<<

Ecology and Conservation Biology of the Baw Baw frosti (Anura: ): Distribution, Abundance, Autoecology and Demography

Gregory J. Hollis

Submitted in total fulfilment of the requirements of the degree of Doctor of Philosophy

January 2004

Department of Zoology University of

Abstract

The decline of populations around the world is a well documented phenomenon. The Philoria frosti belongs to a group of high-elevation, mountain-top in that have undergone recent population declines, but an understanding of the responsible agents is deficient or absent for most . The inability to diagnose agents of decline has mostly been attributed to a paucity of knowledge on the natural history of these species.

The discipline of conservation biology provided a scientific basis for commencing investigation into the decline of P. frosti. This thesis examines the pattern and extent of decline, and the autoecology and demography of the species, in order to provide a basis for evaluating conceivable decline-agents, and to establish a platform to commence diagnosis of the decline.

The results of comprehensive surveys confirm that the population of P. frosti has undergone a significant decline and contraction in range at sub-alpine elevations (> 1300 m), and may have also declined at lower, montane elevations (960 – 1300 m) where previously unknown populations were recorded on the south-western and north-eastern escarpment of the Baw Baw Plateau. The results of monitoring between 1993 – 2002 indicate a continuation of the decline of P. frosti at elevations above 1400 m, whilst populations between 960 and 1400 m appear to have remained relatively stable. Due to a lack of historical base-line data, it is not known if populations from montane elevations have declined to the same extent as at sub-alpine elevations.

The survey technique involving counts of calling males was investigated for its aptness in estimating and monitoring the abundance of P. frosti. The species rarity and cryptic nature, and its occupancy of a harsh climate in a landscape that is difficult to negotiate, excluded other potentially useable survey techniques. Participation by calling males during the breeding season was found to vary daily and seasonally. If not accounted for, this temporal variation can result in under- estimation of abundance. Extreme climatic conditions was also found to reduce the detectability of calling males during surveys. Retrospective power analyses indicate that sufficient effort was attained to detect changes in population size of less than 10% at all elevations examined, except for some subsets of data that comprised a small number of survey transects.

A thorough examination of the distribution and abundance of P. frosti showed that the highest density of calling males occurred on the south-western escarpment of the Baw Baw Plateau between 1300 and 1400 m, followed by populations between 960 and 1300 m. The lowest density of males occurred at elevations above 1400 m, and on the north-eastern escarpment between 1200 and 1400 m. The size of the adult male population was estimated to be 7000 individuals.

iii attributes from extant breeding locations of P. frosti were compared with those from randomly-selected locations chosen within the known domain of potential breeding habitat of the species. A significant difference was recorded between extant and random sites, as determined from attributes recorded at different scales. Particular biophysical and floristic attributes, and associated environmental gradients, were found to be correlated with the differences observed. Breeding located within topographically protected, cool, moist communities on the south- western escarpment of the Baw Baw Plateau appear to represent the core breeding habitat of the extant population. The results suggest that historical populations of P. frosti observed prior to this study were less selective in their preference for breeding habitat compared to extant population, following the decline of the species.

Examination of male calling behaviour over several breeding seasons revealed annual, seasonal and temporal differences in calling activity and participation rates. Additional variation in the call structure of the species is described and quantified. Climatic variation, due to differences in habitat structure and elevation, was shown to influence duration, timing, rate and variance of calling activity. Calling activity was also shown to occur within a relatively narrow climatic window, indicating that reproductive activity may be particularly sensitive to natural or anthropogenic influences.

Radio-tracking revealed that adult P. frosti are relatively sedentary over breeding and post- breeding seasons, but partake in a consistent pattern of movement from aquatic to adjacent, terrestrial habitats following oviposition (females), and at the conclusion of calling activity (males). Extent and probability of movement was shown to be influenced by weather conditions, with overall movement being confined to a narrow range of climatic conditions. Movement activity was found to mostly occur when external climate conditions (ambient temperature) mimicked those experienced by at sheltering sites (substratum temperature). These climatic constraints on movement and dispersal by P. frosti have implications for the management of land use within the habitat of the species.

Skeletochronological techniques were successfully used to determine longevity, maturation and growth in P. frosti individuals. The cryptic nature and rarity of the species precluded other methods as being suitable for acquiring this information. The species is relatively long-lived, and attains sexual maturity at an older age when compared to most other anurans. A skewed age-class distribution of adult samples from the extant population indicated population instability when compared to the binomial age-class distribution of museum samples examined. The dominance of younger-aged adults following the decline of the species suggests a decline-agent that has impacted on the post-metamorphic stage of the population.

Knowledge gained in this study has enabled identification of the most conceivable agents of decline for P. frosti. Information acquired on the ecological requirements of the species, and

iv subsequent inference of sensitivity to natural and anthropogenic influences, has contributed to progress towards the development of a conservation strategy for the species, as well as providing input into designing research programs to assess the potential impact of land management.

v

Declaration

This is to certify that i. The thesis comprises only my original work towards the PhD except where indicated in the preface, ii. Due acknowledgment has been made in the text to all other material used, iii. The thesis is less than 100,000 words in length exclusive of tables, maps, bibliographies and appendices.

------Gregory J. Hollis Department of Zoology University of Melbourne, Parkville

vii

Preface

This thesis is structured as a sequence of five chapters that present primary data (3 – 7), accompanied by chapters comprising general introduction (1), review of existing information on P. frosti and general methods (2) and synthesis and conclusions (8).

Population survey data examined from three years (1993 - 1995) in Chapter 3 were collected by myself and an assistant prior to candidature, the results from 1993 and 1994 being published in Hollis (1995). I was also assisted in the field during population surveys (1996 – 2002) by numerous people (names are included in acknowledgment section) for reasons of logistics, safety, and the narrow time-frame within which to conduct an annual survey of the species.

John Davies assisted in collection, identification and interpretation of botanical and soil data for analysis in Chapter 4, under my supervision. Ordination analyses in Chapter 4 were carried out in collaboration with Peter Minchin. Peter Minchin also modified the algorithm used to derive Bray- Curtis dissimilarity coefficients within DECODA software to incorporate more detailed examination of habitat profiles (referred to as ‘height-difference-weighted Bray-Curtis’). This modified dissimilarity coefficient is cited as Minchin (unpublished). Peter Minchin, John Davies and myself intend to publish a paper based on the contents of Chapter 4, with myself as senior co- author.

Diagrams depicting kernel distributions for home range analyses in Chapter 6 were constructed in collaboration with Rolf Willig.

The laboratory work, and some of the field work, on which Chapter 7 is based, was carried out jointly by Michael Scroggie and myself. Both myself and Michael Scroggie intend to publish a paper based on the contents of Chapter 7, with myself as senior co-author.

Research from this thesis has contributed to the publication of a paper (Osborne et al. 1999), cited elsewhere in this thesis.

ix

Acknowledgments

My sincere thanks goes to my supervisors Graeme Watson and Graeme Coulson for their encouragement, advice and guidance throughout this study.

I am also particularly grateful to friends and colleagues, Graeme Gillespie, John Davies, David Hunter, Michael Scroggie, Gerry Marantelli, Peter Robertson and Murray Littlejohn for their support, advice and stimulating discussions on amphibian ecology, conservation biology and botany over the duration of this study.

There are numerous friends, colleagues and helpers who assisted with various elements of the field work, often under extreme climatic conditions and on difficult-to-negotiate terrain. I wish to thank them all. Special thanks must go to John Silins, who provided invaluable assistance, technical advice and support over many years of the study, and John Davies for his assistance in the collection, identification and interpretation of botanical and soil data. Particular thanks must also go to David Black for his contribution and commitment to the survey monitoring program over the years, and David Hunter for his assistance during the early stages of the study. I also thank Paul Barber, Kylie Bawden, Julia Benjamin, Julia Bolton, Scott Breaden, Allan Caddy, Ryan Chick, Marien Davey, Angela Duffy, Sheree Fickling, Graeme Gillespie, Aaron Grigo, Jodelle La- Combe, Gerry Marantelli, Philip Marantelli, Luke Murphy, Kym Saunders, Natasha Schedvin, Simon Nicol, Donna Nunan, Michael Scroggie, Mark Venosta and Luke Woodford for their assistance during population surveys.

I am particularly grateful to Peter Minchin for making available his knowledge on the use of ordination procedures, and for his assistance in data analyses.

I received valuable statistical advice during various elements of this study from Mick Keough, Peter Minchin and Graeme Gillespie. Erran Seamen provided advice on the use of kernel analyses and James Gibb provided advice on power analyses.

Special thanks must go to Michael Scroggie for his expert-hand in conducting histological work on a large portion of tissue samples, and for conveying his knowledge during examination and analysis of skeletochronological data. Bruce Abaloz kindly allowed me to use facilities in his histology lab.

Graeme Gillespie provided knowledge on the use of materials for constructing harnesses for radio- tracking and general radio-tracking methodology.

xi Rolf Willig and Hans Van Elmpt provided valuable assistance in the use of Geographic Information Systems (GIS) to analyse data and construct maps and diagrams.

Several colleagues and friends commented on drafts of various chapters in this thesis. Graeme Watson and Graeme Coulson provided exhaustive corrections and comment on all drafts of each chapter. Graeme Gillespie, Michael Scroggie, John Davies and Peter Minchin provided comment on specific chapters.

Special thanks must go to Department of Sustainability and Environment (DSE) and Department of Primary Industries librarians, Carole Casey and Dianna Carr for their ongoing efforts in obtaining citations. Melanie Thomas provided assistance in the formatting of chapters.

The Baw Baw Frog Recovery Team provided valuable discussion and guidance in determining priority research actions for the Baw Baw Frog. The then Department of Natural Resources and Environment (DNRE), , kindly allowed me to utilise my research on the Baw Baw Frog for post-graduate study. I also thank DNRE for providing equipment and facilities. The Mt Baw Baw Alpine Resort kindly provided accommodation for myself and other personnel over the years that this study was undertaken. I also thank Stewart Galloway and Greg Boucher for their support at Mt Baw Baw.

This study could not have been undertaken without funding support from the then Australian Nature Conservation Agency, the then National Heritage Commission, Environment Australia and the DSE. The study was conducted under the following research and Ethics permits: 934/063, RP-93-014, 945/010, RP-94-108, 956/031, RP-95-158, 967/074, RP-96-178, NP978/068, RP-97-183, RP-10000421, RP-10000874, RP-10001597, AEEC125/95, AEEC97/002 and AEEC99/005.

xii Table of Contents

Abstract...... iii Declaration...... vii Preface...... ix Acknowledgments ...... xi Table of Contents ...... xiii List of Tables ...... xix List of Figures...... xxiii

CHAPTER 1 GENERAL INTRODUCTION 1.1 Crisis and Conservation Biology...... 1 1.2 Global Phenomenon of Amphibian Declines...... 3 1.2.1 Ecological Importance of Amphibians ...... 3 1.2.2 Causes of Amphibian Declines...... 4 1.2.2.1 High-elevation, Mountain-top Amphibians...... 5 1.2.2.2 Amphibians: Potential Decline Agents...... 8 1.2.3 Interpretation and Evaluation of Amphibian Declines ...... 10 1.3 The Baw Baw Frog (Philoria frosti) ...... 12 1.3.1 and Related Species...... 12 1.3.2 Coexisting Amphibians...... 14 1.3.3 Conservation Status ...... 14 1.4 Study Objectives ...... 15

CHAPTER 2 EXISTING KNOWLEDGE ON PHILORIA FROSTI, DESCRIPTION OF STUDY AREA AND GENERAL METHODS 2.1 Existing Knowledge on Philoria frosti ...... 21 2.1.1 Distribution and Abundance ...... 21 2.1.2 Macro and Micro-habitat ...... 23 2.1.3 Life History and Ecology...... 24 2.2 Study Area ...... 26 2.2.1 Climate...... 26 2.2.2 Geomorphology and Geology...... 26 2.2.3 Soils ...... 27

xiii 2.2.4 Vegetation ...... 28 2.2.5 Fire History ...... 29 2.2.6 Historical landuse...... 29 2.2.6.1 Aboriginal History...... 29 2.2.6.2 European History...... 30 2.2.7 Current Landuse...... 31 2.2.8 Sites of Biological and Ecological Significance ...... 31 2.3 Nomenclature...... 31 2.4 Statistical Analyses ...... 32

CHAPTER 3 PATTERNS OF DISTRIBUTION AND ABUNDANCE, AND AN ASSESSMENT OF COUNTS OF CALLING MALES AS A METHOD FOR MONITORING AND ESTIMATING POPULATION SIZE 3.1 Introduction ...... 35 3.2 Material and Methods ...... 37 3.2.1 Abundance and Distribution ...... 37 3.2.2 Survey Effort and Timing ...... 38 3.2.3 Seasonal Correction of Census Data ...... 39 3.2.4 Analysis of Census Data and Detection of Trends...... 41 3.2.4.1 Statistical Power ...... 42 3.2.5 Estimating Adult Male Population Size...... 43 3.3 Results ...... 44 3.3.1 Distribution ...... 44 3.3.2 Relative Abundance and Population Trends ...... 45 3.3.2.1 Sub-alpine Elevation (>1400 m)...... 45 3.3.2.2 Sub-alpine-montane Elevation (1300 - 1400 m)...... 45 3.3.2.3 Montane Elevation (960 - 1299 m)...... 45 3.3.3 Participation of Calling Males in 1993 ...... 46 3.3.4 Statistical Power and Effect Size ...... 46 3.3.5 Comparison of Raw and Seasonally-corrected Census Data ...... 46 3.3.6 Population Size of Adult Males ...... 46 3.4 Discussion ...... 47 3.4.1 Statistical Power...... 47 3.4.2 Distribution ...... 48 3.4.3 The Use of Counts of Calling Males for Monitoring and Estimating Population Size ...... 49 3.4.4 Population Trends and Conservation Status ...... 52

xiv CHAPTER 4 BREEDING HABITAT FLORISTICS AND STRUCTURE: IDENTIFICATION OF PREFERRED ATTRIBUTES AND ASSOCIATED ENVIRONMENTAL GRADIENTS 4.1 Introduction...... 83 4.2 Materials and Methods ...... 85 4.2.1 Sampling Design...... 85 4.2.2 Selection of Breeding and Random Sites...... 86 4.2.3 Habitat Attributes and Measurement ...... 87 4.2.3.1 Floristics...... 87 4.2.3.2 Structural Variables...... 87 4.2.4 Index of Breeding Habitat Importance ...... 88 4.2.5 Soil Sampling Limitations ...... 88 4.2.6 Data Analysis...... 89 4.2.6.1 Ordination of Floristic and Structural Data ...... 89 4.2.6.2 Selection of a Weighting Scheme for Comparing Quadrat Height Class Pairs...... 91 4.2.6.3 Ordination of Biophysical Data...... 91 4.2.6.4 Vector Fitting ...... 92 4.2.6.5 Analysis Programs...... 93 4.3 Results...... 93 4.3.1 Weighting Scheme for Height Class Comparisons...... 93 4.3.2 Ordination of Floristic and Structural Data from 10 x 10 m Scale Sites ...... 93 4.3.2.1 All Sites...... 93 4.3.2.2 Sub-alpine Sites...... 94 4.3.4 Ordination of Floristic and Structural Data from 1 x 1 m Scale Sites ...... 95 4.3.4.1 All Sites...... 95 4.3.4.2 Sub-alpine Sites...... 96 4.3.5 Ordination of Biophysical Data ...... 97 4.3.5.1 All Sites...... 97 4.3.5.2 Sub-alpine Sites...... 97 4.3.5.3 Montane Sites...... 98 4.4 Discussion...... 99 4.4.1 Preferred Breeding Habitat Attributes ...... 99 4.4.2 Historical Use of Breeding Habitat...... 101 4.4.3 Stable, Persistent and Moist Breeding Habitat ...... 103 4.4.4 Related Taxa ...... 105

CHAPTER 5 CALLING BEHAVIOUR: ANNUAL, SEASONAL AND TEMPORAL VARIATION, AND THE INFLUENCE OF ENVIRONMENTAL FACTORS 5.1 Introduction...... 121 5.2 Methods ...... 123 5.2.1 Automatic Recording Units ...... 123

xv 5.2.2 Selection of Breeding Site...... 124 5.2.3 Quantification of Acoustic Data...... 124 5.2.4 Acquisition of Weather and Micro-climatic Data...... 125 5.2.5 Data Analysis ...... 125 5.2.5.1 Annual and Seasonal Calling Activity...... 125 5.2.5.2 Diel Calling Activity ...... 126 5.2.5.3 Variation in Calling Activity Between Males...... 127 5.2.5.4 Frog Participation and Associations between Calling Rate and Calling Intensity.127 5.2.5.5 Associations Between Calling Activity and Weather...... 128 5.2.5.6 Thermal Dynamics of Micro-habitats...... 129 5.2.5.7 Duration of Calling Activity...... 130 5.2.5.8 Weather during Commencement, Completion and Peak of Calling Activity...... 130 5.3 Results ...... 131 5.3.1 Call Structure and Variance ...... 131 5.3.2 Annual and Seasonal Calling Activity, Duration and Male Participation.....132 5.3.3 Diel Calling Activity and Male Participation...... 133 5.3.4 Calling Activity, Weather and Micro-climate...... 134 5.4 Discussion ...... 136 5.4.1 Call Structure ...... 136 5.4.2 Annual and Seasonal Patterns of Calling Activity...... 137 5.4.3 Patterns of Diel Calling Activity...... 140 5.4.4 Habitat Variability...... 141

CHAPTER 6 BREEDING AND POST-BREEDING SEASON PATTERNS OF MOVEMENT, ACTIVITY AND HABITAT USE 6.1 Introduction ...... 175 6.2 Materials and Methods...... 177 6.2.1 Radio-tracking...... 177 6.2.1.1 Data Collection...... 178 6.2.2 Pitfall Trapping ...... 179 6.2.3 Breeding and Post-breeding Seasons ...... 180 6.2.4 Data Analyses ...... 180 6.2.4.1 Movement...... 180 6.2.4.2 Home Range ...... 181 6.2.4.3 Habitat and Micro-habitat Use...... 182 6.2.4.4 Activity Patterns ...... 183 6.2.4.5 Sheltering Site Selection...... 185 6.3 Results ...... 185 6.3.1 Radio-tracking...... 185 6.3.2 Breeding and Post-breeding Seasons ...... 186 6.3.3 Movement Rate...... 186 6.3.4 Macro and Micro-habitat Use ...... 187

xvi 6.3.5 Home Range ...... 188 6.3.6 Weather and Movement...... 188 6.3.7 Sheltering Sites ...... 190 6.3.8 Pitfall Trapping...... 190 6.4 Discussion...... 191 6.4.1 Study Limitations...... 191 6.4.2 Movement Patterns and Habitat Use...... 192 6.4.3 Weather Associations ...... 196

CHAPTER 7 LONGEVITY, MATURATION AND SEX-SPECIFIC GROWTH 7.1 Introduction...... 225 7.2 Methods ...... 227 7.2.1 Breeding Biology and Life History...... 227 7.2.2 Toe-clipping for Mark-recapture and Ageing...... 228 7.2.3 Histological Procedure...... 229 7.2.4 Age Determination...... 230 7.2.5 Validation ...... 231 7.2.6 Analysis of Growth...... 231 7.3 Results...... 232 7.3.1 Interpretation of Histology...... 232 7.3.2 Validation of Skeletochronolgy ...... 232 7.3.3 Longevity, Maturation and Age Structure ...... 233 7.3.4 Growth ...... 233 7.4 Discussion...... 234

CHAPTER 8 SYNTHESIS AND CONCLUSIONS 8.1 Distribution, Abundance and Pattern of Decline ...... 243 8.2 Factors that Predispose P. frosti to being Sensitive to Environmental Change...... 245 8.2.1 Headwater, Relictual Amphibians ...... 247 8.3 The Decline of P. frosti and Potential Causative Agents ...... 248 8.3.1 Increased UV-B Radiation...... 248 8.3.2 Pathogens...... 249 8.3.3 Climate Change...... 250 8.3.4 Natural Population Fluctuations and Weather Patterns ...... 251

xvii 8.3.6 Atmospheric Pollution ...... 252 8.3.7 Multiple and Interacting Factors ...... 252 8.4 Implications for Conservation...... 253 8.5 Recommendations for Management and Research ...... 255 8.5.1 Management in State Forest...... 255 8.5.2 Management in the and Mt Baw Baw Alpine Resort...... 257 8.6 Priority Research Directions ...... 257

REFERENCES...... 261

APPENDICES ...... 299

Appendix 3.1: Census data of calling males recorded at survey transects at different elevations between 1983 and 2002...... 299

Appendix 3.2: Seasonally-corrected census data of calling males (C prefix) and estimated level of participation (%) by calling males (P prefix) for survey transects censused between 1993 and 1999...... 309

Appendix 3.3: Date in which each survey transect was censused between 1993 and 2002. .317

Appendix 4.1: Results of fitting vectors of maximum correlation for environmental variables in NMDS ordinations...... 325

Appendix 4.2: Alphabetic list of plant taxa recorded in 1 x 1 m and 10 x 10 m quadrats, including generic groups where individuals were unidentifiable to species level during the study...... 341

Appendix 4.3: Matrix of taxa and structural attributes that best discriminated between random and breeding sites at 10 x 10 and 1 x 1 m scales from sub-alpine elevations, as identified using the step-wise, variant of ANOSIM procedure...... 345

Appendix 5.1: Distribution of seasonal and diel calling (calls/min, pulses/call/min) and growling (growls/min) data recorded at sub-alpine and montane breeding sites from 1994 - 1999...... 353

xviii List of Tables

CHAPTER 2

Table 2.1. Relationship between vegetation types described for the study area and habitats utilised by P. frosti, as reported by Malone (1985a)...... 33

CHAPTER 3

Table 3.1. Survey transects and census data of calling males used to examine for population trends between 1983 and 2002...... 55 Table 3.2. Relationship between abundance of calling males (R = raw census data, C = seasonally- corrected census data) and survey year for different monitoring programs (data subsets) at different elevation (S = sub-alpine, SM = sub-alpine-montane, M = montane)...... 57 Table 3.3. Details of survey transects used to derive a density of calling males at sub-alpine elevations (> 1400 m)...... 58 Table 3.4. Details of survey transects used to derive a density of calling males at sub-alpine- montane elevation (1300 – 1400 m)...... 60 Table 3.5. Details of survey transects used to derive a density of calling male at montane elevation in the south-west (sw) (960 – 1299 m) and north-east (ne) (1200 – 1299 m) regions...... 62 Table 3.6. Total stream length, density of calling males (/m of stream ± SE) and estimated population size of adult males within six geographic areas encompassing the distribution of Philoria frosti (see methods for description of areas)...... 67

CHAPTER 4

Table 4.1. List of environmental variables quantified at breeding and random sites, and details of their field measurement...... 107

CHAPTER 5

Table 5.1. Details of sites used to monitor calling activity of Philoria frosti on the Baw Baw Plateau, 1994 - 1999...... 144 Table 5.2. Results of Levene’s test for unequal variances on daily calling and growling rates recorded at 13:00 h over the duration of 1994 - 1999 breeding seasons...... 145 Table 5.3. Results of Levene’s test for unequal variances on daily calling and growling rates recorded over six time periods (01:00, 07:00, 10:00, 13:00, 15:00 and 20:00 h) during the calling-activity peak of 1994 and 1998 breeding seasons...... 145 Table 5.4. Results of the Kolmogorov-Smirnov two sample test on mean daily call rates (calls/min) for 1995 - 1999 breeding seasons...... 146 Table 5.5. Results of the Friedman test on mean daily calling and growling rates and mean calling intensity recorded at 13:00 h over the duration of 1994 - 1999 breeding seasons...... 147

xix Table 5.6. Results of the Kolmogorov-Smirnov two sample test on calling participation by males for 1995 - 1999 breeding seasons...... 148 Table 5.7. Results of one-way ANOVA and Kruskal-Wallis test on the relationship between male participation and mean calling and growling rates, and mean calling intensity, recorded at 13:00 h during the calling-activity peak of breeding seasons 1994 - 1999...... 148 Table 5.8. Results of Spearman rank correlation analysis between mean calling intensity, recorded at 13:00 h during the calling-activity peak in 1994 - 1999 breeding seasons, and mean calling rates and weather...... 149 Table 5.9. Results of the Friedman test on calling and growling rates, calling intensity and participation over six time periods (01:00, 07:00, 10:00, 13:00, 15:00 and 20:00 h) during the calling-activity peak of 1994 and 1998 breeding seasons...... 150 Table 5.10. Results of correlation analysis between weather and mean calling-activity rates (call and growl rates > 0) and frog participation (proportion of calling males), recorded at 13:00 h during the calling-activity peak in 1994 - 1999 breeding seasons...... 151 Table 5.11. Results of univariant logistic regression analysis on weather and calling activity (calls and growls) recorded during the peak of calling activity in 1994 - 1999 breeding seasons, in sub-alpine and montane habitat...... 152 Table 5.12. ANOVA, regression and ANCOVA (homogeneity of slopes test) analyses on ambient and substratum temperatures recorded at 13:00 h from sub-alpine and montane breeding sites between 6 October - 11 November, 1999...... 153 Table 5.13. Results of regression analysis for the relationship between mean daily ambient temperature and date for the period encompassing the 1994 - 1999 breeding seasons...... 154 Table 5.14. Regression-derived mean daily ambient and substratum temperatures for different stages (calling-activity peak, calling commencement and calling completion) of 1994 - 1999 breeding seasons...... 155

CHAPTER 6

Table 6.1. Location, sex, size and radio-tracking details for individual frogs radio-tracked between November 1995 and February 1999...... 200 Table 6.2. Movement rate, radio-tracking intensity and results of home range and asymptotic analyses for individual frogs radio-tracked between November 1995 and February 1999...... 201 Table 6.3. Results of univariant logistic regression analysis on weather variables (collapsed into categories) and movement (Ref' = reference [no movement], Resp = response [movement]) during the breeding and post-breeding seasons, and at sub-alpine and montane elevations...... 202

Table 6.4. Results of correlation analysis (Pearson-r and Spearman rank-rs) between weather variables and daily frog movement > 0 m (natural log transformed)...... 203 Table 6.5. Results of regression analysis between ambient temperature and temperature of frog sheltering sites (surface and substratum) during the breeding and post-breeding seasons and at sub-alpine and montane elevations...... 203

xx Table 6.6. Results of hypothesis test (Ho:β1=1) on regression coefficients derived for the relationship between ambient and substratum temperature recorded at frog sheltering sites during the breeding and post-breeding seasons and at sub-alpine and montane elevations...... 203

