎥ 퐕 푀푀푀 ퟎ 퐑 ⎢ 푆 𝑖�푖 푡, ⎥ ⎢ 푆 𝑖�푖 푡, ⎥ ⎢ ⎥ 푆 𝑖�푖 푌 ⎢0 0 0 0 0 0 0 1 0⎥ 푋 0 ⎢ 푣 ⎥ , , 퐾�퐾�퐾퐾 ⎢ 퐾�퐾�퐾퐾 푡 ⎥ ⎢ ⎥ ⎢ 퐾�퐾�퐾퐾 푡 ⎥ ⎢0⎥ ⎢ 푣 ⎥ 푌 0 0 0 0 0 0 0 0 1 푋 푇�푇푇�푇� ⎢ 푇�푇푇�푇�,푡 ⎥ ⎢ 푇�푇푇�푇�,푡 ⎥ ⎢ ⎥ 푣 푌 ⎢ ⎥ 푋 ⎢ ⎥ ⎢ ⎥ ⎣ ⎦ ⎢ ⎥ ⎣ ⎦ ⎣ 푣푇�푇푇�푇� ⎦ ⎣ 푌푇�푇푇�푇� 푡 ⎦ ⎣ 푋푇�푇푇�푇� 푡 ⎦ 0 0 0 0 0 0 0 0 02 0 0 0 0 0 0 0 푟 ⎡ 0 02 0 0 0 0 0 0 ⎤ 푟 ⎢ 0 0 02 0 0 0 0 0 ⎥ ⎢ 푟 ⎥ = 0 0 0 02 0 0 0 0 ⎢ 푟 ⎥ ⎢ 0 0 0 0 02 0 0 0 ⎥ 퐑 푟 ⎢ 0 0 0 0 0 02 0 0 ⎥ ⎢ 푟 ⎥ 0 0 0 0 0 0 02 0 ⎢ 푟 ⎥ ⎢ 0 0 0 0 0 0 0 02 ⎥ ⎢ 푟 2 ⎥ ⎣ 푟 ⎦

233

Appendix 2

Oasis population model (m = 2):

Process model:

, 0 , , = + + ; ~ ( , ) 0 , 푂𝑂𝑂,푡 퐵푂푂푂푂푂 푂𝑂𝑂,푡−1 푂𝑂𝑂 푂𝑂𝑂 푡 푋 푋 휇 푤 푡 � � � 표표표표� � � � 푂𝑂�� � 푂𝑂� 푡 � 퐰 푀푀푀 ퟎ 퐐 푋푂𝑂� 푡 퐵 푋푂𝑂� 푡−1 휇 푤 0 = 20 푞푂𝑂𝑂 퐐 � 2 � 푞푂𝑂� Observation model:

, 0 1 0 , , 푀푀푀푀 푡 0 1 , 푌 푀푀푀푀 푡 ⎡ 퐶�퐶퐶퐶ℎ,푡 ⎤ 0 1 푣 , ℎ 퐶�퐶퐶퐶ℎ푡 푌 ⎡ ⎤ 퐶�퐶퐶퐶 ⎡ ⎤ ⎢ ℎ, ⎥ ⎡ ⎤ 푣 , 푆�푆� 푡 1 0 � 0 푆�푆� 푡ℎ ⎢ 푌 ⎥ ⎢ ⎥ , ⎢ 푆�푆� ⎥ ⎢ 푣 ⎥ 퐹�퐹�퐹𝐹𝐹ℎ,푡 = 1 0 + � ℎ + ℎ, ; ~ ( , ) ⎢푌 ⎥ ⎢ ⎥ , ⎢ ⎥ ⎢ 푣퐹�퐹�퐹𝐹𝐹 푡⎥ ℎ , ⎢0 1⎥ 푂푂푂푂푂 푡 퐹�퐹�ℎ퐹퐹𝐹� 퐹�퐹�퐹퐹ℎ 𝐹� 푡 ⎢ 퐹�퐹�퐹퐹𝐹� 푡⎥ 푋 ⎢� ⎥ ⎢ 푣 ⎥ 푡 푌 ⎢1 0⎥ � � ⎢ ⎥ ⎢ ⎥ 퐕 푀푀푀 ퟎ 퐑 ⎢ 푆 𝑖�푖 푡, ⎥ 푂𝑂� 푡 푆 𝑖�푖 푆 𝑖�푖 푌 ⎢1 0⎥ 푋 ⎢ � ⎥ ⎢ 푣 ⎥ ⎢ , ⎥ 퐾�퐾�퐾퐾 퐾�퐾�퐾퐾 푌퐾�퐾�퐾퐾 푡 ⎢1 0⎥ ⎢ � ⎥ ⎢ 푣 ⎥ ⎢ , ⎥ 푇�푇푇�푇� 푇�푇푇�푇� 푇�푇푇�푇� 푡 ⎢ ⎥ ⎢ � ⎥ ⎢ 푣 ⎥ 푌 푇�푇푇�푇� ⎢ ⎥ ⎣ ⎦ ⎣ � ⎦ ⎣ 푣푇�푇푇�푇� ⎦ ⎣ 푌푇�푇푇�푇� 푡 ⎦

234

Appendix 2

0 0 0 0 0 0 0 0 02 0 0 0 0 0 0 0 푟 ⎡ 0 02 0 0 0 0 0 0 ⎤ 푟 ⎢ 0 0 02 0 0 0 0 0 ⎥ ⎢ 푟 ⎥ = 0 0 0 02 0 0 0 0 ⎢ 푟 ⎥ ⎢ 0 0 0 0 02 0 0 0 ⎥ 퐑 푟 ⎢ 0 0 0 0 0 02 0 0 ⎥ ⎢ 푟 ⎥ 0 0 0 0 0 0 02 0 ⎢ 푟 ⎥ ⎢ 0 0 0 0 0 0 0 02 ⎥ ⎢ 푟 2 ⎥ ⎣ 푟 ⎦ Productivity population model (m = 3)

Process model:

ℎ, ℎ 0 0 ℎ, ℎ ℎ, , = 0 0 , + + ℎ, ; ~ ( , ) 𝑁𝑁ℎ 푡 𝑁𝑁 ℎ 𝑁𝑁ℎ 푡−1 𝑁𝑁ℎ 𝑁𝑁 푡 푋 퐵 0 0 푋 휇 푤 , , , 푆�푆� 푆�푆� 푡 �푋푆�푆� 푡 � � 퐵푆�푆� � �푋푆�푆� 푡−1� �휇 � �푤 � 퐰푡 푀푀푀 ퟎ 퐐 𝑊𝑊 𝑊𝑊 푡 푋𝑊𝑊 푡 퐵𝑊𝑊 푋𝑊𝑊 푡−1 휇 푤 0 0 ℎ = 20 0 푞𝑁𝑁 ℎ 0 20 퐐 � 푆�푆� � 푞 2 푞𝑊𝑊

235

Appendix 2

Observation model:

, 0 1 0 0 , , 푀푀푀푀 푡 1 0 0 , 푌 0 푀푀푀푀 푡 ⎡ 퐶�퐶퐶퐶ℎ,푡 ⎤ 0 1 0 푣 , 푌 ⎡ ⎤ ℎ ⎡ 퐶�퐶퐶퐶ℎ푡 ⎤ ⎢ 푆�푆� 푡ℎ, ⎥ 0 0 1 ℎ, ⎡ 0 ⎤ 푣 ℎ, 푌 ⎢ ⎥ 푆�푆� ⎢ 푣푆�푆� 푡 ⎥ ⎢ ℎ, ⎥ = 0 0 1 , + ⎢ � ℎ⎥ + ℎ, ; ~ ( , ) 푌퐹�퐹�퐹𝐹𝐹 푡 ⎢ ⎥ 𝑁�푡ℎ 푡 ⎢ 퐹�퐹�퐹𝐹𝐹 푡 ⎥ ⎢ ⎥ 푋 ⎢ ℎ ⎥ 푣 퐹�퐹�ℎ퐹퐹𝐹�, 푡 ⎢0 1 0⎥ , 퐹�퐹�퐹퐹𝐹� ⎢ 퐹�퐹�퐹퐹ℎ 𝐹� 푡⎥ ⎢푌 ⎥ �푋푆�푆� 푡 � ⎢� ⎥ 푣 퐕푡 푀푀푀 ퟎ 퐑 , ⎢1 0 0⎥ 푆 𝑖�푖 ⎢ 푆 𝑖�푖 ⎥ ⎢ 푌푆 𝑖�푖 푡 ⎥ 푋𝑊𝑊 푡 ⎢ � ⎥ 푣 ⎢0 0 1⎥ 퐾�퐾�퐾퐾 ⎢ 퐾�퐾�퐾퐾 ⎥ ⎢ 퐾�퐾�퐾퐾,푡 ⎥ ⎢ � ⎥ ⎢ 푣 ⎥ 푌 ⎢1 0 0⎥ 푇�푇푇�푇� 푇�푇푇�푇� ⎢ 푇�푇푇�푇�,푡 ⎥ ⎢ � ⎥ ⎢ 푣 ⎥ 푌 ⎢ ⎥ 푇�푇푇�푇� ⎢ ⎥ ⎣ ⎦ ⎣ � ⎦ ⎣ 푣푇�푇푇�푇� ⎦ ⎣ 푌푇�푇푇�푇� 푡 ⎦ 0 0 0 0 0 0 0 0 02 0 0 0 0 0 0 0 푟 ⎡ 0 02 0 0 0 0 0 0 ⎤ 푟 ⎢ 0 0 02 0 0 0 0 0 ⎥ ⎢ 푟 ⎥ = 0 0 0 02 0 0 0 0 ⎢ 푟 ⎥ ⎢ 0 0 0 0 02 0 0 0 ⎥ 퐑 푟 ⎢ 0 0 0 0 0 02 0 0 ⎥ ⎢ 푟 ⎥ 0 0 0 0 0 0 02 0 ⎢ 푟 ⎥ ⎢ 0 0 0 0 0 0 0 02 ⎥ ⎢ 푟 2 ⎥ ⎣ 푟 ⎦

236

Appendix 2

Wildfire Model (m = 2):

Process model:

, 0 , , = + + ; ~ ( , ) 0 , 퐵𝐵𝐵 푡, 퐵퐵퐵𝐵� 퐵𝐵𝐵 푡,−1 퐵𝐵𝐵 퐵𝐵𝐵 푡 푋 푋 휇 푤 푡 � � � 푈�푈𝑈𝑈� � � � 푈𝑈𝑈𝑈� � 푈𝑈𝑈𝑈 푡� 퐰 푀푀푀 ퟎ 퐐 푋푈𝑈𝑈𝑈 푡 퐵 푋푈𝑈𝑈𝑈 푡−1 휇 푤 0 = 20 푞퐵𝐵𝐵 퐐 � 2 � 푞푈𝑈𝑈𝑈 Observation model:

, 1 0 0 , , 푀푀푀푀 푡 0 1 , 푌 0 푀푀푀푀 푡 , 푣 ⎡ 퐶�퐶퐶퐶ℎ푡 ⎤ 1 0 ℎ ℎ, 푌 ⎡ ⎤ ⎡ ⎤ ⎡ 퐶�퐶퐶퐶 푡 ⎤ ⎢ 푆�푆� 푡ℎ, ⎥ 1 0 푣 ℎ, 푌 ⎢ ⎥ , ⎢ 푆�푆� ℎ⎥ ⎢ 푆�푆� 푡 ⎥ ⎢ , ⎥ = 1 0 + � ℎ + 푣 ℎ, ; ~ ( , ) 퐹�퐹�퐹𝐹𝐹ℎ 푡 ⎢ ⎥ 퐹�퐹�퐹𝐹𝐹 퐹�퐹�퐹𝐹𝐹 푡 ⎢푌 ⎥ , ⎢� ⎥ ⎢ 푣 ⎥ 퐹�퐹�ℎ퐹퐹𝐹�, 푡 ⎢1 0⎥ 퐵퐵𝐵� 푡 ⎢ 퐹�퐹�ℎ퐹퐹𝐹� ⎥ ⎢ 퐹�퐹�퐹퐹ℎ 𝐹� 푡⎥ ⎢푌 ⎥ � 푋 � � 푣 퐕푡 푀푀푀 ퟎ 퐑 , ⎢1 0⎥ 푈�푈𝑈𝑈 푡 ⎢ 푆 𝑖�푖 ⎥ ⎢ 푆 𝑖�푖 ⎥ ⎢ 푌푆 𝑖�푖 푡 ⎥ 푋 � 푣 ⎢0 1⎥ ⎢ 퐾�퐾�𝑚 ⎥ ⎢ 퐾�퐾�퐾퐾 ⎥ ⎢ 퐾�퐾�퐾퐾,푡 ⎥ ⎢ � ⎥ ⎢ 푣 ⎥ 푌 ⎢0 1⎥ 푇�푇푇�푇� 푇�푇푇�푇� ⎢ 푇�푇푇�푇�,푡 ⎥ ⎢ � ⎥ ⎢ 푣 ⎥ 푌 ⎢ ⎥ 푇�푇푇�푇� ⎢ ⎥ ⎣ ⎦ ⎣ � ⎦ ⎣ 푣푇�푇푇�푇� ⎦ ⎣ 푌푇�푇푇�푇� 푡 ⎦

237

Appendix 2

0 0 0 0 0 0 0 0 02 0 0 0 0 0 0 0 푟 ⎡ 0 02 0 0 0 0 0 0 ⎤ 푟 ⎢ 0 0 02 0 0 0 0 0 ⎥ ⎢ 푟 ⎥ = 0 0 0 02 0 0 0 0 ⎢ 푟 ⎥ ⎢ 0 0 0 0 02 0 0 0 ⎥ 퐑 푟 ⎢ 0 0 0 0 0 02 0 0 ⎥ ⎢ 푟 ⎥ 0 0 0 0 0 0 02 0 ⎢ 푟 ⎥ ⎢ 0 0 0 0 0 0 0 02 ⎥ ⎢ 푟 2 ⎥ ⎣ 푟 ⎦

Single population model (m = 1)

Process model:

, = , + , + , ; , ~ (0, )

푋푆𝑆푆𝑆 푡 퐵�푆𝑆푆𝑆 푡−1 휇푆𝑆푆𝑆 푡 푤푆𝑆푆𝑆 푡 푤푆𝑆푆𝑆 푡 � 푄 Observation model:

, 1 0 , , 푀푀푀푀 푡 1 , 푌 푀푀푀푀 푡 ℎ, 푣 ⎡ 퐶�퐶퐶퐶 푡 ⎤ 1 ℎ ℎ, 푌 ⎡ ⎤ ⎡ 퐶�퐶퐶퐶 ⎤ ⎡ 퐶�퐶퐶퐶 푡 ⎤ ⎢ ℎ, ⎥ 푣 , 푆�푆� 푡 1 � ℎ 푆�푆� 푡ℎ ⎢ 푌 ⎥ ⎢ ⎥ ⎢ 푆�푆� ⎥ ⎢ 푣 ⎥ 퐹�퐹�퐹𝐹𝐹ℎ,푡 = 1 , + � ℎ + ℎ, ; ~ ( , ) ⎢푌 ⎥ ⎢ ⎥ ⎢ 퐹�퐹�퐹𝐹𝐹 ⎥ ⎢ 푣퐹�퐹�퐹𝐹𝐹 푡⎥ , 1 � ℎ ℎ 퐹�퐹�ℎ퐹퐹𝐹� 푡 ⎢ ⎥ ⎢ 퐹�퐹�퐹퐹𝐹� ⎥ ⎢ 퐹�퐹�퐹퐹𝐹� 푡⎥ 푡 ⎢푌 ⎥ 1 � 푆𝑆푆푆� 푡� � 푣 퐕 푀푀푀 ퟎ 퐑 푆 𝑖�푖 푡, ⎢ ⎥ 푋 ⎢ 푆 𝑖�푖 ⎥ ⎢ 푆 𝑖�푖 ⎥ ⎢ 푌 ⎥ ⎢1⎥ ⎢ � ⎥ ⎢ 푣 ⎥ ⎢ , ⎥ 퐾�퐾�퐾퐾 퐾�퐾�퐾퐾 푌퐾�퐾�퐾퐾 푡 ⎢ ⎥ ⎢ � ⎥ ⎢ 푣 ⎥ 1 푇�푇푇�푇� 푇�푇푇�푇� ⎢ 푇�푇푇�푇�, 푡 ⎥ ⎢ ⎥ ⎢ � ⎥ ⎢ 푣 ⎥ 푌 푇�푇푇�푇� ⎢ ⎥ ⎣ ⎦ ⎣ � ⎦ ⎣ 푣푇�푇푇�푇� ⎦ ⎣ 푌푇�푇푇�푇� 푡 ⎦ 238

Appendix 2

0 0 0 0 0 0 0 0 02 0 0 0 0 0 0 0 푟 ⎡ 0 02 0 0 0 0 0 0 ⎤ 푟 ⎢ 0 0 02 0 0 0 0 0 ⎥ ⎢ 푟 ⎥ = 0 0 0 02 0 0 0 0 ⎢ 푟 ⎥ ⎢ 0 0 0 0 02 0 0 0 ⎥ 퐑 푟 ⎢ 0 0 0 0 0 02 0 0 ⎥ ⎢ 푟 ⎥ 0 0 0 0 0 0 02 0 ⎢ 푟 ⎥ ⎢ 0 0 0 0 0 0 0 02 ⎥ ⎢ 푟 2 ⎥ ⎣ 푟 ⎦

239

Appendix 2

Example single population model for rodents with covariates entered via the B matrix.