CHAPTER 7

Table 7.1. Parameter details of von Bertalanffy growth curves derived for male and female samples acquired from sub-alpine (≥ 1270 m) and montane (< 1270 m) elevation...... 238

xxi

List of Figures

CHAPTER 2

Fig. 2.1. Location of study area and associated land tenure...... 34

CHAPTER 3

Fig. 3.1 Distribution of areas (enclosed by blue lines) used to census calling males between 1993 and 2002 on the Baw Baw Plateau...... 69 Fig. 3.2 Distribution of drainage lines (highlighted in blue) used to census calling males between 1996 and 2002 on the south-western and north-eastern escarpments of the Baw Baw Plateau...... 71 Fig. 3.3. Geographic areas from which calling-male density and population estimates were derived...... 73 Fig. 3.4. Distribution of calling males recorded during surveys in 1983 and 1984 (Malone 1985a) and 1993 – 2002, Baw Baw Plateau and escarpment...... 75 Fig. 3.5. Mean proportion of calling males derived from fortnightly, repeated surveys conducted at four survey transects over the duration of the 1993 breeding season, as depicted by distance-weighted least squares smoothing (see methods)...... 77 Fig. 3.6. Relationship between the number of calling males and survey year for subsets of census data recorded between 1983 and 2002 at sub-alpine (> 1400 m) elevation...... 78 Fig. 3.7. Relationship between the number of calling males and survey year for subsets of census data recorded between 1983 and 2002 at sub-alpine-montane (1300 – 1400 m) elevation...... 79 Fig. 3.8. Relationship between the number of calling males and survey year for subsets of census data recorded between 1983 and 2002 at montane (960 – 1299 m) elevation...... 80 Fig. 3.9. Relationship between raw census data of calling males and the same data seasonally- corrected for variation in participation of calling males over the duration of 1993 – 1999 breeding seasons...... 81

CHAPTER 4

Fig. 4.1. Distribution of breeding sites (red) and random sites (black) sampled between 1996 and 1999, Baw Baw Plateau and adjacent escarpment...... 111 Fig. 4.2. Relationship between quadrat height class spacing and ANOSIM R statistic for floristic and structural data collected at 10 x 10 m and 1 x 1 m quadrat scales for all sites (a and c), and separately for only sub-alpine (> 1270 m) sites (b and d)...... 113 Fig. 4.3. Distribution of all 10 x 10 m scale sites within a 2-dimensional NMDS ordination using floristic and structural data, including vectors of maximum correlation for biophysical variables...... 114

xxiii Fig. 4.4. Distribution of sub-alpine 10 x 10 m scale sites within a 2-dimensional NMDS ordination using floristic and structural data, including vectors of maximum correlation for biophysical variables...... 114 Fig. 4.5. Distribution of all 1 x 1 m scale sites within a 3-dimensional NMDS ordination using floristic and structural data, including vectors of maximum correlation for biophysical variables for axes 1 vs 3 (a) and axes 1 vs 2 (b)...... 115 Fig. 4.6. Distribution of sub-alpine 1 x 1 m scale sites within a 3-dimensional NMDS ordination using floristic and structural data, including vectors of maximum correlation for biophysical variables for axes 1 vs 3 (a) and axes 1 vs 2 (b)...... 116 Fig. 4.7. Distribution of all sites within a 2-dimensional NMDS ordination using only biophysical data, including vectors of maximum correlation for variables...... 117 Fig. 4.8. Distribution of sub-alpine sites within a 3-dimensional NMDS ordination using only biophysical data, with vectors of maximum correlation for variables. (a) axes 1 v 3 and (b) axes 1 v 2...... 118 Fig. 4.9. Distribution of montane sites within a 2-dimensional NMDS ordination using only biophysical data, including vectors of maximum correlation for variables...... 119

CHAPTER 5

Fig. 5.1. Recording unit used to monitor calling activity by male Philoria frosti at breeding sites between 1994 and 1999, showing programable timer, cassette recorder and ammunition box...... 156 Fig. 5.2. Relationship between date and mean daily calling activity recorded at 13:00 h over the duration of 1994 - 1999 breeding seasons, as depicted by distance-weighted least squares smoothing...... 157 Fig. 5.3. Comparison of distribution curves (distance-weighted least squares) for mean daily call rates (calls/min) recorded over the duration of 1994 - 1999 breeding seasons...... 159 Fig. 5.4. Comparison of distribution curves for participation by calling males over the duration of 1994 - 1999 breeding seasons, as depicted by distance-weighted least squares smoothing...... 159 Fig. 5.5. Relationship between number of calling males and call rate (pulses/call/min ± SE)...... 160 Fig. 5.6. Relationship between number of calling males and mean growling rate (growls/min ± SE)...... 160 Fig. 5.7. Relationship between date and calling intensity (1 = slow, 2 = medium, 3 = fast, 4 = very fast) recorded at 13:00 h over the duration of 1994 - 1999 breeding seasons, as depicted by distance-weighted least squares smoothing...... 161 Fig. 5.8. Relationship between time and mean call rates (calls/min ± SE) recorded for six time periods (01:00, 07:00, 10:00, 13:00, 15:00, 20:00 h) during the calling-activity peak of 1994 and 1998 breeding seasons. 0 = 24:00, 1 = 01:00, 7 = 07:00 h, etc...... 161 Fig. 5.9. Relationship between time and mean call rates (pulses/call/min ± SE) recorded for six time periods (01:00, 07:00, 10:00, 13:00, 15:00, 20:00 h) during the calling-activity peak of 1994 and 1998 breeding seasons...... 162

xxiv Fig. 5.10. Relationship between time and mean calling participation (± SE) for six time periods (01:00, 07:00, 10:00, 13:00, 15:00, 20:00 h) during the calling-activity peak of 1994 and 1998 breeding seasons...... 162 Fig. 5.11. Mean temperature (± SE) (dotted lines) and mean calling intensity (± SE) (solid lines) (1 = slow, 2 = medium, 3 = fast, 4 = very fast) recorded from six time periods (01:00, 07:00, 10:00, 13:00, 15:00, 20:00 h) during the calling-activity peak of 1994 and 1998 breeding seasons...... 163 Fig. 5.12. Relationship between mean call rate > 0 (calls/min - square root transformed) and weather (ambient temperature and solar radiation) in sub-alpine habitat, for activity recorded at 13:00 h during the calling-activity peak of 1994 - 1999 breeding seasons...... 163 Fig. 5.13. Relationship between mean calling intensity (1 = slow, 2 = medium, 3 = fast, 4 = very fast), ambient temperature (oC) and relative humidity (%) in sub-alpine and montane habitat, as recorded at 13:00 h during the calling-activity peak of 1994 - 1999 breeding seasons...... 164 Fig. 5.14. Relationship between ambient temperature and the probability (with upper and lower bounds) of calling activity (excluding growling activity) occurring in sub-alpine habitat...... 165 Fig. 5.15. Relationship between ambient temperature and the probability (with upper and lower bounds) of growling activity occurring in montane habitat...... 165 Fig. 5.16. Relationship between relative humidity and the probability (with upper and lower bounds) of growling activity occurring in montane habitat...... 166 Fig. 5.17. Diel relationship between ambient and substratum temperature recorded at a sub-alpine (red regression line) and montane (blue regression line) breeding site, 6 October - 11 November, 1999...... 167 Fig. 5.18. Relationship between ambient and substratum temperature recorded at 13:00 h at a sub- alpine (hatched line) and montane (solid line) breeding site, 6 October - 11 November 1999...... 168 Fig. 5.19. Ambient and substratum temperature recorded at hourly intervals from a breeding site located in sub-alpine and montane habitat, 6 October - 11 November, 1999...... 168 Fig. 5.20. Relationship between breeding-season duration (length of calling-activity period) and mean breeding-season weather in sub-alpine habitat for years 1994 - 1999...... 169 Fig. 5.21. Estimated mean daily ambient and substratum temperature for breeding-season commencement and completion in sub-alpine habitat for years 1994 - 1999, as predicted from regression formulae...... 170 Fig. 5.22. Distribution of regression-derived mean daily ambient and substratum temperature recorded during the calling-activity peak (11 day period) of 1994 - 1999 breeding seasons in sub-alpine habitat...... 171 Fig. 5.23. Commencement and completion dates of 1994 - 1999 breeding seasons (horizontal bar with hairs) in sub-alpine habitat in relation to total fortnightly rainfall recorded at Village Flat (1470 m a.s.l.) within the Mt Baw Baw Alpine Resort...... 172

xxv Fig. 5.24. Mean monthly rainfall (± SE) derived from 1932 - 1995 records collected at Erica (40 m a.s.l.), located at the south-eastern end of the Baw Baw Plateau, and the months in which calling activity by Philoria frosti has been recorded...... 173

CHAPTER 6

Fig. 6.1. Adult Philoria frosti with radio-transmitter and harness (drawing: Kate Thompson)...... 204 Fig. 6.2. Duration of male calling activity recorded at breeding sites during 1995, 1997 and 1998 breeding seasons, as depicted by least squares smoothing curves...... 204 Fig. 6.3. Frequency distribution of daily movement recorded for males during the breeding season, and for males and females during the post-breeding season, over the duration of the study...... 205 Fig. 6.4. Distribution of male and female rates of movement recorded during the breeding and post-breeding seasons, and at sub-alpine and montane elevations...... 206 Fig. 6.5. Linear distance moved from site of capture over time by male and female frogs during the breeding and post-breeding seasons...... 207 Fig. 6.6. Accumulated distance moved from site of capture over time by male and female frogs during the breeding and post-breeding seasons...... 208 Fig. 6.7. Mean proportion of days spent by frogs in different macro and micro-habitat types during the breeding and post-breeding seasons, at sub-alpine and montane elevations...... 209 Fig. 6.8a. Utilisation distribution, habitat use and radio-fix locations for frog number 5 over the breeding and post-breeding seasons...... 210 Fig. 6.8b. Utilisation distribution, habitat use and radio-fix locations for frog number 20 over the breeding and post-breeding seasons...... 211 Fig. 6.8c. Utilisation distribution, habitat use and radio-fix locations for frog number 1 over the breeding and post-breeding seasons...... 212 Fig. 6.8d. Utilisation distribution, habitat use and radio-fix locations for frog number 28 over the breeding and post-breeding seasons...... 213 Fig. 6.8e. Utilisation distribution, habitat use and radio-fix locations for frog number 16 over the breeding and post-breeding seasons...... 214 Fig. 6.8f. Utilisation distribution, habitat use and radio-fix locations for frog number 15 over the breeding and post-breeding seasons...... 215 Fig. 6.9. Relationship between daily movement > 0 m (natural-log transformed) and mean daily relative humidity/ambient temperature ratio (natural log transformed) at sub-alpine elevations during the breeding and post-breeding seasons...... 216 Fig. 6.10. Relationship between daily movement > 0 m (natural-log transformed) by frogs and mean daily relative humidity/ambient temperature ratio (natural log transformed) at montane elevations during the breeding and post-breeding seasons...... 217 Fig. 6.11. Relationship between mean daily temperature and relative humidity and the probability (with upper and lower bounds) of frog movement occurring in montane habitat during the breeding season...... 218

xxvi Fig. 6.12. Relationship between mean daily solar radiation, relative humidity (natural-log transformed) and total daily rainfall (natural-log transformed + 1) and the probability (with upper and lower bounds) of frog movement occurring in montane habitat during the post-breeding season...... 219 Fig. 6.13. Relationship between daily frog movement > 0 (natural-log transformed) and the arrival of frontal weather bearing rainfall (natural-log transformed + 1) at sub-alpine and montane elevations during the breeding and post-breeding seasons...... 220 Fig. 6.14. Relationship between the arrival of rain-bearing frontal systems (natural-log transformed) and the probability (with upper and lower bounds) of frog movement occurring during the post-breeding season at sub-alpine and montane elevations...... 220 Fig. 6.15. Relationship between ambient, ground surface and substratum temperature recorded at frog sheltering sites during the breeding and post-breeding seasons at sub-alpine elevations...... 221 Fig. 6.16. Relationship between ambient, ground surface and substratum temperature recorded at frog sheltering sites during the breeding and post-breeding seasons at montane elevations...... 222 Fig. 6.17. Number and type of Philoria frosti captured in pitfall traps during spring and summer periods of the 1994/1995 and 1995/1996 seasons...... 200

CHAPTER 7

Fig. 7.1. Transverse section through the diaphyseal region of secondary phalanx of Philoria frosti aged six months (a) and 12.5 years (b)...... 239 Fig. 7.2. Frequency distribution of age classes for adult frogs from populations comprising (a) male (n = 86) and female (n = 22) museum specimens from sub-alpine elevations (> 1270 m), (b) male (n = 12) and female (n = 10) field samples from sub-alpine elevations, and (c) male (n = 30) and female (n = 5) field samples from montane elevations (≤ 1270 m)...... 240 Fig. 7.3. Relationship between age and size for male and female samples from sub-alpine and montane elevations...... 241

xxvii

Chapter 1

GENERAL INTRODUCTION

1.1 Biodiversity Crisis and Conservation Biology

Biological diversity or biodiversity is the variety of life – the different plants, and micro- organisms, the genes they contain and the ecosystems of which they form a part (Commonwealth of Australia 1992). Recognition of a significant acceleration in loss of biodiversity has recently attracted increasing scientific, government and popular attention (Primack 1993; Perlman and Adelson 1997). The unprecedented loss of biodiversity over a relatively short period has been attributed to a rapidly growing human population and its actions, and has been compared with several other periods during the geological past when huge numbers of species have vanished, leaving behind a greatly impoverished biota (Hunter 1996). Recognition of a ‘biodiversity crisis’ lead to the largest ever gathering of heads of state at the 1992 United Nations Conference on Environment and Development, or ‘Earth Summit’, in which 158 of the countries of the world signed the Convention on Biological Diversity (Perlman and Adelson 1997). In 1997, the Government of released its Biodiversity Strategy to supplement Australia’s National Strategy for the Conservation of Biodiversity and the global Convention (NRE 1999).

The development of conservation biology as a discipline was brought about by the need for a scientific basis for confronting the loss of biodiversity, making it in every sense of the word, a ‘crisis' discipline’ (Primack 1993; Hunter 1996; Meffe and Carroll 1997). Whilst traditional applied disciplines such as agriculture, forestry, wildlife management and fisheries biology were found not to be comprehensive enough to address the critical threats to biodiversity (Primack 1993), conservation biology combined the principles of ecology, biogeography, population genetics, economics, sociology, anthropology, philosophy and other theoretically-based disciplines (Hunter 1996; Meffe and Carroll 1997). Most ecological studies of populations in the past have been focused on abundant species (Krebs 1994), whilst the discipline of wildlife management has focused primarily on vertebrates that were of interest to hunters and anglers (Hunter 1996). Conservation biology differs fundamentally from other traditional natural resource sciences in that it places relatively greater emphasis on all forms of life, rather than forms that are considered to be economically valuable (Soulé 1985). Recognition of a rapid decline in biodiversity has focused the origin and evolution of conservation biology primarily on population decline, scarcity and extinction (Soulé and Wilcox 1980; Soulé 1985, 1986; Caughley 1994; Caughley and Gunn 1996).

1 Two distinct directions in conservation biology have been described: (1) that of the small- population paradigm; and (2) that of the declining-population paradigm (Caughley 1994; Caughley and Gunn 1996). The small population paradigm is largely theoretical, dealing primarily with risks in relation to stochastic influences on the existing dynamics of a population, including demographic and environmental stochasticity, inbreeding depression and genetic drift; how or why the population became small is not the issue. The declining-population paradigm deals primarily with why a population is declining or has declined to small numbers, what has caused the decline, and what can be done to reverse the decline (Caughley 1994; Caughley and Gunn 1996). Both paradigms, however, have their deficiencies, with the declining-population paradigm being characterised by abundant practical experience but a lack of theory, and the small-population paradigm being characterised by limited testing of its strong theoretical basis (Caughley and Gunn 1996).

The First International Conference on Conservation Biology in 1978 is usually recognised as the origin of the conservation biology discipline, followed by the establishment of the Society for Conservation Biology eight years later (Hunter 1996). In 1988, a convention held by the Society for Conservation Biology examined research priorities for the following decade, identifying five major areas (Orians and Soulé 2001): 1. A crash program of extensive surveys and mapping to identify areas critical for the protection of natural and genetic resources, because of their biotic diversity or high levels of endemism, or because of imminent destruction of critical or unusual habitats and/or biotas; 2. Establishment of a small number of research sites in the tropics to develop a coordinated program of comparative research on populations, communities, and ecosystems in relatively undisturbed and secure situations; 3. Studies at all spatial scales to assess the kinds, mechanisms, and magnitude of effects on ecological systems; 4. Enhanced support for research that focuses on the physiology, reproduction, behaviour, ecological interactions, and viability of individuals, populations, and species, especially with regards to species of critical ecological or economic importance; 5. Increased training for both basic scientists and natural resource managers, particularly in tropical developing countries.

A review of research accomplishments and future needs by the Society for Conservation Biology in 2000 identified that although progress had been made in undertaking each of these research priorities, they still remained a high priority (Orians and Soulé 2001). However, the review also recognised that there had been significant changes in the kinds and intensity of human activities over the past decade, and that there were now additional considerations to these research priorities. These considerations include an increased awareness by biologists about the complexity of ecological systems, and of the spatial and temporal scales and contexts in which they operate, and

2 the importance of involvement of conservation biologists in the political process (Orians and Soulé 2001).

1.2 Global Phenomenon of Amphibian Declines

The well documented phenomenon of global declines of amphibians over the past decade has made amphibians one of the better-known taxa associated with the biodiversity crisis. General concern about amphibians originated at the First World Congress of Herpetology in 1989, where it was recognised that amphibian declines and disappearances had commenced as early as the 1970s (e.g., Heyer et al. 1988; Osborne 1989; Ingram 1990; Watson et al. 1991), many in seemingly pristine habitats with no obvious reasons to expect them, suggesting a phenomenon separate from the overall biodiversity crisis (e.g., Barinaga 1990; Blaustein and Wake 1990; Phillips 1990). Following the Congress was an assemblage of reports indicating the widespread decline of amphibian populations, including some celebrated species such as the gastric brooding frogs (Rheobatrachus silus and R. vitellinus) (Czechura and Ingram 1990; McDonald 1990; Tyler 1991), the golden toad (Bufo periglenus) (Crump et al. 1992; Pounds and Crump 1994) and the harlequin frog (Atelopus varius) (Pounds and Crump 1994).

Since the First World Congress of Herpetology over a decade ago, reports of declining amphibian populations have become a focal issue in both the scientific and popular media (Davidson et al. 2001). As early as 1993, such reports encompassed five continents, and included 500 amphibian populations that were listed as either having declined or were of conservation concern (Vial and Saylor 1993). Australia alone has also contributed significantly to the number of reported amphibian declines and disappearances. As of 1999, declines in Australia had been recorded in 33 species (Campbell 1999, and references therein). Tyler (1997) identifies 27 amphibians at threat in Australia, and a further 14 species of concern but poorly understood. These Australian species are primarily from wet tropic and sub-tropic regions of eastern Queensland (e.g., Czechura and Ingram 1990; McDonald 1990; Covacevich and McDonald 1993; Ingram and McDonald 1993; Richards et al. 1993; Trenerry et al. 1994; Laurance 1996; Laurance et al. 1996), temperate regions of eastern Australia (e.g., Watson et al 1991; Mahony 1993; Gillespie and Hollis 1996; White and Pyke 1996; Gillespie and Hines 1999; Hines et al. 1999; Mahony 1999), and high-elevation, sub- alpine regions (Osborne 1989, 1990a; Hollis 1995; Osborne et al. 1999; Hunter et al. 1998).

1.2.1 Ecological Importance of Amphibians

An increasing public awareness of the ecological value of amphibians has also resulted from the recent phenomenon of global amphibian declines. Because amphibians have intimate contact with

3 many components of the environment in which they live, they are considered to be valuable gauges of environmental health or stress (Baringa 1990; Phillips 1990; Blaustein and Wake 1990, 1995; Blaustein 1994). The life history of amphibians, which in most cases comprises of an egg, larval and terrestrial stage, and the possession of highly permeable skins through which to respire, makes them particularly sensitive to many aquatic and terrestrial environments within their habitat (Blaustein and Wake 1990; Blaustein and Wake 1995; Alford and Richards 1999). The position of amphibians within the food chain, where the larvae of many species forage at the base of the detritus food web and adults are upper level consumers, also makes them sensitive to environmental change (Phillips 1990; Wyman 1990). Amphibians should also be good monitors of local environmental conditions because they are relatively sedentary for their entire life by comparison with other vertebrates (Blaustein and Wake 1995). There are, however, some biologists who suggest that there is no substantive evidence that amphibians are better bio- indicators of anthropogenic stresses than other taxa (e.g., Pechmann and Wilbur 1994).

Amphibians are also functionally important for nutrient cycling and ecosystem energy-flow in most freshwater and terrestrial habitats (Gehlbach and Kennedy 1978; Werner and McCune 1979; Beebee 1996). For example, the total biomass of amphibians in some environments has been estimated to equal that of all resident species of mammal combined, more than twice that of all avian species during the peak of breeding activity, and greater than the combined biomass of all species of fish (Burton and Likens 1975a,b; Gehlbach and Kennedy 1978; Beebee 1996). In Britain, two anuran species are approximately as numerous as humans (Beebee 1996). Amphibians are also ecologically important in that they are primary consumers, as larvae (Seale 1980; Alford 1999), and primary predators as adults (Murphy and Hall 1981; Blaustein and Wake 1990; Blaustein et al. 1994a), as well as being prey for other invertebrates and vertebrates (Blaustein and Wake 1990; Duellman and Trueb 1994). Because of the fundamental importance of amphibians to many ecosystems, global decline of amphibians could be detrimental to the maintenance and sustainability of ecosystems at a local and global scale, and to the long-term survivorship of individual species, including humans.

1.2.2 Causes of Amphibian Declines

Reviews of the range of causes of amphibian declines have become increasingly common in the amphibian literature (e.g., Beebee 1996; Alford and Richards 1999, Campbell 1999; Collins and Storfer 2003), with proposed explanations being almost as diverse as the amphibian species in jeopardy (Blaustein and Wake 1995). , over-exploitation and introduced predators have clearly been identified as the main causes of declines and extinctions of mammals, birds and fish (Soulé 1986; Meffe and Carroll 1997; Burgman and Lindenmayer 1998). Similarly, the loss, degradation or modification of habitat is also the most documented cause of amphibian decline (Vial and Saylor 1993; Blaustein 1994; Blaustein et al. 1994a; Pechmann and Wilbur 1994; Blaustein and Wake 1995; Pechmann and Wake 1997; Alford and Richards 1999). Other

4 anthropogenic factors that have been shown to be detrimental to amphibian populations include over-exploitation by humans (Oza 1990), predation by introduced fish (e.g., Bradford et al. 1993; Drost and Fellers 1996; Hecnar and M’Closkey 1996; Fisher and Shaffer 1996; Webb and Joss 1997; Gillespie and Hero 1999; Gillespie 2001a; Kats and Ferrer. 2003), road kills (e.g., Fahrig et al. 1995; Hels and Buchwald 2001), atmospheric pollution (e.g., Beebee et al. 1990; Letnic and Fox 1997), stream sedimentation (e.g., Corn and Bury 1989; Gillespie 2002), and the direct application of toxicants (e.g., Wilson and McCranie 1994; Davidson et al. 2001).

There are, however, documented amphibian declines where the causative factors are more subtle and difficult to identify. For declining amphibian taxa that occur in seemingly pristine environments, a phenomenon independent to the general biodiversity crisis has often been inferred (Barinaga 1990; Blaustein and Wake 1990; Phillips 1990). In the absence of long-term data on populations and climatic, or experimental research, many reports and studies have been able to only review or allude to what the causative factors may be (e.g., La Marca and Reinthaler 1991; Crump et al. 1992; Drost and Fellers 1996; Pechmann and Wake 1997; Woolbright 1997; Lips 1998; Carey et al. 1999; Stallard 2001; Santiago et al. 2003). Examples of potential causative factors that have been investigated for such amphibian declines include, increased ultraviolet-B radiation due to depletion of stratospheric ozone (e.g., Blaustein et al. 1994b; Blaustein et al. 1996, Blaustein et al. 1997; Hays et al. 1996; Anzalone et al. 1998; Davidson et al. 2001; Middleton et al. 2001), acidification (e.g., Bradford et al. 1994; Vertucci and Corn 1996), disease (e.g., Blaustein et al. 1994c; Laurance et al. 1996; Alford and Richards 1997; Berger et al. 1998; Carey et al. 1999), climate change and weather (e.g., Pechmann et al. 1991; Crump et al. 1992; Heyer et al. 1988; Osborne 1989; Pounds and Crump 1994; Drost and Fellers 1996; Laurance 1996; Alexander and Eischeid 2001; Davidson et al. 2001; Stallard 2001) and interactions between, or combinations of the preceding factors (e.g., Carey 1993; Pounds and Crump 1994; Kiesecker and Blaustein 1995; Long et al. 1995; Alexander and Eischeid 2001; Santiago et al. 2003).

1.2.2.1 High-elevation, Mountain-top Amphibians

Amphibians occupying high-elevation environments have contributed significantly to the number of reported cases of amphibian declines in apparent pristine environments (Blaustein and Wake 1990; Tyler 1991). Such declines were reported to have struck suddenly for species in both Australia and other continents (Wake and Morowitz 1990), with proposed causative factors not being related directly to anthropogenic influences (see examples above). On other continents, declines in high-elevation, montane species are also numerous and have occurred in a range of habitats (e.g., Corn and Fogleman 1984; Heyer et al. 1988; Weygoldt 1989; Bradford 1991; La Marca and Reinthaler 1991; Crump et al. 1992; Carey 1993; Fellers and Drost 1993; Vial and Saylor 1993; Pounds and Crump 1994; Drost and Fellers 1996; Lips 1998, 1999; Santiago et al. 2003). In Australia, species reported to have declined occur primarily in montane, mountain-top

5 localities in the wet tropics and sub-tropics of Queensland (e.g., Czechura and Ingram 1990; McDonald 1990; Richards et al. 1993; Trenerry et al. 1994; Laurance et al. 1996), on tablelands in New South Wales (Mahony 1996; Osborne et al. 1996; Hunter and Gillespie 1999; Mahony 1999), and in montane and sub-alpine environments in the Southern Highlands of New South Wales and Victoria (Osborne 1989, 1990a; Gillespie et al. 1995; Hollis 1995; Gillespie and Hollis 1996; Hunter et al. 1998; Osborne et al. 1999).

In Queensland, all 14 montane species of amphibian reported to have declined are known to breed in streams (lotic species) (see Laurance 1996). Richards et al. (1993) excluded water chemistry properties, forestry and mining, low rainfall in wet seasons and phylogeny as factors responsible, or linked to the decline of six wet tropic species, but noted that disturbance and/or predation by feral pigs could have contributed to some of the declines, but was not the sole causative factor. Laurance (1996) dismissed unusual weather as an explanation for the dramatic declines of montane frogs in Queensland, although Alford and Richards (1999) indicate that Laurance’s analysis is likely to have missed rainfall extremes by examining only seasonal rainfall totals, and that a re-examination of the data was warranted. Laurance et al. (1996) proposed that a rapidly- spreading epidemic disease was the most likely agent responsible for declines. However, this interpretation was challenged by Alford and Richards (1997) and Hero and Gillespie (1997) on statistical grounds, arguing that other explanations appeared equally likely. Alford and Richards (1999) report that the involvement of a virus in the decline of these Queensland montane amphibians has now been largely discounted, but that the possible involvement of other pathogenic agents has not.