Data are ‘de-meaned’ so u and a set to zero.

Process model:

, , , , , , 0 , , 0 0 0 0 , 0 , 푋푅𝑅𝑅� 푡 푅𝑅𝑅� 푅𝑅𝑅� 푅𝑅� 푅𝑅𝑅� 푆𝑆� 푅𝑅𝑅� 푆𝑆� 푅𝑅𝑅� 푀𝑀�푀�푀 푋푅𝑅𝑅� 푡−1 푅𝑅𝑅� 푡 퐵 0 퐵 0 퐵 퐵 0 퐵 0 푤 , ⎡ 푅𝑅�,푡 ⎤ = ⎡ ⎤ ⎡ 𝑟𝑟,푡−1 ⎤ + 0 + 푅𝑅� 푡 ; ~ ( , ) 푋 퐵푅𝑅� 푋 ⎡ ⎤ ⎡ 푤 ⎤ ⎢ , ⎥ ⎢ 0 0 0 0 ⎥ ⎢ , ⎥ 0 , 푆𝑆� 푡 푆𝑆� 푆𝑆� 푡−1 ⎢ ⎥ ⎢ 푤푆𝑆� 푡 ⎥ 푡 ⎢ 푋 ⎥ ⎢ 0 0 퐵 0 0 ⎥ ⎢ 푋 ⎥ ⎢0⎥ ⎢ , ⎥ 퐰 푀푀푀 ퟎ 퐐 ⎢ 푆𝑆� 푡 , ⎥ 푆𝑆� ⎢ 푆𝑆� 푡−,1 ⎥ 푆𝑆� 푡 푋 ⎢ 퐵 ⎥ 푋 ⎢ ⎥ ⎢ 푤 ⎥ ⎢ 푀𝑀�푀�푀 푡⎥ 𝑚𝑚𝑚� ⎢ 𝑚𝑚𝑚� 푡−1⎥ 𝑚𝑚𝑚� 푡 ⎣푋 ⎦ ⎣ 퐵 ⎦ ⎣푋 ⎦ ⎣ ⎦ ⎣푤 ⎦ 0 0 0 0 2 0 0 0 0 푞푅𝑅𝑅� = ⎡ 0 20 0 0 ⎤ 푞푅𝑅� ⎢ 0 0 20 0 ⎥ 퐐 ⎢ 푞푆𝑆� ⎥ ⎢ 0 0 0 20 ⎥ 푞푆𝑆� ⎢ 2 ⎥ 𝑚𝑚𝑚� ⎣ 푞 ⎦

240

Appendix 2

Observation model:

, 1 0 0 0 0 0 , ,1 푅𝑅𝑅� 푡 0 1 0 0 0 , 푌 0 푅𝑅𝑅�1 푡 , ⎡ 푅𝑅�1 푡 ⎤ 0 0 1 0 0 0 푣 , 푌 ⎡ ⎤ ⎡ 푅𝑅�1 푡 ⎤ ⎢ 푆𝑆�1,푡 ⎥ 0 0 0 1 0 , ⎡0⎤ 푣 , ⎢ 푌 ⎥ ⎢ ⎥ ⎢ 푆𝑆�1 푡 ⎥ , 0 0 0 0 1 , ⎢0⎥ 푣 , ⎢ 푌푆𝑆�1 푡 ⎥ ⎢ ⎥ 푋푅𝑅𝑅� 푡 ⎢ 푆𝑆�1 푡 ⎥ = , + ⎢ ⎥ + 푣 ; ~ ( , ) 𝑚𝑚𝑚�1 푡 ⎢ ⎥ ⎡ 푅𝑅� 푡 ⎤ 1 ⎢푌 ⎥ 푋 ⎢ ⎥ ⎢ 푣𝑚𝑚𝑚� 푡⎥ , 1 0 0 0 0 ⎢ , ⎥ 0 , ⎢ ⎥ ⎢ ⎥ 푆𝑆� 푡 ⎢ ⎥ 푡 ⎢ 푋 ⎥ ⎢ ⎥ , ⎢ ⋮ , ⎥ ⎢0⋮ 1⋮ 0⋮ 0⋮ 0⋮⎥ 0⋮ ⎢ 푅𝑅𝑅�⋮ 9 푡 ⎥ 퐕 푀푀푀 ퟎ 퐑 푅𝑅𝑅�9 푡 ⎢ 푆𝑆� 푡 , ⎥ 푌 푋 ⎢0⎥ 푣 , ⎢ , ⎥ ⎢0 0 1 0 0⎥ ⎢ ⎥ ⎢ 푅𝑅�9 푡 ⎥ 푅𝑅�9 푡 푀𝑀�푀�푀 푡 ⎢0⎥ 푣 ⎢ 푌 ⎥ ⎢0 0 0 1 0⎥ 푋 ⎢ 9, ⎥ , ⎣ ⎦ 푆𝑆� 푡 ⎢ 푆𝑆�9 푡 ⎥ ⎢0⎥ 푣 , 푌 ⎢0 0 0 0 1⎥ ⎢ 푆𝑆�9 푡 ⎥ ⎢ 푆𝑆�9 푡 , ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ 푣 ⎥ ⎢ 푌 ⎥ ⎣ ⎦ 푣𝑚𝑚𝑚�9 푡 𝑚𝑚𝑚�9 푡 ⎣ ⎦ ⎣ ⎦ ⎣푌 ⎦ Note that subscripts represent each of the nine sites.

0 0 0 0 … 0 0 0 0 0 2 0 0 0 0 … 0 0 0 0 0 푟푅𝑅𝑅�1 ⎡ 2 ⎤ 0 푅𝑅�0 1 0 0 … 0 0 0 0 0 ⎢ 푟 ⎥ 0 0 20 0 … 0 0 0 0 0 ⎢ 푟푆𝑆�1 ⎥ ⎢ 0 0 0 20 … 0 0 0 0 0 ⎥ 푟푆𝑆�1 = ⎢ 2 ⎥ 𝑚𝑚𝑚�1 ⎢ 0 0 0 0 푟 0 … 0 0 0 0 ⎥ ⎢ ⎥ 퐑 0⋮ 0⋮ 0⋮ 0⋮ 0⋮ …⋱ 2 0⋮ ⋮ 0⋮ 0⋮ 0⋮ ⎢ 푟푅𝑅𝑅�9 ⎥ ⎢ 0 0 0 0 0 … 0 20 0 0 ⎥ 푟푅𝑅�9 ⎢ 0 0 0 0 0 … 0 0 20 0 ⎥ ⎢ 푟푆𝑆�9 ⎥ 2 ⎢ 0 0 0 0 0 … 0 0 0 0 9 ⎥ 푟푆𝑆� ⎢ 2 ⎥ 𝑚𝑚𝑚�9 ⎣ 푟 ⎦ Note that R is a 45 x 45 matrix where subscripts represent each of the nine sites.

241

Appendix 2

(a) (b) 1.0 1.0 0.0 0.0 -1.0 -1.0

0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70

(c) (d) 1.0 1.0 Correlation 0.0 0.0 -1.0 -1.0

0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70

(e) 1.0 0.0 -1.0

0 10 20 30 40 50 60 70

Distance (km)

Figure A2.2: Spatial cross-correlations (solid line) up to 70 km from 17–22 years of live- trapping across nine sub-populations of (a) Pseudomys hermannsburgensis, (b) Notomys alexis, (c) Dasycercus blythi, (d) Sminthopsis youngsoni, and (e) Ningaui ridei, in central Australia. Dashed lines represent 95% confidence intervals. Analysis performed in R 3.02 (R Core Team 2014) using the ncf 1.1-5 package (Bjornstad 2013). Due to limited number of sub-populations (9), resampling was limited to 50. This limitation was confirmed by artificially increasing the number of sub-populations to 18 and re-running the analyses with greater resampling effort (50, 100, 1000 and 2000).

242

Appendix 2

Table A2.1: AICc values for MARSS models describing five possible spatial sub-population structures, with density dependence, for two species of small mammals in the Simpson Desert, central Australia. Data are based on 17 years of live-trapping from nine sites. Smallest AICc, highlighted in bold, indicates the best-fitting model. Model With density dependence S. youngsoni Independent 592.66 Oasis 607.82 Productivity 612.49 Wildfire 606.97 Single population 615.26

N. ridei Independent 418.51* Oasis 452.13 Productivity 444.34 Wildfire 463.98 Single population 463.98 *Process error (Q) set to equal across all sub-populations

References:

Bjornstad, O. N. 2013. ncf: spatial nonparametric covariance functions. R package version 1.1-5. R Core Team. 2014. R: a language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria.

243

Appendix 3

Appendix 3:

Figure A3.1: Spatial cross-correlations (solid line) up to 70 km from 1995–2012 of annual rainfall from nine sites in central Australia. Dashed lines represent 95% confidence intervals. Analyses were performed in R 3.02 (R Core Team 2014) using the ncf 1.1-5 package (Bjornstad 2013).