In New South Wales, declines have been reported in three amphibians [Litoria castanea (Litoria flavipunctata of Courtice and Grigg 1975); Litoria booroolongensis and Pseudophryne bibronii] on the northern tablelands (Mahony 1996; Anstis et al. 1998; Gillespie and Hines 1999; Mahony 1999) and three (L. castanea, Litoria raniformis and Litoria aurea) on the southern tablelands (Osborne et al. 1996; Mahony 1999). Litoria booroolongensis appears to have declined from higher-elevation areas on the southern tablelands, but there is no evidence of a decline at lower elevation (Hunter and Gillespie 1999; Gillespie 2000). Anecdotal observations suggest that P. bibronii may have also declined on the southern tablelands (D. Hunter pers. comm.). Introduced fish (Gambusia) have been suggested as the likely cause of the decline of members of the bell frog group (L. castanea, L. raniformis and L. aurea) (e.g., White and Pyke 1996; Morgan and Buttemer 1996; Webb and Joss 1997), whilst introduced trout and anthropogenic disturbances to catchments and riparian habitats have been identified as threats to L. booroolongensis on the southern tablelands (Hunter and Gillespie 1999). Increased ultraviolet radiation, disease and drought have also been implicated as factors explaining the possible decline of these tablelands taxa (Mahony 1999).

6 Factors responsible for the decline of high-elevation amphibians on the Southern Highlands of New South Wales and Victoria currently remain unclear for most taxa. Based on a review of wildlife data base records, Gillespie et al. (1995) indicate that nine of 27 taxa from the Australian Alps were of particular conservation concern. Osborne et al. (1999) provided compelling evidence that three of these nine taxa (Litoria verreauxii alpina, Pseudophryne corroboree and Philoria frosti) have undergone serious populations declines and range contractions after 1970 (the data for P. frosti originate from this study), and that a fourth species (Pseudophryne pengilleyi) has declined at higher altitudes, but has remained common at lower, montane altitudes. An additional species not confined to high-elevation areas (Litoria spenceri) has also declined from areas known to be previously occupied (Gillespie and Hollis 1996), including one montane location (Gillespie and Hines 1999; Gillespie 2001b).

Unlike taxa from the wet tropics and sub-tropics of Queensland and tablelands of New South Wales, a feature of most declining taxa from the Australian Alps is that they breed mostly in lentic situations such as shallow pools and seepages. Only L. spenceri is lotic with respect to its breeding habitat, although one might consider it not to belong to the group of alpine taxa given that a large portion of its distribution is outside the Australia Alps (see Gillespie and Hollis 1996). There does not appear to be a phylogenetic relationship between taxa occurring in the alps that might explain their decline: there are two hylids (L. v. alpina and L. spenceri) and three myobatrachids (Pseudophryne corroboree, Pseudophryne pengilleyi and Philoria frosti). Only P. corroboree and P. pengilleyi are closely related (see Osborne and Norman 1991). There also does not appear to be a single aspect of the life history in common between the taxa (see Osborne et al. 1999 for comparison of L. v. alpina, Pseudophryne corroboree, Pseudophryne pengilleyi and Philoria frosti). Litoria spenceri also does not appear to have any unique life history attributes that would predispose it to population declines more so than the other species (see Gillespie 2001b). Although global phenomenon have not been excluded as factors that may have contributed to, or influenced the decline of L. spenceri, Gillespie (2001b) concluded that introduced trout, increased sediment load in streams and habitat disturbance are probably the primary agents resulting in the decline and current rarity of L. spenceri.

The causative agents responsible for the decline of L. v. alpina, Pseudophryne corroboree, Pseudophryne pengilleyi and Philoria frosti are unclear. Osborne et al. (1999) review factors that may have contributed to the decline of a number of amphibians in the Australian Alps, including P. frosti. The data on the population of P. frosti presented in their review are from this study, excluding an additional five years of monitoring from 1998 - 2002. Osborne et al. (1999) concluded that it was unlikely that natural population fluctuations, acid precipitation and long-term weather patterns have contributed to the decline of three species (Pseudophryne corroboree, Litoria verreauxii alpina and Philoria frosti), and possibly a fourth species (Pseudophryne pengilleyi), but that an introduced fungal pathogen, increased ultraviolet (UV-B) radiation and climate change required further investigation as potential causal agents. Osborne et al. (1999) also

7 identified an increase in the extent of declines at higher elevations in each species as a single common factor (see below for further details and discussion).

1.2.2.2 Australian Alps Amphibians: Potential Decline Agents

Increased UV-B Radiation

Increased ultraviolet radiation (UV-B) due to depletion of stratospheric ozone (Blumthaler and Ambach 1990; Kerr and McElroy 1993) has been investigated for its anatomical and physiological effects on amphibians (e.g., Worrest and Kimeldorf 1976; Grant and Licht 1995; Hays et al. 1996; Blaustein et al. 1997; Fite et al. 1998; Jablonski 1998; Blaustein et al. 2000; Ankley et al. 2002), and as a factor to interpret declines and population persistence (e.g., Blaustein et al. 1994b; Blaustein et al. 1996, Blaustein et al. 1999; Langhelle et al. 1999; Broomhall et al. 2000; Belden and Blaustein 2002a, 2002b; Peterson et al. 2002). Many declines that have been attributed to increased UV-B have occurred at high-elevation or temperate latitudes (e.g., Blaustein et al. 1994b; Broomhall et al. 2000). It has also been demonstrated that species with high levels of photolyase activity are more tolerant of UV-B radiation (e.g., Blaustein et al. 1994b; Blaustein et al. 1996; Blaustein et al. 1999; Smith et al. 2002). Species that are most likely to be susceptible to increased levels of UV-B radiation are those that occur at higher elevation where UV-B intensities are greater (Blumthaler and Ambach 1990), and those that expose themselves to sunlight during either the pre-metamorphic stages of development (embryos and larvae) or the post-metamorphic stage.

Broomhall et al. (2000) reported an impact of increased UV-B radiation by demonstrating increased embryonic mortality in Litoria verreauxii alpina and Crinia signifera with increasing altitude in the Australian Alps of New South Wales. They also showed that developing embryos of L. v. alpina, a declining species, were significantly more sensitive to UV-B than C. signifera, a non-declining species. Both species deposit their eggs in relatively shallow ponds of water, where potentially harmful intensities of UV-B can penetrate (e.g., Schindler et al. 1996). An in situ experiment investigating the response of larvae of C. signifera to four treatments of UV-B exposure (0%, 10%, 55%, 90%) at Mt Baw Baw in the Central Highlands of Victoria demonstrated a sublethal effect on development rate (Gosner Stage) but not on size, with reduced development being recorded for higher exposures of UV-B (Hollis 2002). Embryos of P. frosti were also subjected to the four treatments of UV-B, but results were equivocal, with 100% mortality occurring across all treatments (Hollis 2002).

In the vicinity of Mt Baw Baw, monitoring of UV-B radiation reaching the ground surface in three habitat types utilised by P. frosti showed that lower-elevation (1400 m), montane riparian thicket received significantly lower levels of UV-B radiation than sub-alpine wet heathland at higher elevation (1530 m), and that cool temperate rainforest, at the lowest elevation (1090 m), received

8 only very low levels of UV-B radiation (Hollis 2002). This pattern of UV-B intensity was considered to be the result of the greater canopy and understorey cover offered by montane riparian thicket and cool temperate rainforest (Hollis 2002), and corresponds with the pattern of decline for P. frosti noted by Osborne et al. (1999).

Pathogens

Little is known about diseases in amphibians (Alford and Richards 1999). Various disease agents have been implicated in the cause of mortality and declines in amphibians (e.g., Nyman 1986; Bradford 1991; Blaustein et al. 1994c; Lips 1999). Declines in some of Australia’s frogs occurring in the streams of montane rainforests in Queensland were initially attributed to a virulent pathogen (Trenerry et al. 1994; Laurance et al. 1996), but this causal agent has now been largely discounted (Alford and Richards 1999). Daszak et al. (2003) report that a number of pathogens have been associated with amphibian declines, but the clearest link so far exists for amphibian chytridiomycosis. Support for chytridomycete fungus (Batrachochytrium sp.) as a possible cause of declines in tropical and temperate amphibians in Australia has also increased (Osborne et al. 1999). Experimental studies and observation on ill and dead frogs from Queensland, and other countries (central and south America and Europe), have provided evidence that chytridiomycosis may be the proximal cause of mortality (Berger et al. 1998; Berger et al. 1999; Bosch et al. 2001; Lips et al. 2003; Muths et al. 2003).

In Australia, 46 amphibians have now been recorded infected with Batrachochytrium. These species range from the east coast, the Adelaide area in South Australia, south-west Western Australian, and the Kimberley of north Western Australia (Speare and Berger 2000). Berger et al. (1999) believe that Batrachochytrium may be a recently introduced pathogen based on the nature of recorded amphibian declines, although other possibilities are that the fungus has suddenly become more pathogenic, or that other environmental stresses have lowered the resistance of frogs to the fungus. It is also possible that interaction between pathogens and other factors may be involved (e.g., Kiesecker and Blaustein 1995). There is currently insufficient knowledge about the interaction between Batrachochytrium and its host to interpret its future impact on amphibian populations, but if uninfected areas still exist in Australia, then endemic species in those areas may be at high risk (Berger et al. 1999).

Climate Change

Global climate change is now widely accepted by researchers, policy-makers and conservationists (Hannah et al. 2002). Climate change, in the form of enhanced greenhouse, is considered a threat to high-elevation species (e.g., Busby 1988; Bennett et al. 1991; Brereton et al. 1995; Pounds et al. 1999). Although not included in their BIOCLIM analysis due to its very limited distribution, Bennett et al. (1991) predicted that the bioclimate of P. frosti would disappear with a rise in

9 temperature of 1 - 3 oC. Osborne et al. (1999) reported that south-eastern Australia has experienced a slight increase (less than 0.1 oC) in annual maximum temperatures since 1951. More detailed climatological examinations of weather data from Victorian weather stations have also revealed warming trends, both at ground level and in the upper atmosphere (Smith et al. 1999).

As temperature, along with moisture, is a major factor influencing reproduction in anurans, particularly temperate-zone species (Duellman and Trueb 1994), a shift to earlier spawning might be an expected response to climate warming in some species. Some recent studies examining the breeding phenology of amphibians in association with apparent warming climates have recorded earlier commencement of spawning activity (e.g., Beebee 1995; Gibbs and Breisch 2001). Amphibian reproduction is also considered as probably the most vulnerable of all terrestrial vertebrates to variation in precipitation (Carey and Alexander 2003). Changes to rainfall patterns due to climate change therefore has the potential to impact on amphibian populations. For example, Osborne et al. (1999) summarise the decline of P. corroboree with respect to the distribution of rainfall and elevation, noting a progressive decline of the species from the drier, lower-elevation areas to the wetter, higher-elevation areas of its distribution. A recent review of the strength of evidence linking amphibian declines with climate change (Carey and Alexander 2003) emphasises the requirement for further research to demonstrate direct casual relationships between amphibian declines and climate change.

Atmospheric Pollution

With their water-permeable skins and dependence on aquatic habitats for reproduction, the sensitivity of amphibians to their external environment (see 1.2.1) predisposes them to being vulnerable to pollutants. In the absence of obvious anthropogenic disturbances in areas where declines have occurred, acid deposition has been examined as a potential causal agent (e.g., Bradford et al. 1994; Vertucci and Corn 1996). Other pollutants, such as industrial fluoride and agricultural pesticides, have also been investigated for their impact on amphibians (e.g., Letnic and Fox 1997; Davidson et al. 2002). Although Osborne et al. (1999) reported that there was no evidence to date that acid precipitation was a problem in parts of the Australian Alps, they noted the absence of information on possible levels of contaminants and dust in the general region. A recently-completed 10-year study examining air quality in the (located to the south of the Baw Baw Plateau), where large quantities of sulphur dioxide and oxides of nitrogen were emitted, concluded that power stations in the valley were not contributing significantly to acidification (Holper 1995).

1.2.3 Interpretation and Evaluation of Amphibian Declines

Although most amphibian biologists now agree that many species and communities of amphibians have declined or are undergoing ecological collapse (Alford and Richards 1999; Davidson et al.

10 2001), debate persists over the existence, extent and timing of some reported declines, and how they should be investigated and interpreted (e.g., Pechmann et al. 1991; Blaustein 1994; Blaustein et al. 1994a; McCoy 1994; Pechmann and Wilbur 1994; Roush 1995; Pechmann and Wake 1997; Pounds et al. 1997; Alford and Richards 1999). For example, Pechmann and Wilbur (1994) argued that there was insufficient evidence to separate declines of amphibian populations in isolated, protected habitats from that of natural population fluctuations, and that recorded declines were not independent of the overall biodiversity crisis. They also stated that there was no evidence to substantiate the view that amphibians are good bio-indicators of environmental stress, which is often used to support the presence of a phenomenon independent of the general biodiversity crisis. Blaustein (1994) disagreed with Pechmann and Wilbur (1994) based primarily on interpretation of decline-reviews, the generalised use of the term ‘pristine’ (in reference to habitat), and the likelihood that amphibians are likely to be as good bioindicators of environmental stress as any other organisms. More recent studies suggest that amphibians have exhibited periods of global decline, but the timing of such declines remains in dispute (Houlahan et al. 2000; Alford et al. 2001). Regardless of these differences of opinion, there are some generally-excepted views among most amphibian researchers concerning the interpretation and evaluation of amphibian declines. These include:

1. That the phenomenon of global amphibian declines is part of the overall crisis in biodiversity, where habitat destruction or modification are the single most important cause of declines of species (Blaustein 1994; Pechmann and Wilbur 1994; Houlahan et al. 2000; Alford et al. 2001); 2. That along with other taxa, amphibians are good bioindicators of environmental stress, but may vary between species and types of stress (Blaustein 1994; Storfer 2003); 3. That there is currently not enough long-term data on amphibian populations to assess the overall significance of amphibian declines (Blaustein 1994; Pechmann and Wilbur 1994; Alford and Richards 1999); 4. That there is a requirement for the understanding of the autoecology and dynamics of amphibian populations through integration of studies on, and within, metapopulations to effectively assess declines (Blaustein 1994; Pechmann and Wilbur 1994; Alford and Richards 1999); 5. That long-term monitoring of amphibians (and other taxa), in association with the development and testing of appropriate null and behavioural hypotheses, is required to evaluate declines (Pechmann et al. 1991; Blaustein 1994; McCoy 1994; Pechmann and Wilbur 1994; Alford and Richards 1999); 6. That the kind and amount of evidence required to statistically evaluate the presence or absence of a decline should be examined for a particular magnitude (effect size), and over selected time and spatial scales (e.g., McCoy 1994; Pechmann and Wilbur 1994; Storfer 2003).

11 In Australia, information used to assess the conservation status of amphibians has been based primarily on incidental reports, presence/absence data, or limited survey data, rather than on rigorous assessment using long-term census data (e.g., see Tyler 1997). There are few studies on the Australian frog fauna that have covered the entire, known geographic or environmental range of a taxon (e.g., Malone 1985a; Osborne 1989; Wardell-Johnson and Roberts 1993; Gillespie and Hollis 1996; Roberts et al. 1997; Parris 2001, 2002), or that have involved detailed study on the ecology and conservation biology of species (e.g., Humphries 1979; Odendaal et al. 1982; Odendaal and Bull 1983; Macnally 1985; Trenerry 1988; Osborne 1990b; Driscoll 1996; Gillespie 2001b). This paucity of information on the Australian frog fauna serves to highlight: (1) the lack of available resources to undertake studies on amphibians; (2) the difficulty in assessing the conservation status of most species; and (3) the difficulty in attempting to identify threatening processes or causative agents responsible for declines.

1.3 The Baw Baw Frog (Philoria frosti)

1.3.1 Taxonomy and Related Species

Philoria frosti was discovered in the central highlands of Victoria at Mt Baw Baw in 1898 by the naturalist, Charles Frost. At this time, Frost retrieved five freshly-regurgitated specimens from a captured , and shortly afterwards, collected two more live specimens near Mt Baw Baw (Spencer 1901). The species has it closest affinities with three distant, smaller frogs located in north-eastern New South Wales and south-eastern Queensland (Philoria sphagnicola, Philoria loveridgei and Philoria kundagungan), although Moore (1961) originally placed the two known species, P. sphagnicola and P. loveridgei, in a new genus (Kyarranus) without defining generic differences (Cogger 2000). In this thesis, I have followed the taxonomy of Cogger (2000) who adopts the genus Philoria. However, others retain the genus Kyarranus for species in north-eastern New South Wales and south-eastern Queensland on the basis that they are smaller and have smooth skins, compared to the raised and glandular skin of Philoria (Tyler 1992), and that there are osteological features (particularly of the pelvis) that separate them from the southern frosti (Barker et al. 1995). Barker et al. (1995) noted that both Philoria and Kyarranus have been found in fossil sites in north-west Queensland aged 20 – 30 million years. The fossil records also indicated that Kyarranus was once distributed far more widely compared with the geographically- isolated species that are currently present, and that the presence of Philoria in eastern Victoria suggests that a common ancestor may have been present in the intermediate area (Barker et al. 1995). Tyler (1992) described members of Kyarranus as relics, with their ancestors having a more continuous distribution along the at a time when the whole of Australia had a wetter climate.

12 Recent studies of allozyme and mitachondrial nuleotide sequence variation have revealed that there are more species present in north-eastern New South Wales and south-eastern Queensland than the three that are currently recognised (Knowles 1994; Mahony and Knowles 1994; Knowles et al. in press). Formal recognition of these species is pending a taxonomic revision of the group. Despite survey work conducted in similar environments to that which is used by P. frosti, no species closely related to P. frosti have been discovered elsewhere in Victoria (e.g., Malone 1985a, this study). Morphological descriptions of frogs and larvae of recognised members of the genus Philoria and/or Kyarranus can be found in Littlejohn (1963), Barker et al. (1995), Cogger (2000) and Anstis (2002).

All members of the genus Philoria have a similar life history and occupy similar habitats and environments. All species occupy mountain or plateau environments where they utilise high- elevation forests (rainforest or wet forest) or sub-alpine vegetation for breeding purposes (see Littlejohn 1963; Ingram and Corben 1975; Anstis 1981, 2002; Malone 1985a; de Bavay 1993; Barker et al. 1995; Hines et al. 1999). In late spring to early summer, all lay a relatively small clutch of large, unpigmented eggs within a foam nest, in water-filled or saturated cavities beneath rocks, logs or vegetation. Larvae are non-feeding, relying upon a residual yolk reserve for nutrition, and remain mostly at the site of oviposition to develop through to metamorphosis. Ingram and Corben (1975), Anstis (1981, 2002), Malone (1985a), de Bavay (1993) and Seymour et al. (1995) provide more details on embryos, larvae and oviposition for these species. See below for further details on P. frosti.

The confinement of taxon like Philoria to high-elevation, mountain or plateau environments predisposes them to rarity because they have relatively restricted distributions and narrow ecological requirements, and may be found in small populations (e.g., Malone 1985a; Primack 1993). All species from north-eastern New South Wales and south-eastern Queensland are considered to be rare (Hines et al. 1999), and P. frosti is listed as endangered (Tyler 1997) or critically endangered (VDSE 2003). Hines et al. (1999) report that there are no documented declines in species assigned to the genus Kyarranus, but of six presumed species, two are known from less than ten sites each. This suggests that there may be a lack of population data on some of these species to adequately assess their conservation status, or potential population declines. Although embryonic and larval development are well documented in recognised members of Philoria, or Kyarranus (see references above), and some data are available on diet (Webb 1989; Hollis 2002), very little is known about the autoecology, population dynamics and demography of these species (Tyler 1997; Hines et al. 1999). See further details on the ecology and life history of P. frosti in Chapter 2.

13 1.3.2 Coexisting Amphibians

Philoria frosti coexists with four other amphibians: the Alpine Tree Frog (Litoria verreauxii alpina), Southern Brown Tree Frog (Litoria ewingii), Common Froglet (Crinia signifera) and Southern Smooth Froglet (Geocrinia victoriana). The ecological relationships between these four species and P. frosti are largely unknown. The breeding habitat of P. frosti appears different from that of the other species, P. frosti utilising seepages on slopes and in gullies, whilst L. v. alpina, C. signifera and G. victoriana utilise pools or pockets of accumulated water occurring in the basins of sub-alpine wet heathland or in creek lines. Litoria ewingii occurs at intermediate altitudes (800 - 1300 m a.s.l.) on the Baw Baw Plateau, where it hybridises with L. v. alpina (Watson et al. 1985). The timing of breeding seasons of L. v. alpina, C. signifera and G. victoriana has been described as sequential, with L. v. alpina commencing breeding at the onset of the snow thaw, followed by the C. signifera, whilst G. victoriana breeds in autumn and has over-wintering larvae (Malone 1985b).

There are some quantitative and qualitative data available to examine the status of populations of the four amphibians that coexist with P. frosti on the Baw Baw Plateau. No L. v. alpina have been recorded during surveys undertaken to monitor the population of P. frosti between 1993 and 1997 (Osborne et al. 1999). This observation suggests a dramatic decline in the L. v. alpina population, particularly given that the species was observed to be extremely common during surveys for P. frosti by Malone (1985a) in 1983 and 1984 (B. Malone pers. comm.). The decline of L. v. alpina also appears to be concomitant with the observed decline of P. frosti. According to Gillespie et al. (1995), L. v. alpina was considered abundant throughout its distribution prior to the mid 1980s, but now appears to have declined at many localities, particularly those at higher altitudes. Litoria v. alpina is now listed as a threatened taxon (Tyler 1997; VDSE 2003). In contrast to the decline of L. v. alpina, C. signifera continues to be relatively common on the Baw Baw Plateau (G. Hollis unpublished data). Similarly, the observed presence of G. victoriana larvae in many ponds across the plateau during surveys for P. frosti also suggests that this species is also breeding successfully. However, anecdotal observations of calling activity by G. victoriana during late summer and autumn in 1996, 1997 and 1998 suggests the presence of only small breeding aggregations (G. Hollis pers. obs.).

1.3.3 Conservation Status

In 1983, the then Department of Conservation, Forests and Lands commissioned the Australian Biological Research Group Pty. Ltd. to undertake a study on the ecology of P. frosti (see Malone 1985a). This was due to assessments of status and future prospects of survival that concluded that the species warranted a high priority ranking of Category B – Vulnerable among threatened species of wildlife (IUCN Red Data Book 1979) (Ahern 1982). The allocation of P. frosti to this

14 threat category was due to several reasons: (1) its great scientific interest; (2) its highly restricted distribution; (3) the threat of habitat disturbance and destruction; and (4) the lack of basic ecological information on the species (Malone 1985a). Subsequent listings of the conservation status of P. frosti following the study of Malone (1985a) retained the Vulnerable threat category, including Schedule 1 of the Victorian Flora and Fauna Guarantee Act 1988 (CNR 1993); Schedule 1 of Endangered Species Protection Act 1992; and ANZECC (1991, 1995). Other assessments of the conservation status of P. frosti at the time presumed the species to be secure (e.g., Tyler 1992).

Based on the results of population surveys for P. frosti between 1993 and 1997 (reported on by Hollis (1995) and Osborne et al. (1999), and a component of this thesis), an adjustment to the conservation status of P. frosti was initiated. This adjustment was based initially on criteria outlined in the 1984 IUCN Red List of Threatened Species that examine the extent and magnitude of declines of species, and involved upgrading P. frosti from the Vulnerable threat category to Endangered (CNR 1995; Tyler 1997). The threat category of P. frosti in Victoria was further upgraded to Critically Endangered based on criteria in the 1994 IUCN Red List of Threatened Species (DNRE 2000).

1.4 Study Objectives

Because P. frosti displays all attributes that predispose species to rarity (see 1.3.1), and that a decline in population numbers and contraction in distribution range has recently occurred, there is an urgent need to understand the threats to this endemic frog, and the nature of its decline. Additional support is given to this degree of urgency by the relatively high frequency of reported amphibian declines from other high-elevation environments (see 1.2.2.1), and that high-elevation amphibians possess life-history characteristics that predispose them to being more vulnerable to the process of extinction (Morrison and Hero 2003). Unfortunately, like many other amphibian populations that have undergone declines, attempts to understand the extent and nature of the decline of P. frosti have been hampered by a lack of knowledge about current distribution and abundance, autoecology, and population dynamics and demography. This lack of knowledge has limited the capacity to develop sound management measures to ensure the long-term conservation of P. frosti, as well as the development and testing of appropriate null hypotheses required to evaluate its decline.

This thesis is based on the first four of five principles outlined for the process of Endangered Species Recovery Analysis (Caughley and Dunn 1996):

1. Confirm that the species is presently in decline, or that previously it was more widely distributed or more abundant.

15 2. Study the species' natural history for knowledge of its ecology, context, and status. 3. List all conceivable agents of decline when background knowledge is adequate. 4. Assess conceivable agents by measuring each where the species now is and where the species used to be in time or space. Any contrasting patterns identify a putative agent of decline and thus a hypothesis to be tested.

In 1993, an Action Statement was completed for P. frosti by the then Victorian Department of Conservation and Natural Resources, outlining actions to be undertaken to ensure the species long- term survival (CNR 1993). Unfortunately, this statement was published prior to obtaining knowledge about the decline of P. frosti, and it therefore lacked specific actions to address the decline. In 1996, a national Recovery Program was instigated to investigate the ecology and decline of P. frosti, with the long-term aim of using the gained knowledge to recover the species from the threat of extinction. The details and scheduling of this program are outlined in the Recovery Plan for the species (Hollis 1997). The objectives identified in the Recovery Plan were formulated by a referral group comprising amphibian biologists, ecologists and relevant land managers, known as the Recovery Team. These objectives were subsequently adopted for P. frosti in the Australian Frog Action Plan (Tyler 1997).

The Recovery Plan for P. frosti presents various research and management actions developed by the Recovery Team to contribute to, or achieve its objectives. The research actions were primarily developed to acquire the ecological information on P. frosti and its habitat needed to interpret the nature of the decline and contraction in range of the species, whilst the management actions were developed to identify and minimise known potential threats associated with land management, recreational use, and exotic flora and fauna within the Baw Baw National Park and Mt Baw Baw Alpine Resort. The Recovery Plan also contains criteria, against which progress towards meeting the objectives was to be measured.

A number of the objectives, criteria and research actions identified in the Recovery Plan for P. frosti form the basis of this thesis:

Objectives

• To gain an understanding of those aspects of the biology and ecology of the species that will enable effective management of the population. • To determine reasons for the observed decline across the species geographic range.