Table A3.1: Total captures of each species of reptile from long-term (17–22 years) live- trapping (205 524 trap nights), across nine sites, Simpson Desert, central Australia. Taxonomy follows Cogger (2000).

Species Total captures Ctenophorus isolepis 2514 Ctenophorus nuchalis 1303 Ctenotus ariadnae 748 Ctenotus dux 868 Ctenotus pantherinus 1474 Lerista labialis 5663

244

Appendix 3

Figure A3.2: Spatial cross-correlations (solid line) up to 70 km from 17–22 years of live- trapping across nine sub-populations of (a) Ctenotus pantherinus, (b) Ctenotus ariadnae, (c) Ctenotus dux, (d) Lerista labialis, (e) Ctenophorus isolepis and (f) Ctenophorus nuchalis, in central Australia. Dashed lines represent 95% confidence intervals. Analyses were performed in R 3.02 (R Core Team 2014) using the ncf 1.1-5 package (Bjornstad 2013). Due to limited numbers of sub-populations (9), resampling was limited to 50. This limitation was confirmed by artificially increasing the number of sub-populations to 18 and re-running the analyses with greater resampling effort (50, 100, 1000 and 2000). References:

Bjornstad, O. N. 2013. ncf: spatial nonparametric covariance functions. R package version 1.1-5. Cogger, H. G. 2000. Reptiles and amphibians of Australia. 6th edition. Reed New Holland, Sydney, Australia. R Core Team. 2014. R: a language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria.

245

Appendix 4

Appendix 4: 0.8 Dingo

0.7 Feral cat Red fox 0.6

0.5

0.4

0.3

0.2

Mean numberMean photographs of (SE) 0.1

0 2010.4 2010.5 2010.6 2010.7 2010.8 2010.9 2011.1 2011.2 2011.3 2011.4 2011.5 2011.6 2011.7 2011.8 2011.9 2012.1 2012.2 2012.3 2012.4 2010.10 2010.11 2010.12 2011.10 2011.11 2011.12

Figure A4.1: Numbers of predator photographs obtained per camera night (means ± SE) over two years from twenty-five remote cameras (24 Moultrie i40 and one Reconyx RapidFire) set in the Simpson Desert, central Australia. The two-year study followed predator populations before, during and after an irruption in their rodent prey.

246

Appendix 4

(a) 0.12 0.08 0.04 0.00

0 100 200 300 400 500

(b) 0.15 0.10 Density 0.05 0.00

0 100 200 300 400 500

(c) 0.4 0.3 0.2 0.1 0.0

0 100 200 300 400 500 Time between images (min)

Figure A4.2: Histograms showing the proportions of remote camera photographs taken within 500 minutes with time between recurring photographs of (a) red fox, (b) feral cat and (c) dingo at the same camera trapping location, Simpson Desert, central Australia. Twenty- five remote cameras (24 Moultrie i40 and one Reconyx RapidFire) were active from April 2010 to April 2012 and downloaded 3–4 times a year. Bin size 0, 1, 2, 3, 10, 100, 200, 500 minutes. A two minute break point was chosen to assume independence between individual photographs based on these results.

247

Appendix 4

3 160 Camera

140

2.5 Live-trapping Captures (100 trapnights) 120 2 100

1.5 80

60 1 40

0.5 20 Mean numberMean photographs of (SE) 0 0 2010.4 2010.5 2010.6 2010.7 2010.8 2010.9 2011.1 2011.2 2011.3 2011.4 2011.5 2011.6 2011.7 2011.8 2011.9 2012.1 2012.2 2012.3 2012.4 2010.10 2010.11 2010.12 2011.10 2011.11 2011.12

Figure A4.3: Numbers of rodent photographs (four species, pooled; solid line) obtained per camera night (means ± SE) over two years from twenty-five remote cameras (24 Moultrie i40 and one Reconyx RapidFire) and live-trapping captures (means ± SE per 100 trap nights; dashed line) in the Simpson Desert, central Australia. Rodents comprised the long-haired rat, Rattus villosissimus, sandy inland mouse Pseudomys hermannsburgensis, spinifex hopping mouse, Notomys alexis and house mouse Mus musculus. Both datasets were used to estimate the timing of prey population phases.

248

Appendix 4

Table A4.1: GLM parameter estimates for numbers of detections of animals in photographs obtained during remote camera trapping of the dingo, red fox and feral cat in the Simpson Desert, central Australia. Cameras were active for two years and photographs pooled per population phase. Phase is the population phase of prey (bust, boom or decline).

Interaction Estimate SE Z value P

Dingo-fox interaction

Dingo 0.02 0.004 4.49 <0.01

Phase bust −0.22 0.03 −6.83 <0.01

Phase decline −0.41 0.03 −14.12 <0.01

Dingo × Phase bust −0.36 0.008 −43.65 <0.01

Dingo × Phase decline −0.03 0.004 −7.89 <0.01

Dingo-cat interaction

Dingo −0.02 0.004 −4.80 <0.01

Phase bust −0.54 0.03 −16.82 <0.01

Phase decline −0.37 0.03 −13.12 <0.01

Dingo × Phase bust −0.06 0.006 −9.24 <0.01

Dingo × Phase decline 0.001 0.004 0.270 0.79

Fox-cat interaction

Fox 0.02 0.002 11.09 <0.01

Phase bust −1.00 0.03 −33.14 <0.01

Phase decline −0.53 0.03 −16.06 <0.01

Dingo × Phase bust 0.06 0.003 19.87 <0.01

Dingo × Phase decline 0.03 0.003 10.40 <0.01

249

Appendix 4

Table A4.2: GLM parameter estimates with outliers removed for numbers of detections of animals in photographs obtained during remote camera trapping of the dingo, red fox and feral cat in the Simpson Desert, central Australia. Cameras were active for two years and photographs pooled per population phase. Phase is the population phase of prey (bust, boom or decline).

Interaction Estimate SE Z value P

Dingo-fox interaction

Dingo 0.02 0.004 4.43 <0.01

Phase bust −0.22 0.03 −6.71 <0.01

Phase decline −0.41 0.03 −13.66 <0.01

Dingo × Phase bust −0.36 0.008 −43.32 <0.01

Dingo × Phase decline −0.03 0.006 −4.83 <0.01

Dingo-cat interaction

Dingo −0.02 0.004 −4.76 <0.01

Phase bust −0.54 0.03 −16.68 <0.01

Phase decline −0.31 0.03 −10.55 <0.01

Dingo × Phase bust −0.06 0.006 −9.17 <0.01

Dingo × Phase decline −0.03 0.006 −5.86 <0.01

250

Appendix 5

Appendix 5:

Table A5.1: Total captures of each species of reptile from long-term (17–22 years) live- trapping (205 524 trap nights), across nine sites, Simpson Desert, central Australia. Taxonomy follows Cogger (2000).

Species Total captures Ctenophorus clayi 15 Ctenophorus isolepis 2514 Ctenophorus nuchalis 1303 Ctenotus ariadnae 748 Ctenotus brooksi 145 Ctenotus calurus 139 Ctenotus dux 868 Ctenotus helenae 87 Ctenotus lateralis 41 Ctenotus leae 314 Ctenotus leonhardii 20 Ctenotus pantherinus 1474 Ctenotus piankai 73 Ctenotus regius 21 Ctenotus schomburgkii 3 Delma borea 1 Delma butleri 1 Delma nasuta 48 Delma tincta 32 Demansia psammophis 22 Diplodactylus ciliaris 48 Diplodactylus conspicillatus 87 Diplodactylus elderi 9 Diplodactylus stenodactylus 63 Diporiphora winneckei 200 Egernia inornata 289 Eremiascincus fasciolatus 274 Gehyra variegata 17 Heteronotia binoei 66 Lerista aericeps 53 Lerista labialis 5663 Lerista xanthura 14 Lialis burtonis 38 Lophognathus longirostris 20 Lucasium damaeum 20 greyii 487 Menetia maini 11 Moloch horridus 130 Morethia ruficauda 117 Nephrurus levis 416 Notoscincus ornatus 9

251

Appendix 5

Oxyuranus microlepidotus 6 Pogona vitticeps 123 Pseudechis australis 2 Pseudonaja modesta 16 Pseudonaja nuchalis 25 Pseudonaja textilis 2 Pygopus nigriceps 15 Ramphotyphlops endoterus 365 Ramphotyphlops nigrescens 1 Rhynchoedura ornata 153 Simoselaps fasciolatus 11 Suta punctata 1 Tiliqua multifasciata 26 Varanus brevicauda 508 Varanus eremius 232 Varanus gilleni 14 Varanus gouldii 281 58 17681

252

Appendix 5

Table A5.2: Description of the variables used in the structural equation models. The distribution is given for response variables only. Variables used follow Greenville et al. (2009), Greenville et al. (2014) and Chapters 4 and 5.

Variable Type Distribution Notes

Dasyurid Abundance count; 0–max Poisson Nine sites, ≤ 23 years

Reptiles Abundance count; 0–max Poisson Nine sites, ≤ 23 years

Mulgara Abundance count; 0–max Poisson Nine sites, ≤ 23 years

Rodents Abundance count; 0–max Poisson Nine sites, ≤ 23 years

Cat Abundance count; 0–max Poisson 2 years, pooled by sites

Fox Abundance count; 0–max Poisson 2 years, pooled by sites

Dingo Abundance count; 0–max Poisson 2 years, pooled by sites

Spinifex cover 0–max% Binomial Nine sites, ≤ 23 years

Spinifex seed Index; 0,1,2,3,4,5 Binomial Nine sites, ≤ 23 years

Site Categorical; 1–9 Nine sites, ≤ 23 years

Mean event size 2 Continuous Nine sites, ≤ 23 years months prior

8 months cumulative Continuous Nine sites, ≤ 23 years rainfall

Rain days Continuous Nine sites, ≤ 23 years

Years since wildfire Continuous, integer Greenville et al. (2009)

and NAFI (2013)

253

Appendix 5

Table A5.3: Parameter µ (mean) used to simulate the change in wildfire and rainfall due to climate change over 100 years, using 12 time steps. Simulated values were generated from a negative binominal distribution (rainfall: size (dispersion) = 1.57, n = 2000, and wildfire: size = 3.34, n = 2000; size was estimated from actual datasets for both rainfall and wildfire, as was µ for 2014).

Time step Time since wildfire 8 months cumulative rainfall

(µ) (µ)

2014 21 152

2023 20 176.27

2032 19 200.54

2041 18 224.81

2050 17 249.08

2059 16 273.35

2068 15 297.62

2077 14 321.89

2086 13 346.16

2095 12 370.43

2104 11 394.7

2114 10 418.97

254

Appendix 5

Figure A5.1: Eight month cumulative rainfall data for (a) actual dataset, (b) simulated dataset for current rainfall (n = 2000), and (c) simulated dataset for rainfall in 100 years. See Table A4.3 for parameters used to simulate data from a negative binomial distribution.

255

Appendix 5

Figure A5.2: Years since wildfire for data for (a) actual dataset, (b) simulated dataset for current rainfall (n = 2000), and (c) simulated dataset for wildfire in 100 years. See Table A4.3 for parameters used to simulate data from a negative binomial distribution.

References: Cogger, H. G. 2000. Reptiles and amphibians of Australia. 6th edition. Reed New Holland, Sydney, Australia. Greenville, A. C., C. R. Dickman, G. M. Wardle, and M. Letnic. 2009. The fire history of an arid grassland: the influence of antecedent rainfall and ENSO. International Journal of Wildland Fire 18: 631-639. Greenville, A. C., G. M. Wardle, B. Tamayo, and C. R. Dickman. 2014. Bottom-up and top- down processes interact to modify intraguild interactions in resource-pulse environments. Oecologia 175: 1349-1358. NAFI. 2013. North Australian Fire Information, accessed 12th December 2012, http://www.firenorth.org.au/nafi2/

256

Appendix 6

Appendix 6:

This appendix includes popular science, policy and opinion articles that I have written based

on work in this thesis. Some of these articles have been republished on science blogs and in

newsletters*. The articles, in order, from each chapter are:

Chapter 3: Greenville, A.C. (2014). After the floods come the rats. Biology News, 25.

School of Biological Sciences, University of Sydney. February 2014.

Chapter 6: Greenville, A.C. (2014). Top Dog: How Dingoes Save Native Animals.

Australasian Science, November 2014.

Chapter 6: Greenville, A.C. and Wardle, G.M. (2013). Demise of the dingo: the loss of a top predator would have consequences for native wildlife. ESA Hot Topics in Ecology,

Ecological Society of Australia.

Chapter 6: Greenville A.C. and Wardle G.M. (2013). Will we hunt dingoes to the brink like the Tasmanian Tiger? The Conversation, 21st November 2013.

*A complete list can be found at www.aarontheecolog.wordpress.com/publications/

257

VALE ALAN WALKER

Professor Norman Alan (Alan) Walker FAA, who joined the School of Biological Sciences in 1967 and retired as Challis Professor of Biology in 1993, passed away on 28 October, 2013. Alan made major contributions to plant biophysics and our understanding of transport across plant membranes. Early in his career, he was involved in the development of the technique still used today to measure the electrical potential difference across plant cell membranes. He went on to characterise the mechanism of transport across the cell membrane of all the major ions and compounds needed by plant cells. With collaborator Alex Hope, he developed a quantitative explanation for the electrical potential difference across the plant cell membrane and with Andrew Smith made the first measurements of cytoplasmic pH in plant cells. Alan served on the University Senate (1986-87) and as Head of the School Biological Sciences for two years. Alan leaves a legacy of plant scientists around the world who have been mentored and inspired by him. He was admired for his high intellectual rigour, integrity and insatiable curiosity. A very special person - he will be missed by his family, friends, colleagues and the field of plant biophysics.

Given the conspicuous nature of the rat, it is surprising that AFTER THE FLOODS COME we know little about its biology. Thus we set out to find out more. In our study published in the Greening of arid THE RATS Australia: new insights from extreme years, a special issue of Austral Ecology, we collated historic, media and anecdotal BY AARON GREENVILLE records to work out if we could predict when this rat may They ate the supplies belonging to Burke and Wills, and irrupt. We found there was an eighty-percent chance of a more recently chewed on the electrical wiring throughout rat plague following a flooding annual rainfall of 750 mm. To homesteads. They have even disrupted communication put that in perspective, central Australia has an average networks. A rare invasion of native rats that spreads out annual rainfall of 150-200 mm. across central Australia, holding siege to some of the But where do the rats come from in the first place? They remotest and driest regions of our country. are quite rare in the dry times and disappear from much Once every ten years or more conditions change in central of the desert. Are they holding out in one or multiple Australia and flooding rains fall across the desert. The hideaways? A 1972 study suggested a single refuge in the desert blooms and turns a vivid green – such a contrast Barkly Tablelands. To investigate this we set up live-trapping against the red sand and the deep blue sky. A mass seeding grids at twelve sites. If the rats have only one safe-house, event takes place as grasses pour out their life boats in then they would move in one direction from the north. hope of more rain. These are the conditions the long-haired But the rats came from the south! We also found greater rat, otherwise known as the plague rat (Rattus villosissimus), captures at trapping grids closer to drainage lines. This takes advantage of. suggested that other refuges exist and that drainage lines The long-haired rat feeds on seeds and vegetation, but also provide important corridors for dispersal. anything else they can get to. If camping in central Australia Even though the long-haired rat can be an inconvenience to this rat is known to eat all the paper labels off your tins and some, it has some interesting biology. Perhaps, in another even have a go at your swag – it doesn’t matter if you are ten years time we can find out some more. still trying to sleep in it! Originally published online at aarontheecolog.wordpress.com/

Photo by Aaron Greenville

5 Top Dog How Dingoes Save Native Animals

AARON GREENVILLE

Dingoes are considered a pest by land managers in Central Australia, but it turns out they are effective pest managers of feral cats and foxes – until the rains come.

othing quite gets your blood flowing like an in Oecologia , we found that while dingoes in the Simpson Desert encounter with a large predator, whether you’re limit the number of feral cats and foxes in dry times, an explo - locking eyes with a lion, even from the safety sion in the number of native rodent prey after flooding rains of a safari vehicle, or listening to the haunting weakens this relationship. howl of a dingo in remote central Australia. On the basis of this, we suggest it might make sense for graziers ONur instincts tell us to defend ourselves, be ready to get out of and conservationists to concentrate their control efforts on wet their way, to fight or flee. Not surprisingly, when we consider the seasons. At other times, simply deterring dingo attacks without deep-seated nature of these instincts, humans are responsible for killing the dingoes themselves should keep the smaller predators the persecution of large predators all around the world –and their in their place. populations are declining. However, recent research shows that large predators such as Cascading Effects of Top Predators lions, sharks and wolves play important roles in their ecosystems. Sitting at the top of the food chain, top predators can reduce Sitting at the top of the food chain allows them to cast their influ - populations of herbivores. These effects can then cascade through ence widely. the ecosystem. Such cascades have an important influence on The subtle but pervasive influence of top predators has become maintaining the diversity of many species. increasingly acknowledged through the study of ecosystem inter - For example, sea otters experienced large population declines actions, with important understanding emerging that top pred - after overhunting for the fur trade in the 18th and 19th centuries. ators can limit the populations of smaller “mesopredators”. For The removal of this predator produced an irruption in the sea example, the grey wolf in North America affects populations of urchin population, as this was a favoured meal for the otter. As a the smaller coyote. In Africa, lions and leopards can restrict popu - result, the urchin overgrazed the kelp forests of coastal ecosys - lations of the olive baboon. tems. When the fur industry ended, sea otter populations started When populations of these top predators are reduced, popu - to recover and they reduced urchin numbers. Consequently, kelp lations of the smaller predators increase. In turn, smaller prey increased in abundance again, and fish species that use the kelp also species come under increased predation pressure and can suffer. returned. Like wolves in North America, Australia’s largest terrestrial After the grey wolf was reintroduced into Yellowstone National predator, the dingo, is also the top dog. In a recent study published Park, it started to prey on elk. The reduction in the elk popula -

24 ||NOVEMBER 2014 During droughts, dingoes limit the abundance of the red fox and feral cat.