Criteria

• Secure population by monitoring population trends, distribution and habitat using rigorous experimental design.

16 • Obtain knowledge of those aspects of population dynamics and demography necessary to understand the population decline: (a) extent of habitat use below 1300 m (b) longevity and age structure of population (c) breeding habitat (d) patterns of movement.

Attempts to identify factors responsible for amphibian declines have been confounded by an inability to separate apparent declines from natural population fluctuations (Pechmann et al. 1991; Pechmann and Wilbur 1994). Recognition that a lack of long-term data on populations, weather and climate patterns and low statistical power, are largely responsible for this problem (Blaustein 1994; McCoy 1994; Pechmann and Wilbur 1994; Reed and Blaustein 1995) has resulted in the requirement of ecologists and land managers to review monitoring programs and their assessments of conservation status for species. Assignment of a misleading conservation status can result in the inappropriate allocation of resources and political capital. In Chapter 3, I review the conservation status of P. frosti by examining population trends over a 20-year period (1983 – 2002). I also review the established survey method and monitoring program used to collect data on the population, and attempt to account for temporal variation associated with counts of calling males (documented in Chapter 5) when examining relative abundance data and estimating population size. The results of surveys undertaken by Malone (1985a) in 1983 and 1984 are used as base-line data on distribution and abundance of P. frosti with which to examine population trends and compare the results of monitoring from this study.

To reverse amphibian declines, management of populations will require definition of high-quality habitat for individual species, or groups of species (Knutson et al. 1999). It is also considered that knowledge of the specific adaptations of high-mountain amphibian populations may be important in developing appropriate measures for the prevention of future declines (Vences et al. 2000). Habitats associated with oviposition, and development of embryos and larvae, are particularly important for amphibians, given that these developmental stages are considered to be the most vulnerable of the anuran life cycle (Duellman and Trueb 1994). The sensitivity of P. frosti to disturbance or habitat modification is not well known, although a study carried out by Malone (1985a) found that vegetation structure influenced embryonic and larval survivorship, with higher levels of mortality being recorded in disturbed habitats compared with undisturbed. Given the relationship recorded by Malone (1985a), and the apparent pattern of decline of P. frosti from higher-elevation habitats, this study sought to identify potential differences in breeding habitat between high and low-elevation habitats that may elucidate factors involved in the decline of the species. In Chapter 4, the attributes of extant breeding sites are compared with a sample of randomly-selected sites chosen from within the domain of known breeding habitat for the species. This comparison was stratified for high and low-elevation breeding habitats to examine potential differences between the two. I subsequently test the null hypothesis that the attributes of extant

17 breeding sites do not differ from that of the sample of randomly-selected sites. I also identify specific breeding habitat attributes and environmental gradients that contribute most to observed differences between breeding and random sites, and compare these attributes with breeding habitat described historically for P. frosti by Malone (1985a).

Calling activity and oviposition in P. frosti occur within a relatively brief period in each season (see Chapter 2). Because of the importance of vocal activity to anuran reproductive success (Wells 1988), an understanding of the ecological and distribution limits within which breeding activity takes place in P. frosti is fundamental to its conservation, as well as for investigating population declines. For some sensitive amphibians, effects on breeding patterns due to climate change, such as alterations to the timing of reproduction, could result in significant changes to population structure or population declines (Blaustein et al. 2001). Mating systems can also have a large effect on the way a population responds to natural and human-induced habitat changes, so that an understanding of them is also important for management (Anthony and Blumstein 2000). In Chapter 5, I report on patterns of breeding activity and measure the climatic environment within which breeding activity takes place by monitoring calling activity over several breeding seasons at high and low-elevation breeding sites.

Knowledge of the movement patterns of amphibians is considered fundamental to their ecology (Duellman and Trueb 1994). An understanding of how animals partition their activities and travel among resources is also necessary for species conservation (Pilliod et al. 2002). Because of the cryptic nature of P. frosti, almost all records of the species have been obtained during its breeding season (Atlas of Victorian Wildlife, Department of Sustainability and Environment). Consequently, most descriptions of its habitat are also from the breeding season (see Malone 1985a; Hollis 1995). The few records of P. frosti from outside its breeding season are from beneath building materials, and in disturbed habitats within the Mt Baw Baw Alpine Resort (M. Littlejohn, J. Coventry, P. Robertson unpublished data). There is currently no knowledge on the patterns of movement of P. frosti, and its use of habitat during the non-breeding season. In Chapter 6, using recent advancements in the miniaturisation of radio-transmitters, I examine patterns of movement and habitat use by adult P. frosti in the breeding and post-breeding season, and the influence of weather on activity.

Population demography and life history are considered fundamental elements of conservation biology (e.g., Caughley and Gunn 1996; Meffe and Carroll 1997). To effectively manage and protect populations of species, particularly those that are threatened, or have restricted distributions, an understanding of the potential for natural population fluctuations relative to the time scale and magnitude of reproductive cycles and generation time is required (Primack 1993; Caughley and Gunn 1996). Information on the demography of P. frosti is currently lacking. Mark- recapture studies have typically been successful in documenting population demography, dynamics and life history of amphibians, but are problematic when dealing with secretive or rare

18 species, or those with relatively long lifespans (Trenham et al. 2000). Due to the rarity and cryptic nature of P. frosti, a mark-recapture program using pitfall traps proved to be unsuccessful in capturing sufficient individuals to obtain demographic information (see Chapter 6). In Chapter 7, I use skeletochronological methods as an alternative to determine longevity, maturation and sex- specific growth in P. frosti.

19

Chapter 2

EXISTING KNOWLEDGE ON PHILORIA FROSTI, DESCRIPTION OF STUDY AREA AND GENERAL METHODS

2.1 Existing Knowledge on Philoria frosti

2.1.1 Distribution and Abundance

Following the discovery of P. frosti at Mt Baw Baw in 1898, the earliest post-discovery records of P. frosti in the Museum of Victoria were from Mt Baw Baw in 1955. Subsequent specimens lodged in the Museum of Victoria in the 1960s were also from the vicinity of Mt Baw Baw, collected by museum personnel on expeditions, and from early biological studies on the species (Littlejohn 1963). Including the 1970s and early 1980s, most records of P. frosti were also from locations within the vicinity of the access road to the Mt Baw Baw Alpine Resort, except for a small number from the Mt Whitelaw and Mt Erica areas. From the species' discovery in 1898 up until 1982, it is difficult to assess the species abundance in the absence of survey data. However, some information may be gleaned from some qualitative data and anecdotal observations from this period. Firstly, the fact that Charles Frost retrieved five regurgitated specimens, in fresh condition, from the tiger snake he captured might suggest that P. frosti was either relatively abundant at the time, or tiger snakes are efficient hunters of P. frosti, and secondly, the ease in which frogs were able to be located during collection trips by the Museum of Victoria and early scientific studies also indicates that the species was relatively common during the 1950s, 1960s and 1970s (M. Littlejohn unpublished data; J. Coventry pers. comm.).

In 1983 and 1984, Malone (1985a) conducted the first systematic survey of P. frosti. Using counts of calling males to estimate relative abundance, Malone (1985a) recorded 4248 males, including at sites visited twice in 1983 and 1984. The observation of silent males in choruses, and the frequent occurrence of more than one male at calling site, lead Malone to believe that total population size of P. frosti would be something in the order of 2 – 3 times the number of calling males that he recorded. His estimate of the adult male population size was subsequently 10,000 – 15,000. During these surveys, Malone (1985a) showed that the distribution stronghold of the species was primarily the western (Mt Baw Baw), central (Mt St Phillack) and north-western (Mt Whitelaw) regions of the Baw Baw Plateau, whilst fewer calling males were recorded from the eastern region (Mt St Gwinear), and the species was conspicuously absent from the south-eastern region of the plateau (Mt Erica). Areas surveyed by Malone in both years showed that numbers of calling males within

21 some breeding units remained relatively stable whilst numbers within others did not. Those that varied between 1983 and 1984 were mainly on the eastern side of the plateau, where a contraction in range of calling males was recorded in 1984. Malone speculated that this observation may have been due to the combination of lower spring and summer rainfall received on the Baw Baw Plateau in 1984, and the eastern side being drier in nature than other regions on the plateau. Overall, breeding aggregations of P. frosti were noted at this time as being confined predominantly to elevations above 1300 m, encompassing an area of approximately 80 km2 on the Baw Baw Plateau (Malone 1985a, b), although a small number of breeding aggregations were recorded by Malone (1985a) along the access road to the Mt Baw Baw Alpine Resort as low as 1160 m, and there was an additional record of P. frosti from near Mushroom Rocks (1100 m) near Mt Erica (Atlas of Victorian Wildlife 2003, Department of Sustainability and Environment, 123 Brown St, Heidelberg, 3084).

Between 1985 and 1992, no records of P. frosti were registered on the Atlas of Victorian Wildlife. Anecdotal observations made by biologists in the vicinity of the Mt Baw Baw Alpine Resort between 1987 and 1992 suggest that only low numbers of calling males were present (W. Osborne, G. Marantelli, M. Hutchinson, J. Morey, P. Johnson pers. comm.), signifying a possible decline in population numbers of P. frosti during this period. In 1992, concern was raised over the conservation status P. frosti due to an increasing number of reports documenting the decline and disappearance of other amphibians, particularly those restricted to mountain-top environments in relatively pristine habitats, and a scarcity of records of P. frosti since the survey of Malone (1985a).

In 1993, the then Victorian Department of Conservation and Natural Resources initiated a systematic survey of the population of P. frosti, with the objective of comparing results obtained by Malone (1985a) ten years earlier. Using the same census methodology as Malone (1985a), the survey results indicated that the population of P. frosti had declined by several orders of magnitude, measuring only 2% of former counts from 1983 and 1984 (Hollis 1995). In addition to this decline, Hollis (1995) also recorded the population as having contracted further from the eastern region of the Baw Baw Plateau compared with the distribution recorded by Malone (1985a) in 1984. A subsequent survey of the population of P. frosti in 1994 yielded almost identical results to that obtained in 1993 (Hollis 1995).

The results of further surveys conducted in 1995, 1996 and 1997 verified the decline and contraction in range of the population of P. frosti recorded in previous years (Osborne et al. 1999). However, in 1996, the survey was extended into lower elevation habitats (< 1300 m) not previously surveyed by Malone (1985a). The presence of breeding aggregations that Malone recorded along the access road to the Mt Baw Baw Alpine Resort as low as 1160 m, suggested that other populations may exist in other localities on the south-western escarpment on the Baw Baw Plateau at similar elevation. Populations of P. frosti were subsequently discovered in montane

22 forest habitats as low as 960 m on the south-western escarpment, and later, on the north-eastern escarpment as low as 1200 m, as reported in this thesis.

2.1.2 Macro and Micro-habitat

Due to the cryptic nature of P. frosti, almost all records from which details of habitat have been recorded are from the breeding season, when calling males can be located. Malone (1985a) provided the most detailed account of breeding habitat use by the species when it was considered to be relatively common. By far the greater number of frogs counted by Malone (1985a) were recorded in wet alpine heath-bog ecotones (75%), the remainder being recorded in wet alpine heath (22.5%), bogs (1.6%) and grassland [modified areas within the Mt Baw Baw Alpine Resort] (0.9%). Malone also noted that a small percentage (2.4%) of frogs in vegetation communities dominated by Leptospermum grandifolium and Nothofagus cunninghamii.

In the survey conducted by Hollis (1995) in 1993, a shift in the proportion of calling males using different habitat types, compared to the results of Malone (1985a), was recorded. In wet sub-alpine heathland (equivalent to Malone’s wet alpine-heath-bog ecotone, wet alpine heath and bog area combined), 39.4% of calling males were recorded. Of these, only a small number were in the heath-bog ecotones reported by Malone (1985a), with most being recorded further up-slope in wet alpine heath. A further 39.4% were recorded by Hollis (1995) in ecotonal habitats at the periphery of wet sub-alpine heathland, adjacent to montane riparian thicket, dry sub-alpine shrubland and sub-alpine woodland. Malone (1985a) did not record any aggregations of calling males in habitats such as these. It is possible, however, that some frogs that were recorded by Malone as occurring in wet alpine heath may have actually occurred in these habitats (B. Malone pers. comm.). Another apparent difference in habitat use recorded by Hollis (1995) was the greater proportion of calling males recorded in montane riparian thicket (21.2% ), compared to Malone’s 2.4%. It is also possible that Malone may have overlooked a number of calling males in this habitat type, given the difficulty in accessing thickets, and having to contend with counting large numbers of calling males in wet alpine heath-bog ecotones where it appeared the majority were (B. Malone pers. comm.). No calling males were recorded by Hollis (1995) in the modified areas within the Baw Baw Alpine Resort. The results of surveys conducted by Hollis (1995) suggest that breeding localities utilised by P. frosti at the time were more restricted to topographically protected gully habitats compared with the broader distribution of the species recorded by Malone (1985a).

Oviposition sites utilised by P. frosti have been recorded as natural cavities in or under vegetation, logs, peat and soil, and rock (or combinations of these), that act as catchments for water travelling down slope (Littlejohn 1963; Malone 1985a; Hollis 1995). These breeding micro-habitats occur along seepages or rivulets that drain water from the slopes of sub-alpine wet heathland and in gullies that drain the plateau. Oviposition has also been reported from beneath building materials in the Mt Baw Baw Alpine resort (Malone 1985a; Littlejohn unpublished data; J. Coventry and P.

23 Robertson pers. comm.). Malone (1985a) recorded 79% of oviposition sites in cavities of vegetation consisting of spp., Astelia alpina, Empodisma minus, Epacris paludosa, Richea continentis and Carex spp., with the remaining 21% of egg masses being recorded from beneath logs, rocks and building materials. Hollis (1995) reported a similar proportion of breeding sites in micro-habitats similar to that of Malone, but the assemblage of plants was slightly different. Most calling sites (76%) were located amongst the roots of vegetation, dominated by the species R. continentis, E. paludosa, Baeckea latifolia, Orites lancifolia, Callistemon pityoides, L. grandifolium, N. cunninghamii, Sphagnum cristatum, Wittsteinia vacciniacea, Polytrichum alpinum/commune, A. alpina, Carex gaudichaudiana and C. appressa. The remaining 24% of calling sites were located in peat or soil cavities beneath large granite boulders or logs. The different assemblage of plants recorded by Hollis (1995) can be largely attributed to the proportionally greater number of calling males recorded in montane riparian thicket vegetation, and in ecotonal habitats at the periphery of sub-alpine wet heathland, as noted above.

Other micro-habitats described for P. frosti include those for larvae. Most larvae usually remain at oviposition sites through to metamorphosis (Barker et al. 1995; G. Hollis pers. obs.), although upon hatching, they have also been observed to move small distances in shallow water from oviposition sites, while remaining covered under vegetation and/or woody debris (G. Hollis pers. obs.), or to be washed into nearby pools (B. Malone pers. comm.).

The use of non-breeding habitats by P. frosti is mostly unknown. There are a small number of anecdotal records of the species outside the breeding season. These are from beneath logs, rocks and building materials around the Mt Baw Baw Alpine Resort (M. Littlejohn, P. Robertson and J. Coventry pers. comm.). Over-wintering habitats used by the species are unknown.

2.1.3 Life History and Ecology

Calling activity by P. frosti has been recorded as taking place between late October and late December (Malone 1985a; Barker et al. 1995; Hollis 1995), although some individuals have been heard in January (J. Coventry pers. comm.). Oviposition appears to be confined to a shorter interval of 2 - 3 weeks during this period, when a peak in calling activity occurs (Malone 1985a, b; Hollis 1995). Recorded peaks in calling activity include, 11 - 29 November 1983/84 (Malone 1985a, b) and 17 - 29 November 1993 (Hollis 1995). High levels of calling activity have been recorded diurnally, except when air temperatures fall below 5oC (Malone 1985a). The advertisement call structure of P. frosti is irregular, complex and variable (Littlejohn 1963, 1987), and has been described as a short ‘clunk’, repeated in sequences of up to 30 calls (Littlejohn 1963). Further analyses of the call structure of P. frosti (call duration and frequency) are provided by Littlejohn (1963).

24 The egg mass of P. frosti is deposited in a transparent foam nest at the calling site, or nearby, during inguinal amplexus. The foam nest measures 8 cm in diameter and 3 - 4 cm in height, and is produced by the female beating air bubbles into the mucous and eggs with flanged fingers during egg laying (Littlejohn 1963). It is not known how long male and female P. frosti remain in amplexus. The egg mass may be deposited at varying depths in vegetation, or below the ground surface, depending on the structural attributes of the site. Depths of over a metre have been observed, whilst others are deposited in vegetation very close to the surface (G. Hollis pers. obs.). Clutch sizes reported in the literature range from 50 - 185 (see Littlejohn 1963; Malone 1985a, b; Tyler 1992). The ova are unpigmented and are considered large, measuring on average 4 mm in diameter (Malone 1985b), whilst the capsule measures 5.8 mm diameter (Littlejohn 1963). Oviposition of more than one egg mass may occur at a single site (Malone 1985b; G. Hollis pers. obs.), and it is also possible that females deposit a portion of their eggs at more than one site (Malone 1985a.). Under natural conditions the embryonic period varies from 5 - 8 weeks, with individuals hatching at Gosner stages 22 - 23 (Malone 1985a, b).

Prior to the study by Malone (1985a, b), there was some uncertainty as to whether P. frosti possessed free-swimming larvae (e.g., Martin 1967; Watson and Martin 1973; Littlejohn 1963). Upon hatching, larvae are initially unpigmented, with no external gills (Anstis 2002). They are non-feeding, hatching with a large residual yolk mass to sustain them through to metamorphosis. Under experimental conditions, Malone (1985a) found that the provision of food, and stocking of larvae at different densities, had no significant effect on development. Although larvae are non- feeding, and development usually takes place at the oviposition site with very little water (G. Hollis pers. obs.), the ability to swim has been retained (Malone 1985a, b). Under natural conditions, the larval period varies from 5 - 10 weeks, at which time individuals metamorphose at an average length of 6.72 mm snout-vent length (Malone 1985a). Anstis (2002) provides a more detailed description of larvae at various stages of development.

Very little is known about the population dynamics and demography of P. frosti. The little work that has been conducted documents mortality during embryonic and larval stages of development (Malone 1985b). Mortality during these stages of development is relatively high, measuring 74% and 70.3% respectively, with recruitment to the terrestrial stage measuring only 8.1% (Malone 1985b). Overall survivorship is extremely variable among breeding localities, with periodic drying of oviposition sites suggested as the primary cause of the high mortality rates observed (Malone 1985a, b). Such mortality rates, however, are not significantly different to rates recorded for other anurans (Malone 1985a). Vegetation structure at oviposition sites has been shown to significantly affect rates of mortality of embryos and larvae, with higher levels being recorded in disturbed habitats (80.1% and 68.3% respectively) compared to undisturbed habitats (62.5% and 44.5%) (Malone 1985a). An attempt to approximate the proportion of female P. frosti in the population, by examining secondary sex ratios, proved inconclusive due to an inability to differentiate gonads in metamorphosed individuals (Malone 1985a).

25 2.2 Study Area

This study was conducted on the Baw Baw Plateau and adjacent escarpment, located approximately 120 km east of Melbourne in the Central Highlands of Victoria on the north-eastern edge of the Latrobe Valley (Fig. 2.1). The plateau is one of a number of uplifted erosional plains in the eastern section of the highlands considered to be the remnants of an extensive Mesozoic land surface (Hills 1959). It is bounded by the catchments of the in the east (now the Thomson Reservoir), the East Branch and West Branch of the in the south, the in the west and the upper Thomson River in the north. The area above approximately 1200 m a.s.l. is contained mostly within the Baw Baw National Park, whilst the area below this elevation is mostly State Forest, except for the Mt Baw Baw Alpine Resort which encompasses an area of approximately 3 km2 on the plateau’s western side. This study was confined to elevations above approximately 800 m.

2.2.1 Climate

Unfortunately there are no long-term meteorological data from the Baw Baw Plateau from which to describe the general climate of the study area. Instead, descriptions of climate for the area are inferred from nearby meteorological stations with similar altitude and latitude. The plateau exerts a pronounced effect on the local climate by intercepting rain-bearing westerly and south-westerly winds (DCE 1991). Sibley (1975) estimates annual precipitation for the plateau to be in the vicinity of 1900 mm, making it one of the wetter areas in Victoria. Aldrick et al. (1992) report that the plateau receives a mean annual rainfall of 1500 - 2500 mm. Being a sub-alpine environment, climate on the plateau at higher elevation can be severe, with heavy frosts occurring in the low valleys, and snow on the ground above approximately 1200 m from June to September, with approximately 2 m of cover being recorded in good years (DCE 1991). Estimates of temperature include a mean annual temperature range of 4 - 8 o C, July mean minimum temperatures of -2 - 0 o C and February mean maximum temperatures of 11 - 13 o C (Aldrick et al. 1992). Sibley (1975) also notes that it is doubtful if maximum daily temperature in summer ever exceeds 32 o C, except perhaps on days of exceptionally hot north winds.

2.2.2 Geomorphology and Geology

The Mt Erica-Mt Baw Baw massif (Baw Baw Plateau) comprises of a mass of resistant igneous rock composed of Upper Devonian granodiorite which was intruded into surrounding Siluro-

26 Devonian sedimentary rocks (Sibley 1975; ESAV 1982). This batholith is connected in the north- west to a similar intrusive mass, known as the Gembrook batholith (ESAV 1982). It also connects with a zone of contact metamorphism where the sedimentary rocks in contact with the igneous intrusion have been heated and crystallised to form a zone of hardened rock fringing the granodiorite (Douglas and Ferguson 1976). ESAV (1982) point out that, following its intrusion, the Baw Baw batholith has evolved through a number of stages, including the removal of an unknown depth of sedimentary rocks through weathering and erosional processes, beneath which the granodiorite mass was buried. Because the granodiorite is much more resistant to weathering processes than the sedimentary strata, the existing plateau remnant was formed. Further uplifting and erosional processes resulted in the continued dissection of the softer rocks surrounding the Baw Baw batholith. This has resulted in a number of distinct geomorphological features and patterns of drainage, which have resulted in the present-day form of the Baw Baw Plateau (see below).

The plateau above 1200 m is roughly elliptical in shape, with its major axis running from north- west to south-east for a distance of approximately 20 km, and its minor axis running from north- east to south-west for approximately 7 km (ESAV 1982) (Fig. 2.1). At higher elevation, the plateau comprises of generally low, rounded hills (60 – 90 m high) separated by flat-bottomed valleys with relatively little relief between the hill tops and valley floors (Sibley 1975). Rosengren et al. (1981) identify a variety of geomorphological features associated with the weathering of the granodiorite: broadly concave valleys, peaty flats, tors, stepped valley heads and a distinctive rectangular drainage pattern derived from the cooling joints in the granodiorite. ESAV (1982) report that the average plateau height is approximately 1500 m, which places it well above the Great Dividing Range located about 25 km to the north, with the highest point being Mt St Phillack (1566 m), then Mt Baw Baw (1564 m), whilst Mt St Gwinear, Mt Mueller, Mt Kernot, Mt Tyers and Mt Whitelaw range in height between 1417 and 1509 m. DCE (1991), however, report that Mt Baw Baw is the highest point at 1565.6 m (Australian Height Datum), and Mt St Phillack the second highest. The escarpment surrounding the plateau comprises of steep erosional slopes up to 30 degrees, which fall rapidly to a general elevation of 500 m (DCE 1991). In contrast with the rectangular pattern of drainage on the granodiorite, drainage on the escarpment of the plateau is distinctively dendritic (ESAV 1982).

2.2.3 Soils

Sibley (1975) describes the physical and chemical properties of soils derived from the granodiorite on the Baw Baw Plateau and adjacent escarpment. They can be divided broadly into mineral soils on the rocky hills and escarpment slopes, and organic soils (peats) in the valleys, hill-side drainage lines and lower slopes of the hills. The mineral soils can be classified as either transitional alpine soils or acid brown earths, both generally friable and porous with an open, earthy fabric that

27 affords them an increased capacity to hold water. Differences between the two soils relate to the relative exposure, steepness, supply of organic matter and depth of the parent material above the decomposing granodiorite at a site, with transitional alpine humus soils generally occurring on gently sloping sheltered sites and acid brown earths occurring on steeper and more exposed sites where there is less organic matter and the parent material is shallow. At lower elevation on the escarpment slopes, these soils are characterised by a thick layer of decomposing plant litter at the surface, over a deep organo-mineral horizons that comprise large amounts of organic matter and excellent structure, beneath which lies a mineral horizon comprising of sand and smaller amounts of clay, with only minor amounts of organic matter.

The organic soils can be classified as bog peats or humified peats. Bog peats are confined to permanently wet bogs of sphagnum moss located on the valley floors and hill-side and escarpment drainage lines, and are composed of layers of raw and decomposing bog vegetation. Humified peats occur on the sloping edges of the bogs, in eroded desiccated bogs and on the lower slopes of the hills. These particular sites are characteristically located in areas where the water table is low enough to allow the peats to dry out and become aerated. They comprise generally of peats where the vegetation remains have lost their identity, and the change to a homogeneous mass of colloidal matter has proceeded.

2.2.4 Vegetation

Vegetation types documented for the study area have been described and classified at different scales. More recently, vegetation in Victoria has been classified and described as Ecological Vegetation Classes (EVC’s), which group together one or more floristic communities that exist under a common regime of ecological processes within a particular environment at a regional, state or continental scale (Woodgate et al. 1993). EVC’s that occur in the study area include sub-alpine woodland, sub-alpine wet heathland, sub-alpine damp heathland, sub-alpine shrubland, sub-alpine riparian shrubland, montane riparian thicket, montane wet forest, montane damp forest, wet forest and cool temperate rainforest (LCC 1991; Davies et al. 2002). An additional two vegetation types (montane fen and sub-alpine rocky-outcrop shrubland) have also been described from the study area (Davies 2000; Davies et al. 2002), but are yet to be recognised formally as EVC’s. Prior to the development of the EVC concept, descriptions of vegetation from the study area were in the form of floristic communities and sub-communities (see Cunningham et al. 1971; Gullan et al. 1981; 1984; 1985; Davies et al. 1994; Walsh et al. 1984; Hollis et al. 1995). More recently, Peel (1999) described in more detail the floristics and structure of Victorian rainforests and cool temperate mixed forests. The results of his study recognised cool temperate mixed forest as an additional vegetation type among other rainforest EVC’s. Peel also identified a number of floristic community variants of the cool temperate rainforest EVC that occur in the study area. These were Central Highlands Cool Temperate Rainforest, Central Highlands Montane Riparian Cool

28 Temperate Rainforest and Central Highlands Montane Scrub Cool Temperate Rainforest. Table 2.1 summarises the relationship between vegetation types that occur in the study area, as well as those documented for P. frosti by Malone (1985a) (see 2.1.2).