Credit: Bobby Tamayo tion led to an increase in woody plants like aspen, which the elk This begs the question: given that large predators have a regu - browsed upon. The surprising take-home message is that wolves latory role in suppressing the excesses of smaller predators elsewhere protect trees! in the world, does the dingo play a similar role in limiting popu - Such interactions up and down the food chain are among the lations of the red fox and feral cat in Central Australia? most fascinating and complex areas in ecology, and challenge our thinking about the world’s ecosystems. Predators Under Surveillance There has been a long debate in ecology about the importance of The Accusatory Finger the “top-down” effects of predators compared with “bottom-up” Australians like breaking records, but we hold one record for effects such as booms in resources and prey abundance. More which we should not be proud: we lead the world in mammalian recently, these two schools of thought acknowledged as important extinctions. A total of 29 species, or almost 10% of our mammals, influences on species populations. have become extinct within the past 226 years. The deserts of the world are a great place to study this inter - This is not simply a problem of the past: in 2009 we lost a action between bottom-up and top-down effects. Desert popu - small bat, the Christmas Island pipistrelle. Even more recently a lations generally tick along slowly, particularly in the climatically native rodent, the Bramble Cay melomys, may have squeaked variable environments in the heart of Australia, but every now goodbye. Given the high rate of endemism of our mammals, with and then, perhaps every 10 years or more, flooding rains deluge the so many species found nowhere else on Earth, this is a tragic loss desert and it blooms – and booms! In boom times the desert envi - for our natural heritage. ronment can experience 60-fold increases in the abundance of The accusatory finger apportioning blame for this shameful native rodent prey. We set out to discover if this massive increase record points squarely at two of our introduced predators: the in food changes or modifies the interactions between top and red fox and feral cat. The red fox is considered a threat to an esti - smaller predators. mated 76 native species, while the feral cat places at least 34 threat - To answer questions about the interplay between desert pred - ened native species at high risk of extinction. ators we needed to use a method to monitor populations of Past and predicted extinctions are not confined to the more dingoes, cats, foxes and their prey. We cannot be in the field all the populated areas of Australia. In fact, the remote corners of Central time, but a remote camera can. Australia have suffered the greatest loss of mammals. Remote camera traps are activated by movement. Once switched

NOVEMBER 2014 | | 25 A feral cat is caught out with a native rodent, the sandy inland mouse, by a remote camera trap set in the Simpson Desert. Introduced cats and foxes threaten many native Australian An introduced red fox walks past a remote camera trap species with extinction. one morning in the Simpson Desert.

on, they lie in wait 24 hours per day. When an animal walks past, the trap is triggered. The result is a photograph of the individual, complete with a record of the time, date, temperature and moon phase recorded. This information can be used to establish what – – time, and where, animals are most active. All this information is gathered without any human presence or interference with the animals. + We set our cameras up to capture images along the dirt tracks – + in an 8000 km 2 study area of the Simpson Desert. The cameras have so far collected photographs continuously for more than + + – – 2 years. What we have discovered is that the interaction between dingoes and the smaller introduced predators is more complex than first thought. During droughts, dingoes apply top-down pressure to limit the abundance of the smaller introduced red fox and feral cat. Fewer photographs of these smaller predators were recorded when the number of photographs of dingoes increase. In contrast, when populations of rodent prey boomed, this – relationship broke down for dingoes and foxes. Instead, all pred - ator populations increased due to the bottom-up influence of increased prey availability. + However, dingoes were still able to limit cat numbers in both – + drought and boom conditions, making this study one of the few + + examples of dingoes limiting feral cat abundance. The familiar – – dog chasing the neighbourhood cat seems to hold for wild popu - lations too. As prey populations declined and fell back to drought levels, the top-down interaction between dingoes and both of the smaller The interactions up and down the food chain are some of the most fascinating and complex areas in ecology. In predominantly introduced predators resumed. In contrast, there was no evidence dry times, dingoes in Central Australia can act as a pest species of an interaction between the red fox and feral cat. manager that limits numbers of the introduced red fox and feral By looking at when the photographs were taken, we could also cat. In turn, populations of native animals, such as rodents, can benefit. In boom times this relationship weakens, and this is when work out when dingoes, cats, foxes and their rodent prey were land managers should focus their fox and cat control measures. most active. The red fox and feral cat showed peak activity around The line width in the diagram represents the strength of the interaction (+ is a positive interaction; – is a negative interaction). the middle of the night – 2 hours before the peak activity of the Solid lines represent the direct effects of one species on another, dingo. Although all three predator species showed some activity and dashed lines represent the positive indirect effects of the top predator. Illustration: Alison Foster during all hours of darkness, this staggered activity time between

26 ||NOVEMBER 2014 A storm brings rain to the Simpson Desert and provides the trigger for a boom in wildlife. This increase in rodent food for predators can modify how dingoes, cats and foxes interact with each other. Photo: Aaron Greenville the smaller predators and the dingo would potentially reduce the likelihood of cross-species encounters. The red fox and feral cat were also most active when their rodent prey was active, highlighting again the threat that these introduced species can pose to our native wildlife. The smaller predators were most active at the same time, again suggesting there was no interference between the two species.

Implications of Our Study Our study provides support for the concept that the dingo acts as a top predator in Central Australia, limiting the activity or abun - dance of the introduced red fox and feral cat, and perhaps also changing the time when the smaller predators are most active. In turn, the presence of the dingo may help to protect populations The desert mouse ( Pseudomys desertor ) is one of the rare of native species, such as rodents. A predator protecting prey! rodents that occur in the Simpson Desert, but it is under threat from cat and fox predation. Photo: Aaron Greenville These findings have implications for how we manage intro - duced cats and foxes – and also dingoes. Dingoes are still perse - lethal control do exist. Guardian dogs efficiently protect stock cuted across Australia due to their impact on livestock. Land from dog attack, and returns on the cost of purchasing main - managers spend a great amount of time and money implementing taining guardians can be made within just 1 –3 years. However, if control programs to limit the populations of all three predators. guardian animals do not work in some locations, compensation However, our study suggests that the dingo is an unpaid pest to graziers for their losses may be appropriate. species manager that is at work every day. By leaving them alone, We need to get smarter about how we regulate predators on dingoes can help to control populations of cats and foxes for free. the land given what we know now about the role that they them - Land managers should therefore concentrate their pest control selves play in ecosystem regulation. In addition to the economic programs during the period after large rainfall events. Imagine benefits to land managers from fewer cats and foxes, dingoes also pooling resources that are spent every year on controlling cats play a role in regulating kangaroos, goats and pigs, which reduces and foxes in Central Australia and instead investing them every competition with stock for grass. 10 years or more? If the objective is to cap cat and fox numbers, Last, but not least, by erring from the age-old course of perse - then the dingo may be able to do the work for us during dry times. cuting the top predator, we get to keep one of Australia’s iconic There is no denying that dingoes will attack livestock, espe - animals and all of the ecological benefits that are being presently cially sheep, and industry-funded reports suggest this costs up to uncovered.

$60 million annually. Currently, lethal control methods are used Aaron Greenville is a PhD student with The University of Sydney’s Desert Ecology Research Group. The research paper in Oecologia is published online at http://link.springer.com/ to deter attacks on livestock from dingoes, but alternatives to article/10.1007/s00442-014-2977-8

NOVEMBER 2014 | | 27 HOT TOPICS IN ECOLOGY http://www.ecolsoc.org.au/hot-topics/demise-dingo Demise of the dingo The loss of a top predator would have consequences for native wildlife Aaron Greenville and A. Prof. Glenda Wardle, Desert Ecology Research Group, University of Sydney.

Dingoes are persecuted for similar reasons to the extinct Tasmanian tiger. Dingoes are controlled to reduce stock losses, in turn reducing the range of dingoes. The dingo provides positive conservation benefits to biodiversity, through suppression of feral cats, red foxes and over-abundant herbivores. Alternatives to lethal control and the dingo fence exist, with potential benefits to farmers and biodiversity alike.

Dingo ? killed and left on pole in 2013 (left) Thylacine last seen in wild in 1932 after five decades of lethal control (right). Dingo photograph by Aaron Greenville, thylacine photo taken ~1869, photographer unknown.