2.2.5 Fire History

Records of past fires in the vicinity of the study area are incomplete, although it is apparent that there have been several major wildfires over the past 140 years, including 1851, 1898, 1926, 1932 and 1939 (DCE 1991). The entire Baw Baw Plateau was known to have been burnt in 1851 and 1932 (Williams and Baker 1979; DCE 1991), whilst the southern slopes of the plateau was known to have been burnt in 1926 (DCE 1991). No major wildfires are known to have occurred in the vicinity of the Baw Baw Plateau since 1939 (DCE 1991). A number of fires are known to have started from lightening strikes, but have either been extinguished quickly by the land manager, or through natural means.

Early timber millers lit fires at the end of the timber harvesting season to allow easier forest access the following year (Davies et al. 1994). Fire is currently used as a management tool to promote forest regeneration following clearfell timber harvesting in State Forest. Regeneration burns are conducted following harvesting each year, mostly during Autumn. The Gippsland Fire Protection Plan (NRE 1999) generally excludes fire from areas containing stands of ash (Eucalyptus regnans, E. delegatensis and E. nitens), rainforest and alpine vegetation for fuel reduction purposes.

2.2.6 Historical landuse

2.2.6.1 Aboriginal History

Information on aboriginal history within the study area is scant and anecdotal. The name 'Baw Baw' appears to be of aboriginal origin, meaning 'echo', whilst an early map of Gippsland has the Baw Baw Plateau named as 'Bo Bo' (Waters 1966). No archaeological surveys have been conducted on the Baw Baw Plateau (DCE 1991), although some aboriginal implements have been located on the spur between Tanjil Bren and Mt Whitelaw (Waters 1966). It seems likely that aboriginal people lived in the Thomson catchment and used environments ranging from the coast to the mountains (DCE 1991). If the high elevation areas were used at all by aboriginals, their activity would have been seasonal given the lack of available food during the cold winters. However, there does not seem to be any evidence of any feast by aboriginals on moths as occurred on the Bogong High Plains area (see Flood 1980). Interestingly, Waters (1966) reports that the Traralgon Aborigines refused to go near the Baw Baw mountains due to legends that involved

29 some of their tribe being attacked by a ferocious species of yellow snake, or the possibility that one could be sucked into a boiling pool if they ventured too close.

2.2.6.2 European History

The sketch map by of the terrain between Port Albert and Omeo contains the earliest mention of the name Baw Baw (Waters 1966). made the first ascent of the Baw Baw Plateau during his botanical expedition in 1860, whilst the Geodetic and Coast Survey of Victoria was conducted in 1870, when the cairn built on Mt Baw Baw was used as a trigonometrical station (Waters 1966). Cattle grazing is the family up until 1913, whilst a subsequent grazing lease was held by Fred Jan until the first type of land use documented for the Baw Baw Plateau, with the first grazing lease held by the Rawson 1939 wildfire swept over the plateau (Stephenson 1980). Between 1958 and 1975, Hec Stagg grazed cattle on the plateau under licence, until it was withdrawn to protect part of the Thomson Catchment. An additional cattle muster was conducted in 1985 to remove those left behind in 1975, although this was only partially successful (DCE 1991).

Gold mining commenced in 1860, but most of this activity went on outside the study area at locations at lower elevation due to the lack of gold on the Baw Baw Plateau. As part of the network of access tracks to diggings, the first track was cut onto Mt Baw Baw in 1888 (DCE 1991). With the closure of most gold mines by 1914, sawmilling had grown to become the most significant land use in the vicinity of the study area. A significant network of timber tramways and bush mills was subsequently established on the lower, south-western slopes of the plateau until most were lost in the 1939 wildfire. Following this, timber mills were relocated into townships and an extensive salvage operation, primarily of ash timber, commenced when there was increased demand for wood during the second World War.

The Baw Baw Plateau was promoted as a destination for tourists and recreationalists as early as 1906 (Walters 1966), with activities being predominantly bush walking, snow sports and trips on horseback (DCE 1991). Interest in these activities diminished following the 1939 wildfire and the decline of Whalhalla area in 1942. In 1944, interest in the plateau shifted from the Mt Erica – Talbot Peak area, where access had been gained via the tramway network, to the Mt Baw Baw area following the establishment of a track to Mt Baw Baw from Tanjil Bren. With improved road access to Mt Baw Baw, a village was developed nearby in the 1960s. More recently, the Mt Baw Baw Alpine Resort was proclaimed under the provisions of the Alpine Resort Act 1986. In the 1970s, the Alpine walking Track was established on the Baw Baw Plateau, traversing its length from the south-east to the north-west regions (DCE 1991). Encompassing an area of 5200 ha, an Alpine Reserve was created in 1959 on the plateau above the 1220 m contour by the then Forests Commission, and was further supplemented by an additional 2800 ha of adjacent Crown Land three years later (MMBW 1975). Recognising the area's conservation and recreation value, 13,300

30 ha covering the Baw Baw Plateau and some adjacent land in the Thomson and Aberfeldy River catchments were reserved as National Park in 1979 (DCE 1991).

2.2.7 Current Landuse

The study area is currently managed for a variety of uses: timber production in State Forest, where it is estimated that there is approximately 3168 ha of merchantable timber (Department of Sustainability and Environment internal data), biodiversity conservation, and recreation, including bush walking, camping, picnics, snow activities, sight seeing, four-wheel driving, trail-bike riding, mountain-bike riding and hunting (NRE 1998). The plateau and adjacent escarpment also form a significant component of the water catchment for Moondarra Reservoir and Thomson Reservoir ( supply), and Blue Rock Dam. Sibley (1975) notes that the valley bog on the Baw Baw Plateau regulates the stream flow by absorbing a considerable portion of snow melt during spring, with subsequent flows being gradually released over summer when normal discharge is at its lowest.

2.2.8 Sites of Biological and Ecological Significance

The study area comprises a richness of significant biological and ecological sites. The Baw Baw Plateau and upper Thomson River area is described as a site of global significance with respect to its zoological values (Mansergh and Norris 1982; Lumsden 1991). The Tyers River West Branch, encompassing 500 m either side of the river, is also listed as a site of zoological significance by Lumsden (1991). Sites of botanical significance occur in the Tyers River West Branch, South Cascade Creek, Little Boys Creek, Bell Clear Creek, upper Thomson River and Myrrhee Creek (Moorrees and Molnar 1992). The Baw Baw Plateau is also listed as a site of geological and geomorphological significance by Rosengren et al. (1981).

2.3 Nomenclature

Amphibian taxonomy in this study follows that of Cogger (2000). Nomenclature of Victoria plants follows NRE (2001).

31 2.4 Statistical Analyses

Except where stipulated, all statistical analyses were conducted using SYSTAT versions 7.01 and 8.01 (SPSS, Inc, Evanston Illinois). All presented mean values are followed by ± 1 standard error (SE).

32 Table 2.1. Relationship between vegetation types described for the study area and habitats utilised by P. frosti, as reported by Malone (1985a).

Vegetation Types Ecological Vegetation Rainforest and Cool Floristic Communities Utilised by P. frosti Classes (LCC 1991, Temperate Mixed of Cool Temperate (Malone (1985a) Davies et al. 2002) Forest (Peel 1999). Rainforest (Peel 1999) Wet Alpine Heath Sub-alpine Wet Heathland - - Bog Sub-alpine Wet Heathland - - Wet Alpine Heath/Bog Sub-alpine Wet Heathland - - Ecotone vegetation dominated Montane Riparian Thicket - Central Highlands by Leptospermum Montane Scrub Cool grandifolium and Temperate Rainforest Nothofagus (synonymous in part with cunninghamii Montane Riparian Thicket)

- Cool Temperate Rainforest Central Highlands Cool Central Highlands Temperate Rainforest Montane Scrub Cool Temperate Rainforest

Central Highlands Montane Riparian Cool Temperate Rainforest - Montane Wet Forest - - - Unclassified Moist Forest Central Highlands Cool - Temperate Mixed Forest -Wet Forest- - - Sub-alpine Woodland - - - Sub-alpine Damp Heathland - - - Sub-alpine Shrubland - - - Montane Damp Forest - - Sub-alpine Riparian Shrubland -Fen- - - Sub-alpine Rocky-outcrop -- Shrubland grassland (modified --- areas within Mt Baw Baw Alpine Resort)

33 N Matlock #

Thomson Reservoir Mt Baw Baw Alpine Reserve

# Warburton # Private Land Baw Baw National Park Other Public Land # # # Tanjil Bren # Parks and Reserves Melbourne # Noojee State Forest Powelltown # Other Tyers Junction # Study Area # Walhalla # # Erica Neerim South Dandenong #

Bunyip #Cranbourne # Drouin # # Trafalgar # #

Latrobe Valley Latrobe Valley

2002040KilometersKilometres Fig. 2.1. Location of study area and associated land tenure. Kilometres

Fig. 2.1. Location of study area and associated land tenure.

34 Chapter 3

PATTERNS OF DISTRIBUTION AND ABUNDANCE, AND AN ASSESSMENT OF COUNTS OF CALLING MALES AS A METHOD FOR MONITORING AND ESTIMATING POPULATION SIZE

3.1 Introduction

Recognition of a current global biodiversity crisis has lead to an urgency in assessing the conservation status of many species. This urgency is particularly relevant to amphibians, where there have been numerous reported cases of localised and regional declines and disappearances over the past two decades (reviewed in Chapter 1). One aspect that has confounded attempts to identify factors responsible for observed declines has been the inability to distinguish apparent declines from natural population fluctuations (Pechmann et al. 1991; Pechmann and Wilbur 1994; Beebee and Rowe 2001; Storfer 2003). A lack of long-term population and climatic data (Pechmann et al. 1991; Blaustein 1994; Blaustein et al. 1994; Pechmann and Wilbur 1994; Gillespie and Hollis 1996) and low statistical power (McCoy 1994; Pechmann and Wilbur 1994; Reed and Blaustein 1995) have contributed substantially to this predicament.

Documentation of changes in the distribution and abundance of species is considered a fundamental requirement for understanding the consequences of past actions and in formulating future management plans (Skelly et al. 2003). However, amphibian populations are known to fluctuate greatly over time (Tyler 1991; Duellman and Trueb 1994), increasing the difficulty in establishing their status and distribution. For example, more thorough examinations of distribution and abundance have revealed cases where amphibian populations were reported initially to have declined, after which they were recorded as either recovering from the decline, fluctuating in numbers, or were found to occur over a larger distribution (e.g., Pechmann et al. 1991; Pechmann and Wilbur 1994; Gillespie and Hollis 1996; Parris 2001). Lack of consideration of how amphibian populations vary both spatially and temporally, coupled with limited long-term population data, could therefore lead to the mistaken conclusion that a decline has occurred when normal population fluctuations are being exhibited, or that no reduction in population has occurred when the species has actually declined. Determining the conservation status of amphibians that are long-lived may also be inhibited, where slow growth rates, delayed maturity, and low fecundities can lead to slow recovery from perturbations (Wheeler et al. 2003). Assignment of a misleading conservation status or poor estimates of extinction risk, followed by inappropriate allocation of resources and political capital could occur as a result of such mistaken conclusions (Blaustein

35 1994; Blaustein et al. 1994, McCarthy 1997; Marsh 2001). In addition to this, because amphibians, along with reptiles, are viewed socially and politically as ‘deviants’ compared to other non-human species such as birds and mammals (Czech et al. 1998), mis-information about the status of populations could further disadvantage access to resources. Land managers and conservation biologists are now challenged with the goal of developing monitoring programs that provide statistically defensible estimators of abundance, within a framework of diminishing resources and time.

To establish the conservation status of a species, its distribution and population size have to be estimated accurately and efficiently over time. A variety of survey methods has been identified as being useful for amphibians, each varying with respect to the type of data collected, time, cost and personnel required (see Heyer et al. 1994). The amphibian survey method based on counts of calling males is considered to be a simple, rapid, cost effective method that provides a relative abundance estimate of calling males, and is particularly useful for species that are cryptic and where large amounts of habitat are required to be surveyed (Zimmerman 1994). The precision of this method in providing a relative abundance measure, however, is reliant upon the assumption that calling activity, and site turn-over of individuals, are spatially and temporally uniform both within and between breeding seasons. In circumstances where counts of calling males have been calibrated to consider variation in male calling activity and turn-over, the method is considered to be useful for monitoring and estimating the size of adult populations (Zimmerman 1994; Driscoll 1998; Bridges and Dorcas 2000).

In 1982, the conservation status of P. frosti was listed as Vulnerable among threatened species of wildlife (Ahern 1982). Anecdotal records of P. frosti collected during the 1950s, 1960s and 1970s by the University of Melbourne and Museum of Victoria personnel suggested that the species had then been relatively common, due to the relative ease with which individuals were located (M. Littlejohn unpublished data and pers. comm., J. Coventry pers. comm.). To address a lack of information about population size, distribution and ecology, a systematic survey of the species was undertaken in 1983 and 1984 (Malone 1985a). This survey utilised a count of calling males during the breeding season to derive a relative-abundance estimate of adult males. The method recorded high levels of calling activity throughout the day, except when air temperatures fell below 5oC. However, Malone acknowledged that the method underestimated the absolute male population size due to the observed presence of silent males at chorus sites, and frequent occurrence of more than one at a calling site. His estimate of 10,000 - 15,000 adult males attempted to take these observations into account, where he multiplied his count of 4248 calling males by two and three. However, Malone’s estimate did not account for seasonal variation in participation by calling males.

Using a count of calling males, a monitoring program for P. frosti was established in 1993 to reassess distribution and conservation status of the species (this study). The commencement of this

36 monitoring program was stimulated by the then increasing number of reported amphibian declines and disappearances, particularly among high-altitude and mountain-top species, and a scarcity of records of P. frosti since 1984 (Hollis 1995). At this time, the conservation status of the species was listed as vulnerable (CNR 1993; Schedule 1 of Endangered Species Protection Act 1992). Initial survey results from 1993 and 1994 indicated that the population of P. frosti, as surveyed by Malone (1985a) in 1983 and 1984, had declined by approximately 98% (Hollis 1995). A contraction in range of breeding sites from the eastern side of the Baw Baw Plateau was also recorded during these surveys.

In this Chapter, the conservation status and distribution of P. frosti is reassessed in view of the lack of data on the species since 1984, and preliminary survey results from 1993 and 1994 that indicate a dramatic population decline from higher elevation areas. A count of calling males was considered an appropriate census technique for examining the population, given the cryptic nature of the species, remote location, and logistical difficulties posed by its high-elevation habitat. Such a count would also allow a direct comparison of census data previously collected by Malone (1985a). Establishment of an appropriate null hypothesis or base-line data is pertinent to any monitoring program (Storfer 2003). Counts of calling males from 1983 and 1984 were used as the base-line data set with which to examine trends in distribution and population data from this study (1993 - 2002). Comparisons are made between data collected in the field (raw data) and the same census data corrected for seasonal variation in participation by calling males. The monitoring program established for P. frosti is assessed with respect to its power to detect population trends over time. The total size of the extant population of adult males is also estimated.

3.2 Material and Methods

3.2.1 Abundance and Distribution

The survey method involving a count of calling males was adapted to suit the distribution pattern and habitat of P. frosti. Males typically call from natural cavities in or under vegetation, logs, peat and soil, and rock (or combinations of these), that act as catchments for water travelling down slope (Littlejohn 1963; Malone 1985a). These sites occur along seepages or rivulets that drain water from the slopes of sub-alpine wet heathland, and in gullies that drain the plateau. Malone (1985a) initially developed the technique to suit surveys undertaken predominantly in sub-alpine breeding habitats (> 1300 m a.s.l.) in 1983 and 1984. These breeding habitats comprised relatively broad, treeless areas on the Baw Baw Plateau, and are commonly described as frost hollows. It is within these discrete frost-hollow areas, identified and named by Balkau (1982) (updated later in 1987), that Malone (1985a) quantified the distribution and sizes of males choruses. Malone's

37 census method in frost hollows involved walking along a route in sub-alpine wet heathland, approximately 50 m from, and parallel to the boundary with sub-alpine woodland. The searcher periodically stopped at arbitrary intervals for several minutes to listen for, and count calling males. Surveys were conducted mostly during the day at temperatures above 5oC; below that they were abandoned. The routes surveyed within each frost hollow area are described as transects in the remainder of this chapter. Additional transects were also established in some of the larger frost hollows, separated by a distance of 50 m. This distance was an arbitrary value selected by Malone (1985a), and was not based on any formal assessment of the distance from which frog calls could be heard. It therefore assumes that calls by male P. frosti can be heard from a distance of up to 50 m.

To compare abundance and distribution data collected by Malone (1985a) in 1983 and 1984, this study adopted the same census methodology and survey areas used by Malone (Fig. 3.1). From 1993 - 1995, surveys were undertaken primarily within higher elevation, sub-alpine areas previously surveyed by Malone. However, following the discovery of further populations of P. frosti at montane elevations in 1996, additional transects were established and surveyed on the south-western and north-eastern escarpments of the Baw Baw Plateau from 1996 – 2002 (Fig. 3.2). Montane breeding habitat was more linear in distribution compared with the broader, sub-alpine frost-hollow areas. Survey transects were therefore confined predominantly to known drainage lines (creeks and seepages) following hydrological mapping of Roberts (1996). The census methodology used at montane elevations was also adapted to suit the distribution of breeding habitat present, involving two searchers walking down either bank of each drainage line or gully system, up to 50 m from the water course. For smaller creeks and side-gullies, only a single observer was required. Due to the low numbers of calling males present over the duration of this study, frogs were counted individually with little difficulty. Fewer survey transects were established in the southeastern region of the study area, containing the headwaters of the Tyers River West Branch and Tanjil River West Branch, due to limitations imposed by remote access (see Fig. 3.1).

The area encompassing records of the known distribution of P. frosti from this study, and that of Malone (1985a), was estimated using the Minimum Convex Polygon method for home range analyses (Ranges V software, Kenward and Hodder 1996). The area comprising potential habitat of the species was also estimated by summing the area of six geographic regions considered to be potentially occupied by the species (Fig. 3.3, see 3.2.5).

3.2.2 Survey Effort and Timing

A total of 182 survey transects, comprising 363.9 km of drainage systems, was examined to monitor the relative abundance of calling males and investigate potential breeding habitat from which the species might be recorded (Figs 3.1 and 3.2, Appendix 3.1). Incidental locations of

38 calling males were also recorded. One survey transect, ‘The Morass’, did not have its entire area surveyed in some years compared with the area originally surveyed by Malone (1985a). This resulted in a slightly smaller area being surveyed in all years (1993 – 2002) and the entire area being surveyed in just five years (1997 – 2000, 2002; Fig. 3.1). Census statistics are presented for both modified (‘The Morass excluding 2 sections’) and unmodified areas (‘The Morass’) (Appendix 3.1), however, only the data from the modified transect were used in the analysis of population trends. It was considered that because the 1983 census statistic from the ‘The Morass’ (667) was several orders of magnitude larger than the 1993 – 2002 statistics from the modified transect, there would be little, if any influence on the analysis, whilst allowing inclusion of ‘The Morass’ transect in the analysis.

From 1993 – 1995, two survey personnel were responsible for conducting surveys at established transects occurring mostly at sub-alpine elevations. From 1996 – 2002, the establishment of additional montane survey transects on the north-eastern and south-western escarpments of the Baw Baw Plateau led to a requirement for extra personnel. Teams comprising of 4 - 10 people were involved in conducting annual surveys over this time. The length of time spent conducting surveys in any one season was determined by the period over which males were calling. Survey duration was determined in an arbitrary fashion by regularly inspecting calling activity a number of known breeding sites. Estimates of the duration of each breeding season were determined more accurately by later examining calling activity data from recording units placed at breeding sites in each of years 1994 – 1999 (see Chapter 5).

3.2.3 Seasonal Correction of Census Data

Analysis of the patterns of calling activity by P. frosti from 1994 - 1999 showed that male participation varied over the duration of each breeding season and on a daily basis, and that weather conditions influenced calling rates and calling intensity, and the probability of it occurring (see Chapter 5). The results also showed that, of the total number of adult males known to be present at a breeding site, not all participated in calling activity at any one time, with maximum participation levels ranging from 44% to 78% during the peak of calling activity in each season. For surveys conducted between 1993 and 2002, daily variation in the detection of calling males was minimised by confining censuses to diurnal time periods (08:00 - 17:00 h) when calling participation was high, and abandoning censuses during periods of cold weather (≤ 5oC) when call rates and intensity were reduced (see Chapter 5). Because of the limited opportunity within which calling male censuses could be undertaken due to potentially adverse weather, no constraint was applied to censuses conducted during periods of relatively high temperature at sub-alpine elevations when the probability of recording calling activity was also reduced (Chapter 5). It was also assumed in this study that the intensity and timing of calling activity by P. frosti was geographically uniform over a breeding season. No obvious geographic variation in calling activity

39 by the species was detected during surveys conducted by Hollis (1995) in 1993 and 1994, and Malone (1985a) in 1983 and 1984.

For surveys conducted during each season between 1994 and 1999, seasonal variation in participation by calling males was corrected for by estimating the number of calling males that would have been recorded had each census been conducted during the peak of calling activity in each season, and if all adult males present were participating in calling activity. This method of correction assumes that the maximum number of calling males recorded by automatic recording units at breeding sites (Chapter 5) represents the total number of adult males present. Using smoothing curves depicting levels of participation by calling males over the duration of each breeding season (Fig. 5.4), the proportion of males calling at the time in which each census was undertaken was first estimated (Appendix 3.2). Census data from each survey transect were then corrected to 100% participation. For transects that were censused over more than one day, each census statistic was corrected independently relative to the census date, before being summed. Survey transects from which no calling males were recorded were not subject to correction. In a small number of cases (seven survey transects), calling males were recorded on days not encompassed by the smoothing curves, excluding the possibility of correcting these census statistics (see Appendix 3.2). An arbitrary minimum participation threshold of 10% was applied to these statistics because of the potential for error with calling participation estimates made at the steepest end of each smoothing curve.

During 1993, when recording units were not used to monitor calling activity at breeding sites, repeated censuses were conducted at approximately 1 - 2 week intervals within four survey transects above 1300 m elevation to monitor changes in numbers of calling male over the breeding season. The dates on which surveys were conducted at the four transects were: 17 and 29 November, 6 and 16 December at Baragwanath Flat; 11 and 24 November, 18 December at Pudding Basin; 15 and 24 November, 18 December at Macallister Plain; and 17 and 29 November, 6 and 13 December at Currawong Flat. The dates on which surveys were undertaken were allocated to four time periods for examination: (1) 11 - 17 November; (2) 24 – 29 November; (3) 6 December; and (4) 16 – 18 December. At each transect, a ‘proportion of males calling’ was derived for each survey date by dividing the largest number of calling males recorded during the season by the number recorded at each census (Appendix 3.2). These proportions were then averaged across each contributing transect to derive a mean estimate of calling male participation for each of the four time periods. Distance-weighted least squares smoothing was used to plot the relationship between the mean date for each time period and calling male participation. This relationship was then used to correct for seasonal variation in calling activity in 1993, as detailed above for recording units.

40 For surveys conducted by Malone (1985a) in 1983 and 1984, it was not possible to correct for seasonal variation in participation by calling males. Only raw census data have been used for comparison with the results of this study.

3.2.4 Analysis of Census Data and Detection of Trends

Because of the discovery of populations of P. frosti at lower montane elevations in 1996, and the apparent decline of the species from sub-alpine elevations, comparison and analysis of census data were conducted separately for three elevation groups to facilitate detection of trends across the distribution of the species. These included: (1) sub-alpine (> 1400 m a.s.l.); (2) sub-alpine- montane (1300 - 1400 m); and (3) montane (960 - 1299 m). Because not all survey transects in this study were examined in every year between 1993 and 2002, it was not possible to analyse data from all transects over this period. Subsets of transects from different time periods were subsequently chosen to maximise the number of transects analysed. Similarly, because not all survey transects examined by Malone (1985a) were censused in both 1983 and 1984, examining only those censused in both years would have reduced the sample size for analysis considerably. Sample size was subsequently increased by treating transect data from 1983 and 1984 as a single year (‘1983/84 composite’), using transect data collected in only one year, and averaging data collected over both years. This data manipulation was considered valid because of the large size of transect counts from both these years, relative to counts from 1993 - 2002, and that transects from which counts were obtained in both years were of a similar magnitude in most cases (see Appendix 3.1).

Four subsets of data were examined from sub-alpine elevation: (1) Raw data from seven transects censused for 12 years (1983, 1984, 1993 – 2002), with a 20-year monitoring duration; (2) Raw data from 14 transects censused for 11 years (1983/84 composite, 1993 - 2002), with a 19.5-year monitoring duration (the midpoint of 1983 and 1984 was used to derive monitoring duration when '1983/84 composite' was used); (3) Raw data from 14 transects censused for ten years (1993 – 2002), with a 10-year monitoring duration; and (4) Corrected data from 14 transects censused for seven years (1993 – 1999), with a seven-year monitoring duration (Table 3.1).

Four subsets of data were examined from sub-alpine-montane elevation: (1) Raw data from three survey transects censused for 12 years (1983, 1984, 1993 – 2002), with a 20-year monitoring duration; (2) Raw data from five transects censused for ten years (1993 – 2002), with a 10-year monitoring duration; (3) Corrected data from five transects censused for seven years (1993 – 1999), with a seven-year monitoring duration; and (4) Raw data from seven transects censused for seven years (1996 – 2002), with a seven-year monitoring duration (Table 3.1).

41 Five subsets of data were examined from montane elevation: (1) Raw data from four transects censused for 12 years (1983, 1984, 1993 – 2002), with a 20-year monitoring duration; (2) Raw data from four transects censused for ten years (1993 – 2002), with a 10-year monitoring duration; (3) Corrected data from four transects censused for seven years (1993 – 1999), with a seven-year monitoring duration; (4) Raw data from eight transects censused for seven years (1996 – 2002), with a seven-year monitoring duration; and (5) Raw data from 13 transects censused for six years (1997 – 2002), with a six-year monitoring duration (Table 3.1).

Spearman rank correlation coefficients (rs) were used to examine for possible trends in abundance of calling males (probability values were obtained using look-up tables). Correlation analysis was also used to examine the relationship between raw and seasonally corrected census data, pooled for years 1993 - 1999. Graphs depicting the relationship between numbers of calling males and survey year were also used to compare differences in trends portrayed by both raw and seasonally- corrected data for the different elevation groups.

3.2.4.1 Statistical Power

A common problem in trend detection is that the parameter associated with an ongoing trend is obscured by a source of variability (e.g., climate, season). The probability that a monitoring program will detect a trend in the sample counts when the trend is occurring, despite sources of variation within the count data, represents its statistical power (Gibbs 1995). Power analyses are considered useful for designing new monitoring programs, however, they can also be used to refine existing programs to meet the need of wildlife managers (Gerrodette 1987; Thomas 1997). Using the program MONITOR (version 6.2, Gibbs 1995), a retrospective power analysis (a posteriori) was used to compute statistical power (1 - β, where β = the probability of wrongly accepting the null hypothesis when it is actually false – Type II error ) for the subsets of monitoring data where no trend was detected. MONITOR uses Monte Carlo regression procedures to generate simulated sets of count data and sample counts drawn at random from distributions defined by the user (Gibbs 1995). Model inputs were: (1) number of survey transects; (2) initial count from each transect, or mean initial count across all transects when one or more transects had an initial count = 0; (3) standard deviation from counts across each transect; (4) plot weight = 1; (5) the occasion in which each transect was censused; (6) trend type = non-linear; (7) significance level α = 0.05; (8) constant = 1; (9) trend variation = default; (10) rounding = decimal; (11) replication = 1000 iterations. As there was no established ‘effect size’ from which to assess the extent of change in population size of P. frosti, a power value of 0.8 was used to identify ‘minimum detectable effect size’ (MDES) for magnitudes of change between -10 to +10%/year, and was considered adequate for statistical certainty of an acceptable Type II error (see Cohen 1988). A one-tailed test was used where it was believed a decline or increase in calling males had occurred, and two-tailed t-test where the direction of trend was unclear.

42 3.2.5 Estimating Adult Male Population Size

To account for population trends and differences in abundance of calling males across the distribution of the species when estimating the size of the adult male population, the known distribution of the species was stratified into six geographical areas based on elevation and drainage pattern (Fig. 3.3). Two regions were established based on drainage, including systems flowing either south-west or north-east off the Baw Baw Plateau, with the Alpine Walking Track providing an approximate boundary between the two. These regions were chosen because of observations by Malone (1985a) and Hollis (1995), who noted fewer breeding populations in the north-eastern compared to the south-western region. The three elevation groups adopted to examine population trends were also used as groups from which to derive frog density and population estimates, although the lower elevation limit on the north-eastern escarpment of the plateau was raised to 1200 m due to the absence of records of calling males below this elevation. The six geographical areas used were: (1) Montane (1200 – 1299 m) north-east; (2) Montane (960 – 1299 m) south-west; (3) Sub-alpine-montane (1300 – 1400 m) north-east; (4) Sub-alpine- montane (1300 – 1400 m) south-west; (5) Sub-alpine (> 1400 m) north-east; and (6) Sub-alpine (> 1400 m) south-west.

In the Montane (960 – 1299 m) south-west geographic area, the area boundary diverged from the 960 m contour on the north-west fringe to predominantly exclude drainage systems located in the Yarra Catchment, where a change from igneous (granite) to metamorphic geology occurs. In this location, the area boundary was chosen to align with the Toorongo and Thomson valley roads, which approximates this change in geology. An additional area of potentially suitable habitat between Link Road and the 960 m contour was also excluded from this geographic area because of its isolated nature, and because no calling males were recorded from this area (see Block 10 Myrrhee Ck survey transect, Fig. 3.2 – map code 13).

For each geographic area, a mean density of calling males was derived from a sample of survey transect counts, calculated from seasonally-corrected census data (see section 3.2.3), and the length of water course (stream length) within each survey transect. Drainage pattern was considered to provide the best available approximation of the extent of breeding habitat across the distribution of the species, from which density estimates could be calculated (number of calling males per metre of stream). The spatial layer containing information on drainage patterns used to estimate density of calling males was a refined version of hydrological mapping by Roberts (1996) (Dept of Natural Resources and Environment, 71 Hotham St, Traralgon, 3844). The total length of water course within each survey transect and geographic area was calculated using ArcView version 3.2a (ESRI- Environmental Systems Research Institute, Inc, Redlands, CA, 92373, USA). Drainage systems present within vegetation comprising of Montane Fen (see Davies 2000) at the north-western end of the Baw Baw Plateau (see Fig. 3.3) were excluded from the analysis, because they did not have established survey transects within them, and because they appeared not to

43 represent breeding habitat of P. frosti. Drainage systems also occurred in non-breeding habitats in other areas encompassing the breeding distribution of P. frosti. These were in the form of secondary water courses, or water courses located within non-breeding vegetation types (e.g., sub- alpine woodland, sub-alpine shrubland). At montane elevations, sections of stream that were fast- flowing and wide represented unsuitable breeding habitat for P. frosti, but because these were contained within transects used to estimate male density and population size, they had no influence on the derived estimates. At sub-alpine elevations (> 1300 m), however, the locations of water courses in vegetation types considered not to represent breeding habitat were, in most cases, contained outside established transects, and therefore not used to derive density estimates of males. These particular water courses (described as > 1300 m non-breeding habitat stream length in Table 3.6) were therefore excluded from the stream length measurements used to estimate adult male population size in each geographic area above 1300 m.

The results of trend analyses within the three elevation groups for data recorded between 1993 and 2002 were used to determine which survey years, transects and census data to use for estimating density of calling males. For groups where a significant trend was detected, the most recent corrected data from 1998 and 1999 were used, whilst for groups where a non-significant conclusion resulted, census data from all years in which surveys were undertaken between 1993 and 1999 were used. Census data from 2000 - 2002 were not used in the analysis because seasonal correction for male participation was not conducted.

3.3 Results

3.3.1 Distribution

Calling males were recorded from all land tenures, including newly-discovered, montane localities in State Forest, with widespread populations across the south-western escarpment as low as 960 m elevation and a record as low as 1200 m on the north-eastern escarpment in close proximity to the boundary of the National Park (Cascade Creek below 1300 m A, Fig 3.2, Fig. 3.4). Other populations were confined predominantly to the south-western region of the Baw Baw National Park and Mt Baw Baw Alpine Resort (Fig. 3.4). No calling males were recorded from one-off surveys conducted on the south-eastern escarpment of the Toorongo Plateau and headwaters of the Yarra water catchment (not illustrated). The area encompassing extant and historic records of P. frosti (Fig. 3.4) was estimated to be 93.4 km2 (100% of Minimum Convex Polygon), whilst the area encompassing the distribution of potential habitat of the species (Fig. 3.3) was estimated to be 134.5 km2.

44 3.3.2 Relative Abundance and Population Trends

3.3.2.1 Sub-alpine Elevation (>1400 m)

Analysis of the abundance of calling males recorded from survey transects censused between 1983 and 2002 indicates that a significant decline in the size of the population at sub-alpine elevations > 1400 m has occurred (Fig. 3.6, Table 3.2). The extent of this decline appears to be of several orders of magnitude, with post-1993 indices representing an approximate 98% decline from those recorded in 1983 and 1984. This trend is apparent in subsets of census data with both seven and 14 survey transects (Fig. 3.6). The negative coefficients of association for surveys conducted between 1993 and 2002, including the trend for seasonally-corrected data, suggests that the adult male population has continued to decline over this period (Table 3.2).

3.3.2.2 Sub-alpine-montane Elevation (1300 - 1400 m)

Survey transects censused between 1983 and 2002 also indicate that a decline in population size has occurred at elevations between 1300 and 1400 m, although this trend was not significant (Fig. 3.7, Table 3.2). In contrast, between 1993 and 2002 an increase in the number of calling males appears to have occurred. This trend was significant for all but one of the subsets of data examined, including the seasonally-corrected data (Table 3.2, Fig. 3.7). The graph illustrating raw and corrected types of census data shows a dramatic increase in the number of calling males in 1999. This increase was also recorded for transects surveyed at sub-alpine elevations > 1400 m above (Fig. 3.6). The magnitude of this increase was primarily due to the contribution of 29 calling males from ‘The Morass’ survey transect at a time when participation by calling males was estimated to be only 13.5% on the majority of the days when this transect was surveyed (Appendix 3.2). Correction of the raw census data resulted in a substantial increase in contribution by this transect from 29 to 176 calling males (Table 3.1).

3.3.2.3 Montane Elevation (960 - 1299 m)

Figure 3.8 illustrates that a decline in calling males may have also occurred at montane elevations (960 and 1299 m) between 1983 and 1993, although this trend was not significant (Table 3.2). Between 1993 and 2002, no trend in relative abundance of calling males was apparent within the data subsets examined, including the seasonally-corrected data. The recorded increase in abundance of calling males evident at other elevations in 1999 was not as distinct at montane elevations. A slight decrease in numbers of calling males appears to have occurred when examining the larger sample of transects from 1997 – 2002 (Fig. 3.8), but this was not significant (Table 3.2).

45 3.3.3 Participation of Calling Males in 1993

Although only from a small sample size of four repeat-census sampling points, the smoothing curve depicting levels of participation by calling males in 1993 suggests a non-linear relationship, with maximum participation of 80% occurring between approximately the 25 – 29 November, before and after which participation was less (Fig. 3.5).

3.3.4 Statistical Power and Effect Size

Power analyses show that for the data subset representing sub-alpine-montane elevation (1300 – 1400 m) from 1983 and 2002, there was a high certainty (86%) that monitoring could detect a declining trend of 5%/year (Table 3.2), although the correlation coefficient was only poor (rs = - 0.32). The result was similar for the data subset representing montane elevation (960 – 1299 m) from 1983 – 2002 (rs = -0.26), where there was a high certainty (84%) that monitoring could detect a declining trend of 3%/year (Table 3.2). Although not significant, a stronger correlation coefficient (rs = 0.63) was derived for the data subset representing sub-alpine-montane elevation (1300 - 1400 m) from 1996 – 2002, where there was a high certainty (83%) that monitoring could detect an increasing trend of 9%/year (Table 3.2). Similarly, a correlation coefficient of -0.54 was derived for the data subset representing montane elevation (960 – 1299 m) from 1997 – 2002, where there was a high certainty (85%) that monitoring could detect a declining trend of 10%/year (Table 3.2). All other montane subsets of data examining trends between 1993 and 2002, where a non-significant trend was obtained, had low power and therefore a high probability of Type II errors occurring.

3.3.5 Comparison of Raw and Seasonally-corrected Census Data

Raw and seasonally-corrected data, pooled for years 1993 – 1999, were significantly correlated with each other (rs = 0.89, n = 248, p < 0.0001, Fig. 3.9). This correlation indicates that overall, trends in the relative abundance of calling males, as revealed by the raw census data, were not dissimilar from trends in abundance of calling males following seasonal correction. This relationship is illustrated in Figures 3.6, 3.7 and 3.8, with seasonally-corrected data in most cases being proportionally greater than the raw data in each year.

3.3.6 Population Size of Adult Males

Details of individual survey transects used to derive seasonally-corrected densities of calling males within each of the six geographic areas are contained in Tables 3.3, 3.4 and 3.5. Recent census data from 1998 and 1999 were used to derive density estimates for sub-alpine (> 1400 m) and sub-

46 alpine-montane (1300 – 1400 m) geographic areas where population trends were detected (above). For the sub-alpine (> 1400 m) north-east area, using census data from just 1998 and 1999 would have resulted in a reduced sample of survey transects (5) from which to estimate density of calling males. To increase the sample size in this area, survey transects censused in 1993, 1995, and 1997 were also included (Table 3.3). Census data from all years (1996 – 1999) were used to estimate density in montane (1200 – 1299 m) north-east and montane (960 – 1299 m) south-west geographic areas (Table 3.5) where no population trends were detected.

The sum total of adult male frogs from all six geographic areas was estimated to be 6,728 (Table 3.6). The montane (960 – 1299 m) and sub-alpine-montane (1300 – 1400 m) geographic areas located on the south-western escarpment of the Baw Baw Plateau comprised 86% of the total estimated number of adult males, with the remaining 14% of the population from the other four geographic areas (Table 3.6). The geographic area with the smallest estimated adult male population (5) was the montane (1200 – 1299 m) area located on the north-eastern escarpment of the Baw Baw Plateau. Estimates for both sub-alpine (> 1400 m) north-east and south-west geographic areas, and the remaining sub-alpine-montane (1300 – 1400 m) area located on the north-eastern escarpment, ranged between 261 and 334 adult males. Although only comprising 21% of the total estimated adult male population, the sub-alpine-montane (1300 – 1400 m) geographic area located on the south-western escarpment was estimated to have the highest density of calling males per metre of stream (26.6/km) (Table 3.6). The sub-alpine (> 1400 m) north-east area unexpectedly contained the second highest density of calling males (12.3/km), although the sample size from this area, even with the addition of survey transects from other years, remained small compared to other areas examined (Table 3.3). The montane (960 – 1299) south-west area contained the next highest density of calling males (11.7/km).

3.4 Discussion

3.4.1 Statistical Power

Power analyses show that there has been adequate sampling from four subsets of data to detect changes in population size of less than 10% where a non-significant trend was recorded (see results), but insufficient sampling for the remaining subsets of data from montane elevation to conclude that no trends were present (Table 3.2). Connell and Sousa (1983) argue that to demonstrate population stability, or that a population has become extinct, it must be monitored for at least one population turn-over. However, Stirrat et al. (2001) point out that the relevant life- history parameter is generation time, defined as the mean time between birth of parents and the birth of their offspring. This study has examined survey data spanning a period longer than estimates of longevity (14.5+ years) and maturation (3.5 years for males and 4.5/5.5 years for

47 females) for P. frosti (Chapter 7), spanning 20 years (1983 - 2002). As breeding populations of P. frosti were discovered only recently (1996) in lower-elevation habitats, a lack of power for a number montane subsets of data examined is due primarily to the currently low monitoring duration relative to the number of survey transects and monitoring variances observed. Between 1993 and 1996, only four survey transects located at montane elevations (1160 - 1250 m) had been censused, these being historical sites examined by Malone (1985a) in 1983 and 1984. Continued monitoring of the 13 post-1996 survey transects appears to have adequate power to examine changes in population size of less than 10% at montane elevations in the future (Table 3.1).

3.4.2 Distribution

Records of P. frosti from this study, and those collected historically, show that the distribution of the species is confined entirely to the granodiorite Baw Baw land system (see Aldrick et al. 1992) where organic soils (comprised of raw and decomposing vegetation) develop in valleys and hillside drainage lines (Sibley 1975). Although contained within the Baw Baw land system, no breeding populations of P. frost were recorded during surveys conducted on the Toorongo Plateau. It was also noted during surveys of the Toorongo Plateau that the granite geology was limited, with soils being derived predominantly from weathered, metamorphosed sedimentary rocks. An absence of breeding populations in the headwaters of the Yarra catchment at the north-western and northern end of the Baw Baw Plateau also seems to be related to a change in soil type from organic to mineral soils.

On the south-western escarpment of the Baw Baw Plateau, there appears to be a close association between the lower-elevation limit of P. frosti (approximately 960 m) and a change in vegetation type from forest dominated by Alpine Ash (Eucalyptus delegatensis) or Shining Gum (E. nitens) (montane wet forest, cool temperate rainforest and cool temperate mixed forest – see Gilbert 1959; LCC 1991; Peel 1999) to forest dominated by Mountain Ash (E. regnans) (wet forest - LCC 1991). At this elevation, the saturated, organic soils used by the species for breeding also appears to recede. This pattern of distribution is not apparent on the north-eastern escarpment of the plateau, where the species is confined to elevations above approximately 1200 m, even though suitable breeding habitat extends to lower elevations similar to the south-western escarpment. At 1200 m elevation, the vegetation type generally changes from a sub-alpine to a montane floristic composition.

The absence of breeding populations on the north-eastern escarpment at elevations below 1200 m suggests that climate may be a limiting factor. According to DCE (1991), the Baw Baw Plateau exerts a pronounced effects on the local climate, intercepting rain-bearing westerly and south- westerly winds. This observation suggests that the north-eastern escarpment of the plateau may be climatically drier than the south-western escarpment, and that its proximity to the plateau may locate it within a rain-shadow. This climatic pattern is supported by a study on post-frontal rain in

48 the vicinity of the Baw Baw Plateau, showing that a ‘cap cloud’, associated with connective cells with high liquid water content, is formed above Mt Baw Baw and on the western escarpment (Abbs and Jensen 1993). The types of vegetation communities present on the north-eastern side of the plateau also suggest that it is drier, with larger areas of montane damp forest present at montane elevations, and large areas of sub-alpine shrubland and granite out-cropping present at elevations above 1200 m (see mapping by Roberts 1996 and descriptions by LCC 1991). The results of Chapter 4 also show preference by P. frosti for breeding habitats that are topographically protected, cool and moist. If the north-eastern side of the Baw Baw Plateau is significantly drier and warmer than other plateau areas, this could explain both the absence of breeding populations at montane elevations, and the smaller breeding populations recorded at sub-alpine elevations in this study. Except for the sub-alpine (> 1400 m) north-east geographic area, which had a low sample size, density estimates for calling males were higher for south-western regions (Table 3.6). This density pattern was also observed by Malone (1985a) who noted during his 1984 survey that breeding was largely restricted to the windward side of the Baw Baw Plateau, with fewer choruses being heard on the lee side.

A contraction in distribution range of calling males from high-elevation areas in the north-eastern region of the plateau, compared to the distribution reported by Malone (1985a), was also recorded from surveys conducted in this study. In addition to this, an extension of the known distribution range of the species, from predominantly high-elevation habitats > 1300 m to montane habitats on the south-western escarpment of the plateau as low as approximately 960 m was also recorded. Malone (1985a) estimated the area of the Baw Baw Plateau from which he recorded breeding populations of P. frosti to be approximately 80 km2. This estimate contrasts with estimates of 93.4 km2 and 134.5 km2 from this study, which include the area encompassing historic and current records of P. frosti and potential habitat of the species, respectively.

3.4.3 The Use of Counts of Calling Males for Monitoring and Estimating Population Size

This chapter has endeavoured to assess the suitability of the monitoring program established for P. frosti for the purpose of detecting population changes across the species geographic range, over time. There are only a limited number of documented cases within the Australian frog fauna where systematic surveys have covered the entire, known geographic or environmental range of a taxon (e.g., Malone 1985a; Osborne 1989; Wardell-Johnson and Roberts 1993; Gillespie and Hollis 1996; Roberts et al. 1997; Parris 2001). For a number of these surveys, the use of counts of calling males has been effectively used as a method to examine distribution and relative abundance (Malone 1985a; Osborne 1989; Roberts et al. 1997; Parris 2001). However, although considered a simple and cost-effective method for monitoring relative abundance, its use has also been considered problematic due to the possibility of turn-over of calling males, the presence of silent

49 and satellite males, and potential variation in participation by calling males on a daily and seasonal time scale (Zimmerman 1994; Driscoll 1998).

In using counts of calling males to monitor relative abundance and estimate population size of P. frosti, this study has attempted to correct for some of the aforementioned concerns: (1) seasonal variation in participation by calling males; (2) daily variation in participation by calling males; (3) variable climatic conditions; and (4) geographic variation in abundance. No attempt was made to estimate and correct for turn-over of calling males at breeding sites. However, in the absence of this information for P. frosti, it may be possible to indirectly assess the potential contribution of male turn-over to the population using other sources of data. Driscoll (1998) discussed a number of attributes of calling male Geocrinia alba and G. vitellina that related to the low turn-over of males he recorded using mark-recapture and removal studies. These were: (1) persistent calling activity by males, with most individuals likely to be calling during the peak of the calling season; and (2) extreme philopatry, with most males being displaced by less than 5 m during the breeding season, resulting in a reduced possibility of new individuals entering a breeding site and few individuals leaving a breeding site. Analysis of the movement patterns of adult male P. frosti showed that it also had relatively low dispersal tendencies during the breeding season, being confined to home ranges of less than 11 m2 (Chapter 6), and that males were relatively persistent in calling activity over the duration of the breeding season. These attributes suggest that male turn- over in P. frosti may be relatively low. However, observations by Malone (1985a) on the presence of silent males, and the frequent occurrence of more than one male at calling sites, suggests that there may be an additional portion of the male population of P. frosti that could be attributed to these individuals. In this study, more than one calling male was also recorded incidentally at some calling sites that were inspected, but their occurrence would not have been categorised as frequent. Perhaps the enormity of calling male counts recorded by Malone in 1983 and 1984 contributed proportionally to the frequent observation of silent males, compared to this study when much smaller counts were recorded. It is also possible that male turn-over is reduced in terrestrial species such as P. frosti compared with aquatic species where populations are known to fluctuate more (Alford and Richards 1999; Marsh 2001).

This study’s total population estimate of 6,728 adult males excludes any consideration of satellite males that may have been present. However, it is possible that the attendance of these males at breeding sites may have been accounted for, to some extent, during the process of seasonal correction, if they participated in calling activity at breeding sites where recording units were placed. Like other age classes not examined in this study, perhaps the portion of the population relating to satellite males should not be considered in an estimate of adult males, but separately as immature, sub-adult or subordinate males. Malone (1985a) multiplied his count of 4,248 calling males by approximately three to account for silent and satellite males in his estimate of the adult male population above 1300 m. In this study, the 2,352 adult males estimated to be present above 1300 m elevation does not include silent males. Although Malone surveyed a large extent of the

50 area considered to include breeding habitat for P. frosti (i.e., frost hollow areas), a considerable area of breeding habitat within montane gullies between 1300 and 1400 m elevation was not surveyed. It is within these breeding habitats that the largest portion of males contributing to this study’s estimate above 1300 m was recorded (72%). This observation suggests that Malone’s count of 4,248 males, and subsequent population estimate, was considerably lower than what would have been present at sub-alpine elevations at the time.

There are a number of other factors that may have influenced estimates of density and population size in this study. The first of these includes the method used for correcting census data for seasonal variation in participation by calling males. Because there was a reasonable degree of scatter in the points depicting participation variation over the duration of each season, and that correction of raw census data was based on an estimate of the trend (distance-weighted smoothing curves in Chapter 5, Fig. 5.2), it is possible that some census data were over, or under-estimated. Once source of potential error not accounted for could have originated from small-scale (site-to site) differences in calling participation, which were evident but not examined in Chapter 5. Another factor that may have influenced estimates was the assumption that the calling data used to derive smoothing curves depicting male participation in each year was representative of the geographic distribution of calling males (large-scale differences). However, because transect censuses were undertaken at the beginning of each season for reconnaissance, and continued until no further calling males were being recorded at the end of each season, it was possible to compare dates from which calling males were first and last recorded in each season with that of commencement and completion dates estimated by the smoothing curves in each year. This comparison shows that the smoothing curves portray a good representation of calling activity duration from different locations and habitat types, in each year. Only two survey transects, out of all survey transects censused in all years, had calling males recorded from them on days not encompassed by the smoothing curve areas. In these cases, calling activity was recorded one day beyond that estimated by the smoothing curve. A further five transects had calling males recorded from them on days that the smoothing curves suggested completion of calling activity (Appendix 3.2). The selection of a minimum participation estimate of 10% to be applied during correction of census data from these transects is therefore probably justified.

Possible over and under-estimation of census data would have been particularly relevant to survey transects that were censused near to the commencement or completion of each breeding season when smoothing curves were steepest, and the likelihood of a larger error, therefore increased. When comparing both raw and corrected census statistics for individual survey transects over each year, a number of examples where over-estimation may have occurred include The Morass (1999), Cascade Creek (below 1300 m A) (1997), Faith Creek (1998) and Whitelaw Creek (frost hollow) (1995). In particular, the corrected census statistic from the 1999 survey of The Morass transect was so prominent that it influenced the overall seasonally corrected statistic in 1999 considerably (Fig. 3.7).

51 The use of raw counts of calling males was shown to depict similar trends compared with the seasonally-corrected census data (Figs 3.6 – 3.8, Table 3.2). As a method for assessing broad trends in population, raw counts appear to be suitable for measuring changes in relative abundance of P. frosti over time. However, a discrepancy in the use of raw counts may arise when the following conditions occur in combination: (1) survey transects that are censused at the beginning or end of a breeding season when participation by calling males is considerably reduced; (2) survey transects that are located in areas with a high density of calling males (e.g., sub-alpine- montane [1300 – 1400 m] south-west geographic area); and (3) survey transects that are of a large length or size. Raw census data collected from such transects are more likely to under-estimate numbers of calling males by a considerable amount compared with shorter or smaller transects located in areas with a lower density of calling males. Seasonally-corrected census data therefore provide a more realistic estimate of the total number of males present relative to actual counts, as well as reducing seasonal variability due to differences in the time that censuses are undertaken.

3.4.4 Population Trends and Conservation Status

This study confirms the decline and contraction in range of the population of P. frosti from high- elevation, sub-alpine habitats (> 1400 m) on the Baw Baw Plateau reported by Hollis (1995) and Osborne et al. (1999), and suggests probable declines from elevations between 1300 and 1400 m and 960 – 1300 m as well. This result contributes to the list of cases reporting of the decline or disappearance of amphibian populations from high altitude, montane and mountain-top environments (e.g., Osborne 1989, 1990a, 1992; Bradford 1991; La Marca and Reinthaler 1991; Crump et al. 1992; Richards et al. 1993; Pounds et al. 1997; Lips 1998, 1999; Alexander and Eischeid 2001; Stallard 2001). By comparison, Hines et al. (1999) report that there are no declines evident in other members of the genus Philoria in north-eastern New South Wales and South- eastern Queensland, although the data for some species appears limited.

The recorded decline of P. frosti from elevations above 1300 m represents an approximate 98% decrease in the relative abundance of calling males compared to counts recorded by Malone (1985a) in 1983 and 1984. A comparison of the population estimate made by Malone (1985a) (a maximum of 15,000 adult males) and the corrected estimate from the same area this study (2,352 adult males from elevations above 1300 m, Table 3.6) indicates that the male population has declined by 84%. At elevations between 1300 and 1400 m, this study also suggests that a decline has occurred since the survey of Malone, although the small sample size of transects used to examine for trends could have resulted in the non-significant statistic (Table 3.2). Inspection of all census data from this elevation, although not obtained in every year, alludes to a decline of similar magnitude to that recorded at elevations above 1400 m (see Appendix 3.1).

52 Survey data collected between 1993 and 2002 suggest a continuation of the decline in adult males from elevations above 1400 m. In contrast to this, an increase in males appears to have occurred at elevations between 1300 and 1400 m, although this is only a very slight increase compared to the high numbers recorded by Malone in 1983 and 1984 (Fig. 3.7). At montane elevations (960 – 1299 m), the distribution of P. frosti has been extended considerably compared to that which was known prior to this study (Fig. 3.4). This range extension has occurred almost entirely on the south- western escarpment of the Baw Baw Plateau, with a small number of calling males being recorded as low as 1200 m on the north-eastern escarpment in the catchment containing South Cascade Creek. It can not be determined with certainty if this montane population has suffered a similar decline to that recorded above 1300 m. Four survey transects censused along the access road to the Mt Baw Baw Alpine Resort in this study, and by Malone in 1983 and 1984, suggest that a decline at montane elevations could have also occurred (Fig. 3.8), but this was not statistically significant (Table 3.2).