Because it threatens livestock, the native dingo Canis dingo is classified as a pest species. Barrier fencing and lethal methods are used to control the dingo, which is one of the few remaining mammalian top predators in Australia. Dingo management therefore parallels persecution of the Tasmanian tiger: a top-predator that was hunted for its alleged impact on livestock. Dingo persecution has reduced the species' distribution and disrupts its social structure, potentially accelerating the demise of this species through hybridisation with feral dogs. Whatever the genetic intermixing, it is clear that for over 4000 years dingoes have played a functional role in the Australian landscape. Top predators are now understood to be ecologically important in maintaining diversity. The loss or suppression of top predators can lead to additional extinctions through the food chain, and an increase of cats and foxes that negatively impact native wildlife in Australia. A review of the literature on the role of the dingo Ecological Society of Australia http://www.ecolsoc.org.au in the Australian environment identified the following key points:

Further information about this topic 1. Dingoes kill and compete with cats and foxes, and alter the foraging behaviour of feral cats. contact: This suppression of smaller predators can have net positive benefits for populations of Mr Aaron Greenville threatened species. [email protected] 02 9351 8577 2. Dingoes can control populations of herbivores. Kangaroo numbers increase with rainfall and, when present, dingoes limit their population through predation. This reduces grazing pressure Chair, Hot Topics Editorial Board on grasslands. Dr Don Driscoll 3. Lethal control of dingoes can disrupt the social structure of packs that would otherwise limit dispersal, hybridisation and attacks on stock.

4. Alternatives to lethal control and the dingo fence exist. For example, livestock guardian dogs can protect stock from dingo predation. This alternative is cost effective as it will provide a return on investment in 1 to 3 years. HOT TOPICS IN ECOLOGY http://www.ecolsoc.org.au/hot-topics/demise-dingo Demise of the dingo The loss of a top predator would have consequences for native wildlife Aaron Greenville and A. Prof. Glenda Wardle, Desert Ecology Research Group, University of Sydney.

Supporting Evidence

ID Title Type Location

6649 Allen B. L., Fleming P. J. S., Allen L. R., Engeman R. M., Ballard G. & Review paper Australia -wide Leung L. K. P. (2013) As clear as mud: A critical review of evidence for the ecological roles of Australian dingoes. Biological Conservation 159, 158-74.

6627 Brook L. A., Johnson C. N. & Ritchie E. G. (2012) Effects of pre-existing contrasts and North and central Australia predator control on behaviour of an apex predator and indirect manipulative experiment. consequences for mesopredator suppression. Journal of Applied Ecology 49, 1278-86.

6647 Caughley G., Grigg G., Caughley J. & Hill G. (1980) Does dingo Pre-existing contrasts ("natural SA, Qld and NSW border junction. predation control the densities of kangaroos and emus? Wildlife experiment") Research 7, 1-12.

6628 Choquenot D. & Forsyth D. M. (2013) Exploitation ecosystems and Modelling. Simulation. Arid Australia trophic cascades in non-equilibrium systems: pasture ? red kangaroo ? dingo interactions in arid Australia. Oikos 122, 1292- 306.

7049 Corbett L. (2001) The conservation status of the dingo Canis lupus correlational with no pre-determined Australia-wide dingo in Australia, with particular reference to New South Wales: gradient or contrast, threats to pure dingoes and potential solutions. In: A Symposium on the Dingo. Royal Zoological Society of New South Wales

6648 Eldridge S. R., Shakeshaft B. J. & Nano T. J. (2002) The impact of Manipulative experiment station, Lyndavale station and wild dog control on cattle, native and introduced herbivores and Umbearra station. introduced predators in central Australia. Unpublished report to the Bureau of Rural Sciences, Canberra.

6629 Glen A. S. & Dickman C. R. (2005) Complex interactions among Review paper Australia mammalian carnivores in Australia, and their implications for wildlife management. Biological Reviews 80, 387-401.

6630 Glen A. S., Dickman C. R., SoulÉ M. E. & Mackey B. G. (2007) Review paper Australia, excluding Tas. Evaluating the role of the dingo as a trophic regulator in Australian ecosystems. Austral Ecology 32, 492-501.

6632 Johnson C. N. & VanDerWal J. (2009) Evidence that dingoes limit correlational with no pre-determined Eastern NSW abundance of a mesopredator in eastern Australian forests. gradient or contrast Journal of Applied Ecology 46, 641-6.

6631 Johnson C. N., Isaac J. L. & Fisher D. O. (2007) Rarity of a top Pre-existing contrasts Australia, excluding Tas. predator triggers continent-wide collapse of mammal prey: dingoes and marsupials in Australia. Proceedings of the Royal Society B: Biological Sciences 274, 341-6.

6633 Kennedy M., Phillips B. L., Legge S., Murphy S. A. & Faulkner R. A. Manipulative experiment Central Kimberley of Western Australia (2012) Do dingoes suppress the activity of feral cats in northern and the Top End of the Northern Territory. Australia? Austral Ecology 37, 134-9.

6634 Letnic M. & Crowther M. S. (2013) Patterns in the abundance of Manipulative experiment Central Australia kangaroo populations in arid Australia are consistent with the Exploitation Ecosystems Hypothesis. Oikos 122, 761-9.

6636 Letnic M. & Dworjanyn S. A. (2011) Does a top predator reduce the Manipulative experiment predatory impact of an invasive mesopredator on an endangered rodent? Ecography 34, 827-35.

6638 Letnic M. & Koch F. (2010) Are dingoes a trophic regulator in arid Manipulative experiment Strzelecki Desert Australia? A comparison of mammal communities on either side of the dingo fence. Austral Ecology 35, 167-75.

6658 Letnic M., Crowther M. S. & Koch F. (2009) Does a top-predator Manipulative experiment Strzelecki Desert provide an endangered rodent with refuge from an invasive mesopredator? Anim. Conserv. 12, 302-12. ID Title Type Location

6637 Letnic M., Greenville A., Denny E., Dickman C. R., Tischler M., Pre-existing contrasts Qld, WA and Gordon C. & Koch F. (2011) Does a top predator suppress the abundance of an invasive mesopredator at a continental scale? Global Ecology and Biogeography 20, 343-53.

6639 Letnic M., Koch F., Gordon C., Crowther M. S. & Dickman C. R. Manipulative experiment Strzelecki Desert (2009) Keystone effects of an alien top-predator stem extinctions of native mammals. Proceedings of the Royal Society B: Biological Sciences 276, 3249-56.

6640 Letnic M., Ritchie E. G. & Dickman C. R. (2012) Top predators as review paper Australia wide biodiversity regulators: the dingo Canis lupus dingo as a case study. Biological Reviews 87, 390-413.

6641 Moseby K. E., Neilly H., Read J. L. & Crisp H. A. (2012) Interactions Manipulative experiment Roxby Downs, SA between a top order predator and exotic mesopredators in the Australian rangelands. International Journal of Ecology 2012.

6643 Ritchie E. G. & Johnson C. N. (2009) Predator interactions, review paper Global review, with emphasis on dingoes mesopredator release and biodiversity conservation. Ecology Letters 12, 982-98.

6642 Ritchie E. G., Elmhagen B., Glen A. S., Letnic M., Ludwig G. & review paper Global review, with emphasis on dingoes McDonald R. A. (2012) Ecosystem restoration with teeth: what role for predators? Trends in Ecology & Evolution 27, 265-71.

6644 van Bommel L. & Johnson C. N. (2012) Good dog! Using livestock survey Australia. Case study Qld. guardian dogs to protect livestock from predators in Australia?s extensive grazing systems. Wildlife Research 39, 220-9.

6645 Wallach A. D., Johnson C. N., Ritchie E. G. & O?Neill A. J. (2010) Manipulative experiment Arid Australia Predator control promotes invasive dominated ecological states. Ecology Letters 13, 1008-18.

6656 Wallach A.D, Murray B.R and O?Neill A.J.(2009). Can threatened correlation with no pre-determined Flinders Ranges, Ferries McDonald CP, species survive wherethe top predator is absent? Biological gradient or contrast Bakara CP and Innes NP. Conservation 142, 43-52.

6695 Wallach A.D, Murray B.R and O?Neill A.J.(2009). Can threatened correlation with no pre-determined Flinders Ranges, Ferries McDonald CP, species survive wherethe top predator is absent? Biological gradient or contrast Bakara CP and Innes NP. Conservation 142, 43-52.

6698 Wallach A.D, Murray B.R and O?Neill A.J.(2009). Can threatened correlation with no pre-determined Flinders Ranges, Ferries McDonald CP, species survive wherethe top predator is absent? Biological gradient or contrast Bakara CP and Innes NP. Conservation 142, 43-52.