In addition to estimating the population size of adult P. frosti, Malone (1985a) also attempted to indirectly estimate the population size of adult female P. frosti by endeavouring to determine the sex ratio of clutches of recently-metamorphosed frogs. However, histological examination of the gonads showed that they remained undifferentiated at this stage of development, and therefore precluded the use of secondary sex ratios for estimating female population size. By determining that the number of clutches deposited by G. alba at two breeding sites was approximately the same as the estimated number of adult males, Driscoll (1996) concluded that the species probably had a sex ratio of 1:1, assuming that most females produce a single clutch each year. Unfortunately, data are not available to determine if the number of clutches deposited by P. frosti at a breeding site is approximately the same as the number of calling males present, and if only a single clutch is produced. Observations made during this study show that not all calling males were successful in producing a clutch of eggs, whilst other males were successful in producing multiple clutches (e.g., one male radio-tracked over the breeding season produced three clutches). Duellman and Trueb (1994) point out that a 1:1 sex ratio in anurans is not unusual. If this is the case for P. frosti, then the extant adult population might total approximately 14,000. ______This chapter confirmed that the population of P. frosti has undergone a significant decline and contraction in range at sub-alpine elevations (> 1300 m), and may have also declined at lower, montane elevations (960 – 1300 m), where previously unknown populations were recorded on the south-western and north-eastern escarpment of the Baw Baw Plateau. The results of monitoring between 1993 – 2002 suggest a continuation of the decline of P. frosti at elevations above 1400 m, whilst populations between 960 and 1400 m appear to have remained relatively stable. It is not known if recently-discovered populations at montane elevations have declined to the same extent as at sub-alpine elevations. Counts of calling males represent a satisfactory survey technique for monitoring and estimating abundance of P. frosti. However, if not accounted for, temporal variation in activity by calling males can result in under-estimation of abundance. Retrospective

53 power analyses show that sufficient effort was attained to detect changes in populations size of less than 10% at all elevations examined, except for subsets of data that comprised a small number of survey transects.

Analysis of the distribution and abundance of P. frosti showed that the highest density of calling males occurred on the south-western escarpment of the Baw Baw Plateau between 1300 and 1400 m, and the largest number between 960 and 1300 m. The lowest density of males occurred at elevations above 1400 m, and on the north-eastern escarpment between 1200 and 1400 m. The size of the adult male population was estimated to be approximately 7000. Due to the magnitude of the decline from sub-alpine habitats, and possible decline from other elevations, the conservation status listing of P. frosti as critically endangered (VDSE 2003) and endangered (Tyler 1997) is defensible (see the 2000 IUCN Red List of Threatened Species).

54 Table 3.1. Survey transects and census data of calling males used to examine for population trends between 1983 and 2002.

Survey transects are grouped by their elevational distribution, where S = sub-alpine (> 1400 m), SM = sub-alpine-montane (1300 - 1400 m), and M = montane (960 - 1299 m). Prefix identifiers: R= raw census data, C = seasonally-corrected census data. Year abbreviations: 83 = 1983, 84 = 1984, 8384 = 1983/1984 composite, etc. Census data presented for R8384 represents an average of the summed value for R83 and R84. N.B. Historical locations comprising Access Road 2 and Access Road 3 transects are small subsets of Hope Creek (Access Rd 2 above and below rd and Access rd 3 above and below rd), and have not been analysed together.

Survey Transect Number of Calling Males Elevation R83 R84 R8384 R93 C93 R94 C94 R95 C95 R96 C96 R97 C97 R98 C98 R99 C99 R00 R01 R02 Group Access Road 1 M 43 21 32 1 3.57 2 4.44 1 3.70 1 2.35 0 0.00 2 3.92 1 1.65 1 0 2 Access Road 2 M 30 26 28 0 0.00 0 0.00 0 0.00 0 0.00 0 0.00 0 0.00 2 3.31 2 0 2 Access Road 3 M 0 0 0 0 0.00 0 0.00 0 0.00 0 0.00 0 0.00 3 6.74 2 3.31 0 0 3 Barnies Creek M 0 0.00 0 0.00 8 13.30 3 7.30 8 0 1 Chairlift Corner M 6 8 7 0 0.00 1 4.00 2 4.60 0 0.00 0 0.00 0 0.00 0 0.00 0 0 0 Charity Creek (right branch below rd) M 5 7.40 0 0.00 2 10.80 3 2 2 Charity Creek (right branch below Ck Corner) M 00.0000.001 1.78 0 3 Hope Creek (access rd 2 above rd) M 00.0000.000 0.007 0 1 Hope Creek (access rd 2 below rd) M 12 25.00 4 7.84 9 14.88 0 1 1 Hope Creek (access rd 3 above rd) M 2 4.20 10 22.50 14 23.20 12 0 5 Hope Creek (access rd 3 below rd) M 5 7.41 2 3.92 8 13.22 0 0 3 Long Creek M 9 46.64 53 80.30 19 31.40 9 17.65 14 31 27 Long Creek (left branch side Ck) M 21 31.82 28 46.28 17 33.33 8 18 38 Tanjil River West Branch 2 (above aqua.) M 4 40.00 18 50.00 30 50.42 27 81.82 28 47 13 Tyers River East Branch (tributary) M 0 0.00 0 0.00 1 8.33 0 0.00 1 0 0 Baragwanath Flat S 167 245 206 9 11.84 4 9.88 4 8.89 0 0.00 1 2.90 1 2.08 5 6.17 0 0 0 Currawong Flat/Sandys Flat/Pauciflora Flat S 536 536 8 10.26 2 4.44 5 11.90 2 3.70 0 0.00 2 3.80 7 8.60 2 4 2 Currawong Flat S 174 231 202.5 2 2.63 1 2.22 0 0.00 0 0.00 0 0.00 0 0.00 2 2.45 0 0 0 Freemans Flat/Tullicoutty S 41 104 72.5 0 0.00 0 0.00 0 0.00 0 0.00 0 0.00 1 1.67 2 2.65 1 0 1 Gwinear Flat S 93 2 47.5 0 0.00 0 0.00 0 0.00 0 0.00 0 0.00 0 0.00 0 0.00 0 0 0

55 Survey Transect Number of Calling Males Elevation R83 R84 R8384 R93 C93 R94 C94 R95 C95 R96 C96 R97 C97 R98 C98 R99 C99 R00 R01 R02 Group La Trobe Plain S 206 206 3 3.95 2 5.48 0 0.00 4 8.16 5 7.41 0 0.00 1 1.27 2 0 2 Macallister Plain S 82 82 5 6.49 2 4.94 0 0.00 1 1.85 0 0.00 0 0.00 0 0.00 0 0 0 McMillians Flat S 64 64 0 0.00 0 0.00 0 0.00 0 0.00 0 0.00 0 0.00 0 0.00 0 0 0 Moondarra Flat S 225 225 0 0.00 0 0.00 0 0.00 0 0.00 0 0.00 0 0.00 1 1.27 0 0 0 Mustering Flat S 57 0 28.5 0 0.00 0 0.00 0 0.00 0 0.00 0 0.00 0 0.00 2 5.48 1 0 0 Neulyne Plain S 24 24 2 2.63 2 5.48 1 2.27 0 0.00 1 1.48 0 0.00 0 0.00 1 0 0 Pudding Basin S 101 101 2 2.60 1 2.47 1 2.27 0 0.00 0 0.00 0 0.00 0 0.00 0 0 0 Tanjil Plain S 120 120 9 11.54 3 5.94 3 7.50 3 5.56 1 2.90 1 1.82 2 2.47 1 0 0 Tyers River 2 frost hollow S 1.8 1.3 1.55 0 0.00 0 0.00 0 0.00 0 0.00 0 0.00 0 0.00 0 0.00 0 0 0 Village Flat S 183 149 166 1 2.38 0 0.00 0 0.00 0 0.00 0 0.00 0 0.00 0 0.00 0 0 0 Barnies Creek SM 0 0.00 0 0.00 0 0.00 0 0.00 0 0 1 Creek Corner SM 52 49 50.5 0 0.00 0 0.00 2 5.00 4 8.00 0 0.00 2 3.92 2 10.81 4 0 1 Creek Corner 2 SM 9 12 10.5 5 12.50 0 0.00 1 1.61 2 3.92 2 10.81 4 0 9 East Tanjil SM 71 71 1 1.28 0 0.00 0 0.00 2 3.70 0 0.00 1 1.67 5 6.25 7 0 0 The Morass (2 sections excluded) SM 667 667 30 38.46 26 55.32 10 39.22 23 51.69 35 63.60 31 53.00 29 176.60 44 35 49 Tyers River 2 frost hollow SM 5.3 1.7 4 0 0.00 0 0.00 0 0.00 0 0.00 0 0.00 0 0.00 0 0.00 0 0 0 Tyers River 1 frost hollow SM 14 9 11.5 0 0.00 0 0.00 0 0.00 0 0.00 0 0.00 0 0.00 0 0.00 0 0 0

56 Table 3.2. Relationship between abundance of calling males (R = raw census data, C = seasonally-corrected census data) and survey year for different monitoring programs (data subsets) at different elevation (S = sub-alpine, SM = sub-alpine-montane, M = montane).

Spearman rank correlation coefficients (rs) are presented as a measure of association. For monitoring programs with a non-significant conclusion (n.s), statistical power (1-β) is presented for Minimum Detectable Effect Size (MDES) values ranging between 10 and -10%/year. Power values below 0.8 indicate a lack of confidence in concluding that no trend was present (see methods). For data subsets comprising 1983/84 composite survey data, the midpoint of 1983 and 1984 was used to derive monitoring duration.

Data Subset Data Elevation No. of Survey Years Monitoring rs p Power MDES t-test Status Group Transects Monitored Duration (years) (1-β) 1983, 1984, 1993 - 2002 R S 7 12 20 -0.72 < 0.05 - - - 1983/84 composite, 1993 - 2002 R S 14 11 19.5 -0.81 < 0.05 - - - 1993 - 2002 R S 14 10 10 -0.74 < 0.05 - - - 1993 - 1999 C S 14 7 7 -0.79 < 0.05 - - - 1983, 1984, 1993 - 2002 R SM 3 12 20 -0.32 n.s 0.859 5% decrease 1-tailed 1993 - 2002 R SM 5 10 10 0.82 < 0.05 - - - 1993 - 1999 C SM 5 7 7 0.86 < 0.05 - - - 1996 - 2002 R SM 7 7 7 0.63 n.s 0.828 9% increase 1-tailed 1983, 1984, 1993 - 2002 R M 4 12 20 -0.26 n.s 0.842 3% decrease 1-tailed 1993 - 2002 R M 4 10 10 0.30 n.s 0.200 10% increase 2-tailed 0.084 10% decrease 1993 - 1999 C M 4 7 7 0.04 n.s 0.148 10% increase 2-tailed 0.090 10% decrease 1996 - 2002 R M 8 7 7 0.25 n.s 0.506 10% increase 2-tailed 0.450 10% decrease 1997 - 2002 R M 13 6 6 -0.54 n.s 0.849 10% decrease 1-tailed

57 Table 3.3. Details of survey transects used to derive a density of calling males at sub-alpine elevation (> 1400 m). Survey transects are grouped by south- west (sw) and north-east (ne) regions.

* denotes survey transects from which census data were used from other years when not censused in 1998 and 1999 (see results).

Survey Transect Total Stream Region 1998 Seasonally- 1999 Seasonally- Mean No. of Mean No. of Length (m) corrected No. of corrected No. of Calling Calling Males/km Calling Males Calling Males Males Cascade Ck frost hollow 311 ne 0.00 0.00 0.00 Cascade Ck frost hollow side branch 69 ne 9.10 9.10 132.24 Gwinear Flat 4169 ne 0.00 0.00 0.00 0.00 Jeep Track 2* (1993) 295 ne 0.00 0.00 Jeep Track 3* (1993) 182 ne 0.00 0.00 Little Boys A 1031 ne 0.00 0.00 0.00 Mustering Flat 6594 ne 0.00 5.50 2.75 0.42 Talbot Ck frost hollow* (1997) 1056 ne 0.00 0.00 Whitelaw 1* (1995) 669 ne 0.00 0.00 Whitelaw 2* (1995) 1016 ne 15.3 15.06 Whitelaw Ck (frost hollow) – ne section* (1995) 3650 ne 0.00 0.00 Whitelaw Ruins* (1995) 3438 ne 0.00 0.00 Baragwanath Flat 4109 sw 2.10 6.20 4.15 1.01 Chairlift + Corner 276 sw 0.00 0.00 0.00 Creek Corner 2 68 sw 0.00 5.41 2.71 39.84 Currawong Flat/Sandys Flat/Pauciflora Flat 8742 sw 3.80 8.60 6.20 0.71 East Tanjil 406 sw 1.70 0.00 0.85 2.09 Freeman Flat/Tullicoutty 4730 sw 1.70 2.60 2.15 0.45 Hope Ck (access rd 3 right branch)) 158 sw 0.00 0.00 0.00 0.00 Hope Ck (access rd 2 above rd) 137 sw 0.00 0.00 0.00 0.00 Hope Ck left branch 196 sw 0.00 0.00 0.00 0.00

58 Survey Transect Total Stream Region 1998 Seasonally- 1999 Seasonally- Mean No. of Mean No. of Length (m) corrected No. of corrected No. of Calling Calling Males/km Calling Males Calling Males Males Hope Ck 162 sw 1.67 19.51 10.59 65.21 La Trobe Plain 785 sw 0.00 0.00 0.00 0.00 Macallister Plain 703 sw 0.00 0.00 0.00 0.00 McMillians Flat 687 sw 0.00 0.00 0.00 0.00 Moondarra Flat 1158 sw 0.00 1.30 0.65 0.56 Neulyne Plain 395 sw 0.00 0.00 0.00 0.00 Pudding Basin 421 sw 0.00 0.00 0.00 0.00 Tanjil Plain 849 sw 1.80 2.50 2.15 2.53 Tyers River 1 frost hollow 286 sw 0.00 0.00 0.00 0.00 Tyers River 2 frost hollow 158 sw 0.00 0.00 0.00 0.00 Village Flat 2 559 sw 0.00 1.30 0.65 1.16 Village Flat/Dam Valley/Big Hill 640 sw 0.00 0.00 0.00 0.00 Wombat Flat 65 sw 0.00 13.70 6.85 105.18

59 Table 3.4. Details of survey transects used to derive a density of calling males at sub-alpine-montane elevation (1300 – 1400 m).

Survey transects are grouped by south-west (sw) and north-east (ne) regions.

Survey Transect Total Stream Region 1998 Seasonally- 1999 Seasonally- Mean No. of Mean No. of Length (m) corrected No. of corrected No. of Calling Calling Males/km Calling Males Calling Males Males Bell Clear Ck A 1411 ne 0.00 0.00 0.00 Cascade Ck frost hollow 1210 ne 0.00 0.00 0.00 Cascade Ck FH side branch 1219 ne 9.10 9.10 7.46 Cascade Ck side branch A 329 ne 36.36 36.36 110.61 Little Boys 3049 ne 0.00 0.00 0.00 Little Boys A 1439 ne 0.00 0.00 0.00 Little Girls Ck B 1179 ne 0.00 0.00 0.00 Mustering Flat 533 ne 0.00 0.00 0.00 0.00 North Cascade 2 298 ne 0.00 0.00 0.00 Rum Ck 1217 ne 0.00 0.00 0.00 South Cascade 687 ne 0.00 0.00 0.00 South Cascade A 746 ne 0.00 0.00 0.00 South Cascade B 841 ne 0.00 0.00 0.00 Whitelaw Ck (right branch A) 1494 ne 0.00 0.00 0.00 Whitelaw Ck (right branch B) 392 ne 0.00 6.62 3.31 8.45 Whitelaw Ck 1 678 ne 0.00 0.00 0.00 0.00 Whitelaw Ck 2 450 ne 0.00 0.00 0.00 Barnies Ck 343 sw 0.00 0.00 0.00 0.00 Chairlift + Corner 367 sw 0.00 0.00 0.00 Charity Ck (right branch above rd) 539 sw 15.70 32.43 24.07 44.62 Charity Ck (side Ck D) 600 sw 5.00 3.31 4.16 6.94 Charity Ck (side Ck F) 247 sw 0.00 6.61 3.31 13.40 Charity Ck left branch top 549 sw 3.40 3.40 6.19

60 Survey Transect Total Stream Region 1998 Seasonally- 1999 Seasonally- Mean No. of Mean No. of Length (m) corrected No. of corrected No. of Calling Calling Males/km Calling Males Calling Males Males Chinamans 6 947 sw 0.00 0.00 0.00 Creek Corner 566 sw 3.90 10.80 7.35 13.00 Creek Corner 2 414 sw 3.90 5.41 4.66 11.25 East Tanjil 693 sw 0.00 3.75 1.88 2.71 Faith Ck 78 sw 29.63 29.63 378.88 Faith Ck left branch 308 sw 0.00 0.00 0.00 Hope Ck (access rd 3 above rd) 345 sw 11.24 26.45 18.85 54.59 Hope Ck (access rd 3 right branch) above chairlift 1024 sw 0.00 5.41 2.71 2.65 Hope Ck (access rd 2 above rd) 793 sw 0.00 1.65 0.83 1.05 Hope Ck 395 sw 1.67 26.83 14.25 36.06 Hope Ck left branch 640 sw 0.00 7.32 3.66 5.72 La Trobe Plain 285 sw 0.00 1.30 0.65 2.28 The Morass 17540 sw 58.60 228.40 143.50 8.18 Tanjil River East Branch 216 sw 12.50 12.50 57.81 Tanjil River west Branch 5 above aqueduct 268 sw 3.50 3.50 13.06 Tanjil River west Branch 6 above aqueduct 592 sw 14.04 14.04 23.71 Tanjil River west Branch 9 above aqueduct 146 sw 0.00 0.00 0.00 Tyers River 1 frost hollow 1259 sw 0.00 0.00 0.00 0.00 Tyers River 2 frost hollow 179 sw 0.00 0.00 0.00 0.00 Wombat Flat 1217 sw 3.40 30.14 16.77 13.78 X-Sandys Flat 145 sw 1.65 4.91 3.28 22.70

61 Table 3.5. Details of survey transects used to derive a density of calling male at montane elevation in the south-west (sw) (960 – 1299 m) and north-east (ne) (1200 – 1299 m) regions.

Census statistics for each year are seasonally corrected.

Survey Transect Total Stream Region 1996 1997 1998 1999 Mean No. of Mean No. of Length (m) Calling Males Calling Males/km Bell Clear Ck 1530 ne 0.0 0.0 0.00 Bell Clear Ck A 2395 ne 0.0 0.0 0.00 Cascade Ck B 559 ne 0.0 0.0 0.00 Cascade Ck below 1300 a 1537 ne 10.0 1.7 0.0 3.9 2.54 Cascade Ck below 1300 b 1216 ne 0.0 0.0 0.00 Cascade Ck frost hollow side branch 216 ne 0.0 0.0 0.00 Cascade Ck side branch A 208 ne 0.0 0.0 0.00 Little Boys Ck upper 1513 ne 0.0 0.0 0.0 0.0 0.00 Little Girls Ck 677 ne 0.0 0.0 0.0 0.0 0.00 Little Girls Ck a 1128 ne 0.0 0.0 0.0 0.0 0.00 Little Girls Ck b 1640 ne 0.0 0.0 0.00 North Cascade 1 1149 ne 0.0 0.0 0.00 North Cascade 2 1151 ne 0.0 0.0 0.00 Rum Ck 883 ne 0.0 0.0 0.00 Rum Ck Ezards mid branch 1069 ne 0.0 0.0 0.0 0.0 0.00 South Cascade 1013 ne 0.0 0.0 0.0 0.00 Swift Ck above Rd 1371 ne 0.0 0.0 0.0 0.00 Talbot Ck below 1200 603 ne 0.0 0.0 0.00 Whitelaw Ck 2448 ne 0.0 0.0 0.0 0.00 Whitelaw Ck (middle branch) 2751 ne 0.0 0.0 0.0 0.00 Whitelaw Ck (right branch B) 1699 ne 0.0 0.0 0.0 0.0 0.00 Whitelaw Ck (right branch C) 1459 ne 0.0 0.0 0.00 Whitelaw Ck 1 1596 ne 0.0 0.0 0.0 0.00

62 Survey Transect Total Stream Region 1996 1997 1998 1999 Mean No. of Mean No. of Length (m) Calling Males Calling Males/km Whitelaw Ck 2 794 ne 0.0 0.0 0.00 Whitelaw Ck right branch A 2094 ne 0.0 0.0 0.0 0.00 Barnies Ck 1634 sw 0.0 0.0 13.3 7.3 5.2 3.18 Block 10 a 886 sw 5.8 5.8 6.55 Block 10 b 888 sw 15.4 15.4 17.34 Block 10 c 1799 sw 7.7 7.7 4.28 Block 10 d 621 sw 1.9 1.9 3.06 Block 10 e 4458 sw 5.8 5.8 1.30 Block 10 f 3282 sw 0.0 0.0 0.00 Block 10 Myrrhee Ck 5468 sw 0.0 0.0 0.00 Block 10 section a 174 sw 5.8 5.8 33.40 Block 10 section b 531 sw 3.8 3.8 7.15 Block 10 section c 1022 sw 3.8 3.8 3.72 Block 10 section d 1197 sw 1.9 1.9 1.59 Chairlift + Corner 385 sw 0.0 0.0 0.00 Charity Ck (below sewage pond) 459 sw 0.0 0.0 0.0 0.0 0.00 Charity Ck (bog below sewage farm) 212 sw 3.0 6.6 4.8 22.69 Charity Ck (left branch Ck A) 566 sw 0.0 0.0 0.0 0.00 Charity Ck (left branch) 1284 sw 4.0 8.7 8.4 5.0 6.5 5.06 Charity Ck (right branch above rd) 110 sw 0.0 0.0 0.0 0.0 0.00 Charity Ck (right branch below rd) 816 sw 7.4 0.0 10.8 6.1 7.47 Charity Ck (side Ck A) 395 sw 24.2 24.2 61.27 Charity Ck (side Ck B) 287 sw 0.0 3.4 1.7 5.92 Charity Ck (side Ck C) 220 sw 0.0 6.7 3.3 15.02 Charity Ck (side Ck D) 421 sw 0.0 0.0 0.0 0.0 0.00 Charity Ck (side Ck E) 210 sw 0.0 6.7 3.3 15.75 Charity Ck (side Ck F) 512 sw 0.0 0.0 0.0 0.0 0.00 Charity Ck (side Ck G) 244 sw 3.3 3.3 13.50 Charity Ck (upstream sewage farm) 1499 sw 12.5 0.0 14.9 9.1 6.07 Charity Ck left branch top 363 sw 0.0 0.0 0.0 0.00

63 Survey Transect Total Stream Region 1996 1997 1998 1999 Mean No. of Mean No. of Length (m) Calling Males Calling Males/km Charity Ck right br below Ck Corner 818 sw 0.0 0.0 1.7 0.6 0.73 Chinamans 1 1102 sw 16.2 16.2 14.70 Chinamans 2 4386 sw 43.7 43.7 9.96 Chinamans 3 1813 sw 20.0 20.0 11.03 Chinamans 4 1141 sw 2.9 2.9 2.54 Chinamans 5 231 sw 2.9 2.9 12.54 Chinamans 6 3145 sw 13.0 13.0 4.13 Ellery Ck 538 sw 24.5 24.5 45.53 Faith Ck (960-1000) 740 sw 0.0 0.0 0.00 Faith Ck 685 sw 0.0 7.4 3.7 5.40 Faith Ck left branch 788 sw 2.0 7.4 4.7 5.96 Faith Ck right branch 1121 sw 3.8 3.8 3.39 Hope Ck (access rd 3 above rd) 354 sw 8.3 22.5 13.2 14.7 41.51 Hope Ck (access rd 3 right branch) 1489 sw 3.0 1.7 2.3 1.54 Hope Ck (access rd 1 above & below rd) 549 sw 4.2 5.9 6.6 5.6 10.19 Hope Ck (access rd 2 above rd) 347 sw 0.0 0.0 1.7 0.6 1.73 Hope Ck left branch 2523 sw 0.0 5.9 3.3 14.6 5.9 2.34 Hope Ck 1113 sw 0.0 0.0 1.7 4.9 1.6 1.44 Hope Creek (Access Rd 2 below rd) 864 sw 25.0 7.8 14.9 15.9 18.41 Hope Creek (Access Rd 3 below rd) 1098 sw 7.4 3.9 13.2 8.2 7.47 Long Ck 2234 sw 46.6 80.3 31.4 17.6 44.0 19.70 Long Ck below Rd 1989 sw 0.0 0.0 0.00 Long Ck side branch 659 sw 31.8 46.3 33.3 37.1 56.29 Mt Tyers 1 574 sw 6.7 6.7 11.67 Mt Tyers 2 690 sw 11.1 11.1 16.08 Mt Tyers 3 842 sw 0.0 0.0 0.00 Mt Tyers 4 444 sw 0.0 0.0 0.00 Newlands 1 2081 sw 57.7 57.7 27.72 Newlands 2 2184 sw 115.4 115.4 52.84 Newlands 3 650 sw 0.0 0.0 0.00

64 Survey Transect Total Stream Region 1996 1997 1998 1999 Mean No. of Mean No. of Length (m) Calling Males Calling Males/km Newlands 4 1423 sw 31.5 31.5 22.14 Newlands 5 686 sw 73.1 73.1 106.53 Tanjil River East Branch 1542 sw 0.0 0.0 0.0 0.00 Tanjil River tributary 1 2050 sw 18.2 32.0 25.1 12.25 Tanjil River tributary 2 1674 sw 6.1 7.3 7.3 45.5 16.5 9.86 Thomson River 1 4029 sw 30.6 8.6 114.8 51.3 12.73 Thomson River 2 4312 sw 88.9 38.6 46.3 57.9 13.43 Thomson River 3 3443 sw 17.4 17.4 5.05 Thomson River 4 4785 sw 0.0 0.0 0.00 Thomson River 5 3648 sw 0.0 0.0 0.00 Tanjil River West Branch 1a above aqueduct 1144 sw 0.0 0.0 0.00 Tanjil River West Branch 1a below aqueduct 3027 sw 3.2 14.9 9.0 2.97 Tanjil River West Branch 1b above aqueduct 755 sw 10.1 10.1 13.37 Tanjil River West Branch 1b below aqueduct 2856 sw 22.2 20.1 15.2 19.2 6.72 Tanjil River West Branch 2 above aqueduct 2276 sw 40.0 50.0 50.4 81.8 55.5 24.39 Tanjil River West Branch 2 below aqueduct 3441 sw 5.6 5.6 1.63 Tanjil River West Branch 3 above aqueduct 923 sw 0.0 5.6 18.5 72.7 24.2 26.21 Tanjil River West Branch 4 above aqueduct 972 sw 0.0 6.7 3.0 3.2 3.29 Tanjil River West Branch 5 above aqueduct 678 sw 0.0 0.0 0.0 0.00 Tanjil River West Branch 6 above aqueduct 2058 sw 10.0 7.0 8.5 4.13 Tanjil River West Branch 7 below aqueduct 3387 sw 34.9 52.6 142.4 76.6 22.61 Tanjil River West Branch 7b below aqueduct 540 sw 3.2 3.2 5.93 Tanjil River West Branch 8 above aqueduct 990 sw 6.3 6.3 6.37 Tanjil River West Branch 9 above aqueduct 520 sw 12.3 12.3 23.64 Tyers River East Branch tributary 553 sw 0.0 0.0 8.3 0.0 2.1 3.80 Tyers River East Branch tributary 1 1841 sw 23.5 10.9 17.2 9.34 Tyers River East Branch tributary 2 1902 sw 23.5 2.2 12.8 6.73 Tyers River East Branch tributary 3 2666 sw 5.9 8.7 7.3 2.74 Tyers River West Branch tributary a 3589 sw 21.6 21.6 6.02 Tyers River West Branch tributary b 849 sw 62.8 62.8 73.97

65 Survey Transect Total Stream Region 1996 1997 1998 1999 Mean No. of Mean No. of Length (m) Calling Males Calling Males/km Tyers River West Branch tributary c 6822 sw 26.3 26.3 3.86 Tyers River 1 right branch 1985 sw 0.0 0.0 8.3 2.2 2.6 1.31 Tyers River 2 left branch 955 sw 0.0 0.0 0.00 Tyers River 2 right branch 784 sw 0.0 0.0 0.00 Tyers River 3 3804 sw 116.4 116.4 30.60 Tyers River 4 2632 sw 9.8 9.8 3.72 Tyers River 5 2719 sw 31.2 31.2 11.48

66 Table 3.6. Total stream length, density of calling males (/m of stream ± SE) and estimated population size of adult males within six geographic areas encompassing the distribution of Philoria frosti (see methods for description of areas).