6722 Wallach A.D, Murray B.R and O?Neill A.J.(2009). Can threatened correlation with no pre-determined Flinders Ranges, Ferries McDonald CP, species survive wherethe top predator is absent? Biological gradient or contrast Bakara CP and Innes NP. Conservation 142, 43-52.

6646 Wang Y. & Fisher D. O. (2012) Dingoes affect activity of feral cats, Correlational with no pre-determined Taunton National Park(11 626 ha) in but do not exclude them from the habitat of an endangered gradient or contrast central Queensland macropod. Wildlife Research 39, 611-20. Will we hunt dingoes to the brink like the Tasmanian tiger?

Edition: AU UK US Job Board Events Newsletter    Become an author Sign up as a reader Sign in

The Conversation 

Academic rigour, journalistic flair

Arts + Culture Business + Economy Education Environment + Energy Health + Medicine Politics + Society Science + Technology

Follow Topics Lima climate summit US-China climate deal World Parks Congress IPCC Fifth Assessment Report Indonesia Renewing Federalism

21 November 2013, 6.20am AEDT Will we hunt dingoes to the brink like the Tasmanian tiger?

AUTHORS

Aaron Greenville PhD student and Research Assistant, Desert Ecology Research Group at University of Sydney

Glenda Wardle Associate Professor in Ecology at University of Sydney

DISCLOSURE STATEMENT

Aaron Greenville receives funding from Australian Postgraduate Award and Paddy Pallin Science Grant funded by Humane Society International, Royal Zoological Society of NSW. A dead dingo in 2013 (left) and a Tasmanian tiger, last seen in the wild in 1932. Dingo Glenda Wardle receives funding from the photography by Aaron Greenville; a hunted thylacine in 1869, photographer unknown. Australian Research Council The last Tasmanian tiger died a lonely death in the Hobart REPUBLISH THIS ARTICLE Zoo in 1936, just 59 days after new state laws aimed at protecting it from extinction were passed in parliament. We believe in the free flow of But the warning bells about its likely demise had been pealing information. We use a Creative Commons license, for several decades before that protection came too late - and so you can republish our today we’re making many of the same deadly mistakes, only articles for free, online or in now it’s with dingoes. print.

Earlier this month the Queensland government announced Republish Provides funding as a Member of it would make it easier for farmers to put out poison baits for The Conversation AU. sydney.edu.au “wild dogs”. In Victoria, similar measures have already been taken. SHARE

Lethal methods of control have lethal consequences. It is time  Email

JOBS to rethink our approach in how we manage our wild predators.  Twitter 69

 Facebook Senior Lecturer (Social Pedagogy) Victoria University A deadly history lesson  LinkedIn

Graduate Services Coordinator Commonly known as Tasmanian tigers because of their Reddit University of Melbourne striped backs, thylacines were hunted due to the species Senior Research Fellow | Part Time alleged damage they were doing to the sheep industry in the Monash University  Sign in to Favourite

http://theconversation.com/will-we-hunt-dingoes-to-the-brink-like-the-tasmanian-tiger-19982[22/02/2015 12:21:31 PM] Will we hunt dingoes to the brink like the Tasmanian tiger?

state. However, the thylacine’s actual impact on the industry Lecturer Art In Public Space (Part  Comment Time Urgent Start) was likely to have been small. RMIT University

Postdoctoral Research Fellow In Instead, the species was made a scapegoat for poor TAGS Community Engaged Learning And Research management and the harshness of the Tasmanian Macquarie University environment, as early Europeans struggled implementing Dingoes, Victoria,  MORE JOBS foreign farming practises to the new world. Queensland, Farmers, sheep

RELATED ARTICLES UNIVERSITY OF SYDNEY The tiger [thylacine]… received a very bad character in EVENTS the Assembly yesterday; in fact, there appeared not to Sydney Ideas - HAL FOSTER: be one redeeming point in this animal. It was described Contemporary Art and Mimetic Excess — Sydney, New South Wales as cowardly, as stealing down on the sheep in the night

Sydney Ideas - Can You Afford to Live and want only killing many more than it could eat… All Where You Choose? — Sydney, New sheep owners in the House agreed that “something If you want to cut South Wales crime, you can’t ignore should be done,” as it was asserted that the tigers have the evidence Sydney Ideas - Spanish Literature — Sydney, New South Wales largely increased of late years. - The Mercury, October 1886. Sydney Ideas - A Precariat Charter: combating insecurity and inequality — Sydney, New South Wales

 MORE EVENTS Let’s move the world’s longest fence to settle the dingo debate

Queensland’s biggest economic challenge isn’t debt – it’s growth

Grainy footage is all we have left of the thylacine.

More than a century later, and it’s now the dingo in the firing line.

Labor’s first test: putting integrity before politics in Queensland Since 1990, the number of sheep shorn in Queensland has crashed 92 per cent, from over 21 million to less than 2 million. Although there have been rises and falls in the wool price and droughts have come and gone, it’s the dingoes that have been the last straw. - ABC Radio National, May 2013

An ancient predator vs modern farmers

Producing sheep is an incredibly tough business, with droughts, international competition and volatile markets for wool and meat – mostly factors that are well beyond the control of an individual farmer.

http://theconversation.com/will-we-hunt-dingoes-to-the-brink-like-the-tasmanian-tiger-19982[22/02/2015 12:21:31 PM] Will we hunt dingoes to the brink like the Tasmanian tiger?

Dingoes are seen as one of the few threats to livelihood that producers can fight back against. As a result, the dingo has experienced a severe range contraction since European settlement and there is mounting pressure to remove the dingo from the wild, despite dingoes calling Australia home for 4000 years.

Dingoes are now rare or absent across half of Australia due to intense control measures. While they are more common in other areas, we have seen how species populations can collapse quickly. For example, bounty records from Tasmania showed the thylacine population suddenly crashed in 1904-1910 due to hunting pressure from humans.

Will the dingo’s demise be like that of the thylacine? We simply do not know, but the social conditions and a rapidly changing environment mirror the story of the thylacine.

It’s true that dingoes have an impact on livestock. Estimates from industry-funded reports range from A$40 million to A$60 million, which include damage to livestock and cost of control measures.

And the emotional cost to farmers should not be underestimated. As authors, one of us has sheep farmers in the family, and knows the pride people gain from having a happy and healthy flock.

The choice is whether we want to follow the old colonial attitude of trying to conquer our environment, or find new and cheaper methods to live with our environment.

Dingoes and wild dogs

The issue of how to manage one of the few remaining mammalian top predators in Australia is further complicated by the suggestion that dingoes are not distinct from “wild dogs” due to interbreeding.

In eastern Australia dingo purity is low, but it is still high in many regions, such as central Australia.

But whether you call them dingoes or wild dogs, these predators work as unpaid pest species manager that works around the clock, effectively controlling feral cat and red fox numbers.

Even in eastern Australia, there is evidence that dingoes are fulfilling this role by reducing fox numbers.

Dingoes can also control kangaroo numbers, reducing grazing pressure. Reducing pests and grazing pressure are a win for farmers and conservation alike.

http://theconversation.com/will-we-hunt-dingoes-to-the-brink-like-the-tasmanian-tiger-19982[22/02/2015 12:21:31 PM] Will we hunt dingoes to the brink like the Tasmanian tiger?

Learning to live with dingoes

As CSIRO researchers suggested a decade ago, we need to get better at dealing with genetically ambiguous animals, such as those that could be classified as dingoes or wild dogs. Instead, they argued that better approach to conservation decisions would involve protecting animals based on their role in the environment, as well as their cultural value.

Traditionally, barrier fences and lethal control (such as poisoning) have been used as methods to reduce livestock losses from dingoes.

However, the costs of removing the dingo as our free pest species manager, and the impact of fences as barriers to other wildlife, need to be taken into account when assessing the true cost of maintaining these approaches.

Alternatives to lethal control do exist. Guardian dogs can protect stock from dog attack and have a return on investment between one to three years. Such cost-effective strategies can allow both the dingo and grazing to co-exist.

Over thousands of years, dingoes have played a functional

role in the Australian landscape and can provide benefits for SHARE farmers, traditional Indigenous owners and to the

conservation of native wildlife.  Email

 Twitter 69 It is time to learn how to live with the dingo. If not, we risk  Facebook eventually driving dingoes out of the wild and into lonely zoo enclosures, just like the thylacine.  LinkedIn Reddit

 Like us on Facebook  Sign up to our free daily newsletter Job Board

Find career opportunities  Follow us on Twitter Australia from intelligent institutions.

Community Company Contact Community standards Who we are Editorial [email protected] Republishing guidelines Our charter Support [email protected] Friends of The Conversation Our team Subscribe to our Newsletters  Job Board Our audience

 Events Our blog AU UK US

 Our feeds Partners and funders Contributing institutions  Donate Contact us

http://theconversation.com/will-we-hunt-dingoes-to-the-brink-like-the-tasmanian-tiger-19982[22/02/2015 12:21:31 PM]