N.B., for elevations > 1300 m, only lengths of stream from within known breeding habitats were used to estimate population size of adult males. The distribution of an individual survey transect may occur in more than one geographic area.

Geographic Area Total Stream > 1300 m > 1300 m Mean density of Estimated No. of Census Data Length (m) Non-breeding Breeding Calling Males (/km Adult Male Survey Used Habitat Habitat of stream + SE) Population Transects Stream Length Stream Length Size (m) (m) Sub-alpine (> 1400 m) north-east 37417.5 10259 27159 12.31 ± 10.97 334 12 1993, 1995, 1997 - 1999 Sub-alpine (> 1400 m) south-west 42602.4 9642 32961 9.94 ± 5.65 328 22 1998, 1999 Sub-alpine-montane (1300 – 1400 m) north-east 50076.5 15020 35057 7.44 ± 6.48 261 17 1998, 1999 Sub-alpine-montane (1300 – 1400 m) south-west 88974.8 35280 53695 26.61 ± 13.93 1429 27 1998, 1999 Montane (1200 – 1299 m) north-east 53584.8 - - 0.10 ± 0.10 5 25 1996 - 1999 Montane (960 – 1299 m) south-west 374876.0 - - 11.66 ± 1.80 4371 94 1996 - 1999 Total = 647532.0 Total = 6728 Total = 197

67

159 # Map Code Survey Transect 158 1 Access Road 1 # N 2 Access Road 2 3 Access Road 3 4 Baragwanath Flat 160 23 Cascade Creek frost hollow # 155 26 Chairlift + corner # 27 Chairlift corner 157 # 50 Creek Corner $ 51 Creek Corner 2 Mt W hitelaw 52 Currawong/Sandys/Pauciflora Flat 179 53 Currawong Flat

# # 54 Currawong Flat ds 162 170 55 East Tanjil # # 61 Freemans Flat/Tullicoutty 161 163 # 62 Gwinear Flat # 166 71 Jeep Track 1 # 72 Jeep Track 2 Baw Baw National Park # 164 73 Jeep Track 3 # 74 Jeep Track 4 # 167 168 75 La Trobe Plain 165 76 Little Boys State Forest/Mt Baw Baw Alpine Resort # 77 Little Boys A 77 # 156 # 93 86 Macallister Plain 182 76 # 87 McMillians Flat # 52 88 Moondarra Flat Area not surveyed between 1993 and 1995 93 Mustering Flat # 94 Neulyne Plain # # 131 102 Pudding Basin 62 # 71 # 61 105 South Cascade 132 Mt St Phillack $ # 106 South Cascade A Boundary enclosing individual survey transects 55 87# 107 South Cascade B # # 111 # # 53 72 110 Talbot Creek frost hollow 50 111 Tanjil Plain # # # 74 # # # 131 The Morass # $ Mt Baw Baw 54 51 154 4 # 73 132 The Morass (2 sections excluded) Potential breeding habitat # 1 # 153 23 139 Tyers River 1 frost hollow # # 142 Tyers River 2 frost hollow # 2 # 181 153 Village Flat/Dam Valley/Big Hill # 102 # # 142 154 Village Flat 2 # 3 94 # 26 # # 155 Whitelaw 1 88 180 # 10123Kilometers27 75 86 156 Whitelaw 10 Kilometres # 107 # 157 Whitelaw 11 158 Whitelaw 12 105 159 Whitelaw 13 139 # 160 Whitelaw 2 161 Whitelaw 3 162 Whitelaw 4 # 106 163 Whitelaw 5 164 Whitelaw 6 165 Whitelaw 6a 166 Whitelaw 7 (Potential breeding habitat mapping modified from aerial photography mapping by Roberts 1996) 167 Whitelaw 8 # 110 168 Whitelaw 9 170 Whitelaw Creek (frost hollow) 179 Whitelaw Ruins $ 180 Wombat Flat Fig. 3.1. Distribution of calling male survey transects censused between 1993 and 2000 on the Baw Baw Mt Erica 181 Wombat Flat A Plateau. Survey transects are identified by name and number in adjacent table. 182 X-Sandys Flat

Fig. 3.1 Distribution of areas (enclosed by blue lines) used to census calling males between 1993 and 2002 on the Baw Baw Plateau.

Transects are identified by name and number in adjacent table.

69

137 136 15 178 Map Code Survey Transect Map Code Survey Transect

# N 5 Barnies Creek # 101 North Cascade 2 6 Bell Clear Creek 103 Rum Creek # 174 7 Bell Clear Creek (below rd) 104 Rum Creek (Ezards mid branch)

# 8 Bell Clear Creek A 108 Swift Creek (above rd) 173 177 9 Block 10a 109 Talbot Ck (below 1200 m) 10 Block 10b 112 Tanjil River East Branch # 11 Block 10c 113 Tanjil River Tributary 1 175 # 99 # 12 Block 10d 114 Tanjil River Tributary 2 13 Block 10 Myrrhee CK 115 Tanjil River West Branch 1A (above aqueduct) 133 171 14 Block 10e 116 Tanjil River West Branch 1A (below aqueduct) 14 15 Block 10f 117 Tanjil River West Branch 1B (above aqueduct) # 129 # # 16 Block 10 section a 118 Tanjil River West Branch 1B (below aqueduct) # 169 17 Block 10 section b 119 Tanjil River West Branch 2 (above aqueduct) # 115 18 Block 10 section c 120 Tanjil River West Branch 2 (below aqueduct) 95 # 134 19 Block 10 section d 121 Tanjil River West Branch 3 (above aqueduct) # 7 20 Cascade Creek (below 1300 m A) 122 Tanjil River West Branch 4 (above aqueduct) # # 21 Cascade Creek (below 1300 m B) # 130 123 Tanjil River West Branch 5 (above aqueduct) 22 Cascade Creek B # 16 96 124 Tanjil River West Branch 6 (above aqueduct) # # # 24 Cascade Creek frost hollow-side branch 125 Tanjil River West Branch 7 (below aqueduct) # 13 # 135 12 # # 17 25 Cascade Creek (side branch A) 126 Tanjil River West Branch 7b (below aqueduct) # # 98 # # 81 28 Charity Creek (left branch) 127 Tanjil River West Branch 8 (above aqueduct) # 97 117 8 29 Charity Creek (left branch Ck A) 128 Tanjil River West Branch 9 (above aqueduct) # 30 Charity Creek (left branch-top) # # 129 Tanjil River West Branch 10 (above aqueduct) 9 116 # 175 # 31 Charity Creek (right branch-above rd) # # # 130 Tanjil River West Branch 11 (above aqueduct) # 78 19 # 119 # 32 Charity Creek (right branch-below rd) # 133 Thomson River 1 10 # 176 6 121 33 Charity Creek (bog below sewage farm) 134 Thomson River 2 11 # # 34 Charity Creek (below sewage pond rd) 126 # # 135 Thomson River 3 # # 122 172 35 Charity Creek (side ck A) 136 Thomson River 4 36 Charity Creek (side ck B) # # 137 Thomson River 5 # 18 123 Mt Whitelaw # 37 Charity Creek (side ck C) 138 Tyers River 1 (right branch) 125 # 38 Charity Creek (side ck D) 140 Tyers River 2 (left branch) 128 39 Charity Creek (side ck E) 141 Tyers River 2 (right branch) # 40 Charity Creek (side ck F) 143 Tyers River 3 120 82 80 41 Charity Creek (side ck G) 144 Tyers River 4 118 124 # 42 Charity Creek (up-stream sewage pond rd) 145 Tyers River 5 43 Charity Creek (right branch-below Ck Corner) 146 Tyers River East Branch (tributary 1 ) Baw Baw National Park 127 44 Chinamans 1 147 Tyers River East Branch (tributary 2 ) 79 45 Chinamans 2 44 # 148 Tyers River East Branch (tributary 3 ) 46 Chinamans 3 149 Tyers River East Branch (tributary) 47 47 Chinamans 4 150 Tyers River West Branch (tributary A)

# 108 48 Chinamans 5 151 Tyers River West Branch (tributary B) # 49 Chinamans 6 152 Tyers River West Branch (tributary C) State Forest/Mt Baw Baw Alpine Resort 56 Ellery Creek 45 # 49 169 Whitelaw Creek # # 57 Faith Creek 171 Whitelaw Creek (lower section) 58 Faith Creek (800-1000 m) # 172 Whitelaw Creek (right branch A) 100 59 Faith Creek (left branch) 173 Whitelaw Creek (right branch A-lower section) 46 60 Faith Creek (right branch) 174 Whitelaw Creek (right branch B) Boundary enclosing location of survey transects 48 112 63 Hope Creek 175 Whitelaw Creek (right branch C) when not independent of other transects 64 Hope Creek (access rd 1 above & below rd) 176 Whitelaw Creek 1 # 65 Hope Creek (access rd 2 above rd) 177 Whitelaw Creek 2 38 # 30 40 Mt St Phillack 66 Hope Creek (access rd 2 below rd) 178 Whitelaw Creek Middle Branch 114 67 Hope Creek (access rd 3-above rd) 39 # 101 68 Hope Creek (access rd 3-below rd) Survey transect location # # 43 # 69 Hope Creek (access rd 3-right branch) # 41 31 # 65 70 Hope Creek (left branch) 37 78 Little Boys Creek (lower) # Mt St Gwinear # # 79 Little Boys Creek (upper) # 22 # Mt Baw Baw # # 80 Little Girls Creek Stream/creek/seepage/soak 113 # # # 81 Little Girls Creek A # 64 # 82 Little Girls Creek B # 67 # 83 Long Creek # 24 20 29 # 35 69 # # # 84 Long Creek (below road) # # 70 85 Long Creek (left branch side Ck) # 36 # 89 Mt Tyers 1 28 # 21 90 Mt Tyers 2 33 # 63 59 25 91 Mt Tyers 3 32 92 Mt Tyers 4 42 83 95 Newlands 1 34 # 68 # 57 96 Newlands 2 # 97 Newlands 3 66 98 Newlands 4 56 109 99 Newlands 5 # 60 5 100 North Cascade 1 # 90 # 89

#

# # # 91 # 303Kilometers # # 58 # 104 92 # # 84 152 Mt Erica 85 # # 151 103 # 150 146 145 140 148

# # Fig. 3.2. Distribution of drainage lines (highlighted in blue) used to census calling males between 1996 and 2002 on the south-western # 147

# and north-eastern escarpments of the Baw Baw Plateau. Transects are identified by name and number in adjacent table. # # # #

# # 149 144 143 141 138

Fig. 3.2 Distribution of drainage lines (highlighted in blue) used to census calling males between 1996 and 2002 on the south-western and north-eastern escarpments of the Baw Baw Plateau.

Transects are identified by name and number in adjacent table.

71

N

Mt W hitel aw

Stream/creek/seepage/soak

Montane (1200 - 1299 m) north-east

Mt St Phillack Montane (960 - 1299 m) south-west

Mt St G wi near Sub-alpine-montane (1300 - 1400 m) north-east Mt Baw Baw

Sub-alpine-montane (1300 - 1400 m) south-west

Sub-alpine (> 1400 m) north-east

Sub-alpine (> 1400 m) south-west Mt Eric a

Montane Fen (excluded from analysis) 2 0 2 KilometersKilometres

Fig. 3.3. Geographic areas from which calling male density and population estimates were derived. N.B. The area identified as Montane Fen was excluded from analysis. Other non-breeding habitat areas > 1300 m elevation were also excluded from analysis, but are not illustrated here (see methods).

Fig. 3.3. Geographic areas from which calling-male density and population estimates were derived.

N.B. The area identified as Montane Fen was excluded from analysis. Other non-breeding habitat areas located above 1300 m elevation were also excluded from analysis, but are not illustrated here (see 3.2).

73

# # # # # # # 1 # # #### 1 # ### ## # # # 0 ### # # 0 ## # ## 0 ## # 0 ## # # # # # ## 2 1 #### # # ## ## ## # # # # ### ###### # # # # ## # # ## # # ## # ## # # # # # #### # # ## # 00 #### # ## 12 # ## # ### # ## ## ## ##### # ##### # # # # #### ## ## ##### # # #### ### ##### # # ## ## ## # ## ######## # ## # ##### ####### # # # # # #### # # Mt Whitelaw ## ########### # #â â 1 ## # # #ââ# 1 0 # # # # ââ# â#â ### ## # # #â # 0 0 # ## ## # # ## â# â# â 0 #### # â â 0 ### ## ## ââ â # # # â 1 # 0 # # #â â â â 0 # #â â â ## ââ 0 â â â â â â 1 # â 0 â â â 4 â âââ â â ââ 1 â #â â â 1 â# ââ ââ 10 â â â â â 0 â â# â# â ââ 0 â ââââ â â â ââ ââ â ## â â # ââ 9 9 # â â â â ## â â â ââ 0 0 # # â â â 0 0 #### â â â â # â # â # â â ## â ââ â 13 ââ âââ â â ## 00 â â# â â â #â â â # â â #â â â ââ â â â â â ### # â # â ââ ###â## â â âââ â # #â#â #âââ# â ââ â â ###ââ ââ ââ 14 # â## ### â â â â â# â# #â 0 â#â â# #â#â## â â 0 # â ââ##### â â â â# ##â## ## â â â ââ## â â â #### â# â â# # â Baw Baw National Park â â# â â# ââ â â #â#â# # â ## #ââ â â âââ â # â # # â ââ â â# â #â##â â ## â â â# 1500 # #â# # ## â ââ# â â # #â## â ## ## âââ â â â ââ # #â# â##### # ## # # # ââ# â â ââ ââ # ###â#â# â#ââ #â# #### #â â ââ â â # #â ââ# #â ââ#â â â ââ ## â####â â 1â â â# â 1 ####### ### ## â # ââ â5# ââ ââ â â 500 State Forest/Mt Baw Baw Alpine # # # #â âââ 0ââ## 1 ââ â â # ## ## â## ââ â#â â â 0# 5 â# â ### ### â # ââ 0 â â â ## ## # #####â## â#â# ââ####â 0 â Resort ### #### â# #â## â # # â â âââ ââââ # ###ââ # # #â â## â # # ### ## ââ ââââ## # Mt St Gwinear # # # â ââââ â â # # ## ## # â â ââ â â # ## ## ## # â ââ â â # #â# â ### # # ##â## # ââ #ââ ## â ### #â# ##### â â ââ â # Contour Height (metres a.s.l.) #### â â â ââ # # # # # # ââ ââ# # ## ## â# # # ââ â ## #â# ### ## â#â ââ ## # # # # â#â# â ââ Mt Baw Baw ## ### ## ââ# â â# â â â â# # # ### â ââ#â##â # â# # ## # ââ# ââ # â# # # ââ# # ## â â 0 # # # â## â âââ 0 ## # ââ ââ â 14 # # â ââ Locations of calling males recorded in 1983 # # # â â â â # â â ## â # # â â and 1984 surveys (Malone 1985a) â â # ## â# # # # 1000 # # Locations of calling males recorded in 1993 - 2002 1 # 2 # # # surveys # ##### # 0 ## ##### 0 ###### ##### # # # ### # # 1 # ## # # # # # # # 3 # # ##### 0 # #

# 0 0

0

5 ## # # ### # 1 # ## 1400 20246KilometersKilometres 11 ## Mt Erica 0 # # 0 # # # # 11 # # 0#0 1000 # ## ## # # # ## 0 # ## # # 0 ## # ## # 1 # # ##### # 1 # # # # # 1200 ## # # ## # ## # # # # #### # # # ## # #

Fig. 3.4. Distribution of calling males recorded during surveys in 1983 and 1984 (Malone 1985a) and 1993 – 2002, Baw Baw Plateau and escarpment.

75

100

80

60

40

20

Mean proportion of calling males (%) of calling males Mean proportion 0 20 30 9 19 29 9 19 29 OctOct Nov Nov Nov Dec Dec Dec Date (day/month)

Fig. 3.5. Mean proportion of calling males derived from fortnightly, repeated surveys conducted at four survey transects over the duration of the 1993 breeding season, as depicted by distance-weighted least squares smoothing (see methods).

77 7 survey transects (1983, 1984, 1993 - 2002) 800

600

400

200 Number of calling males

0 1980 1984 1988 1992 1996 2000 2004 Year

14 survey transects (1983/84 composite, 1993 - 2002) 2000

1500

1000

500 Number of callingmales

0 1980 1984 1988 1992 1996 2000 2004 Year

14 survey transects (1993 - 2002)

60 Data Type Seasonally Corrected 45 Raw

30

15 Number of calling males

0 1992 1994 1996 1998 2000 2002 2004 Year

Fig. 3.6. Relationship between the number of calling males and survey year for subsets of census data recorded between 1983 and 2002 at sub-alpine (> 1400 m) elevation.

78 3 survey transects (1983, 1984, 1993 - 2002) 80 70 60 50 40 30 20 Number of calling males 10 0 1980 1984 1988 1992 1996 2000 2004 Year

5 survey transects (1993 - 2002) 200 Data Type Seasonally Corrected 150 Raw

100

50 Number of calling males

0 1992 1994 1996 1998 2000 2002 2004 Year

Fig. 3.7. Relationship between the number of calling males and survey year for subsets of census data recorded between 1983 and 2002 at sub-alpine-montane (1300 – 1400 m) elevation.

79 4 survey transects (1983, 1984, 1993 - 2002) 100

80

60

40

Number of calling males 20

0 1980 1984 1988 1992 1996 2000 2004 Year

4 survey transects (1993 - 2002) 20 Data Type 16 Seasonally Corrected Raw

12

8

Number of calling males 4

0 1992 1994 1996 1998 2000 2002 2004 Year

13 survey transects (1997 - 2002)

250 Data Type 200 Seasonally Corrected Raw

150

100

Number of calling males 50

0 1996 1998 2000 2002 2004 Year Fig. 3.8. Relationship between the number of calling males and survey year for subsets of census data recorded between 1983 and 2002 at montane (960 – 1299 m) elevation.

80 s

e

l

a 100

m

g

n

i

l

l

a

c

f

o

r

e b 10

m

u

n

d

e

t

c

e

r

r

o

C 1 1 10 100 Number of calling males

Fig. 3.9. Relationship between raw census data of calling males and the same data seasonally- corrected for variation in participation of calling males over the duration of 1993 – 1999 breeding seasons. Note logged x and y scales.

81

Chapter 4

BREEDING HABITAT FLORISTICS AND STRUCTURE: IDENTIFICATION OF PREFERRED ATTRIBUTES AND ASSOCIATED ENVIRONMENTAL GRADIENTS

4.1 Introduction

Identifying the location of plants and animals, and the factors affecting their patterns of distribution and abundance plays a key role in managing and conserving biodiversity (Burgman and Lindenmayer 1998; Scribner et al. 2001). For the evolving fields of landscape ecology and conservation biology, the study of habitat is also considered essential (Lindenmayer and Franklin 2002). Knowledge of what constitutes habitat for a species, particularly those that are threatened and vulnerable, is critical given that habitat loss is the largest factor contributing to the current global extinction of species (Fahrig 2001). For amphibians, habitat destruction and modification is also considered to be the biggest cause of population declines and losses (Blaustein et al. 1994; Alford and Richards 1999). An increasing number studies have examined habitat components believed to play a role in the ecology or conservation of amphibians (e.g., Cooke and Frazer 1976; Strijbosch 1980; de Fonseca and Jocque 1982; Corn and Fogleman 1984; Laan and Verboom 1990; Welsh and Lind 1995, 1996; Marnell 1998; Babik and Rafiński 2001; Bull and Marx 2002; Hamer et al. 2002; Nyström et al. 2002; Beard et al. 2003; Lecis and Norris 2003). However, the global phenomenon of amphibian declines has highlighted the paucity of ecological and natural history information on most species (Sartorius and Rosen 2000), and the requirement to gain a comprehensive understanding of habitat and adaptations (Monello and Wright 1999).

To reverse amphibian declines, it is considered that management of populations will require definition of high-quality habitat for individual species or groups of species, followed by efforts to retain or restore these habitats on the landscape (e.g., Knutson et al. 1999). An understanding of factors that regulate populations and limit distributions is also required if conservation measures are to be successful (Nyström et al. 2002). Obtaining such information may be additionally relevant to species occupying high-elevation, mountain-top environments, given the prevalence of declines in these areas (e.g., Osborne 1990a; La Marca and Reinthaler 1991; Crump et al. 1992; Richards et al. 1993). Knowledge of the natural adaptations of mountain-top amphibian populations may also be important in developing adequate measures for the prevention of future declines in these particular environments (Vences et al. 2000). For anurans, breeding habitats associated with oviposition and development of embryos and larvae are particularly important, given that it is in these locations where embryonic and larval development occur, and that these

83 developmental periods are considered to be the most vulnerable phases of the anuran life cycle (Duellman and Trueb 1994).

Due to its cryptic nature, records of P. frosti to date are almost exclusively from the period encompassing the breeding season, when calling males can be located. Subsequently, most habitat descriptions relate specifically to breeding habitat (e.g., Littlejohn 1963; Malone 1985a, b). These descriptions infer a preference for breeding habitat that is characterised by intermittent water courses or rivulets on the slopes of shallow valleys on the Baw Baw Plateau and adjacent slopes. Most breeding sites have been recorded in unforested areas on the Baw Baw Plateau in wet alpine health-bog ecotones (Malone 1985a), and others from small creeks on the plateau slopes (Littlejohn 1963). The micro-habitat of breeding sites has been described as natural cavities that act as catchments for water travelling down slopes, where they have been recorded in association with a wide range of features, including vegetation, logs, rocks and man-made materials (Littlejohn 1963; Malone 1985a, b). Although the breeding habitat of the species has been described, no studies have examined it in detail, and no observations have been made since the population decline of the species from sub-alpine elevations (see Chapter 3).

The sensitivity of P. frosti to breeding habitat disturbance or modification is not well known. A study carried out on embryonic and larval survivorship suggested that vegetation structure was very important to survivorship of the species, with lower terrestrial recruitment, through high levels of embryonic and larval mortality, being recorded in disturbed habitats compared with undisturbed (Malone 1985a). Desiccation through periodic drying of oviposition sites was acknowledged as the factor that resulted in this observation. However, results from a second season of study produced equivocal results, in which there was a significant increase in embryonic mortality at disturbed sites, but no difference in levels of larval mortality, and in the proportion of individuals completing metamorphosis, at disturbed and undisturbed sites (Malone 1985a). These observations suggest that habitat disturbance may be detrimental to recruitment, but that other factors such as prevailing climate and associated habitat structure, may result in development that varies temporally.

Although some of the fundamental characteristics of the breeding habitat of P. frosti have been described previously, detailed knowledge of preferred breeding habitat attributes associated with the extant population remains unknown. Given the magnitude of its recent decline from higher elevation (Hollis 1995; see Chapter 3), and that particular elements of vegetation structure have been suggested previously as being important for terrestrial recruitment, detailed information on habitat is considered essential for future management and conservation of the species. The aims of this study were to: (1) describe and quantify breeding habitat from the historic and extant distribution of P. frosti, (2) identify preferred breeding macro and micro-habitat attributes and associated environmental gradients of the extant population, and (3) compare current use of breeding habitat with that of breeding habitat used historically.

84 It was not the purpose of this study to demonstrate cause and effect relating to the recorded decline of P. frosti, but rather, to explore and identify habitat attributes and environmental gradients that best characterise the breeding habitat of the extant population. It was considered that this information would be useful for management and conservation purposes, as well as providing direction for the development and design of further research to answer more specific questions about the decline and ecology of P. frosti.

4.2 Materials and Methods

4.2.1 Sampling Design

To identify the key breeding habitat attributes of P. frosti, it was not possible to base an analysis on a comparison of breeding and non-breeding sites due to the difficulty in classifying habitat as ‘non-breeding’. The use of breeding habitat by P. frosti is likely to vary from season to season such that one cannot categorically label a patch of habitat as ‘non-breeding’ without observing it over many breeding seasons. Such an exercise would be further compromised given the recent population decline of the species from higher elevations (Chapter 3). The presence and absence of historical breeding records, including those recorded in population surveys in 1983 and 1984 by Malone (1985a) prior to the species decline, could therefore not be used as a means of establishing the locality of non-breeding sites from which environmental data could be collected. The design of this study was subsequently based on comparing the ecological attributes of a sample of extant breeding habitat of P. frosti with that of a sample of randomly-selected habitat from within the known breeding distribution of the species. This comparative analysis attempts to test the null hypothesis that attributes of currently-preferred breeding habitat are located randomly across the historic and extant breeding habitat distribution of P. frosti.

The collection of breeding and random site information was conducted at three scales to facilitate the identification of potential variables that might best discriminate between breeding and random sites: (1) a micro-habitat scale - 1 x 1 m quadrat, for the purpose of quantifying accurately micro- habitat characteristics of calling (oviposition) sites, and (2) a macro-habitat scale - 10 x 10 m quadrat, for the purpose of quantifying breeding habitat immediately surrounding breeding sites, and (3) variables collected at a 'local' or sub-catchment scale, for the purpose of quantifying sub- catchment characteristics in the vicinity of breeding sites. The placement of 1 x 1 m quadrats centrally within the boundary of the 10 x 10 m quadrat resulted in habitat attributes of one being a subset of the other, although data collected from the lower profile height classes of 1 x 1 quadrats was conducted in a different manner (see habitat attributes and measurement below).

85 4.2.2 Selection of Breeding and Random Sites

The sample of breeding sites selected for examination was chosen from positions of calling males located accurately during annual population surveys. Observations made during this study, and previous studies (B. Malone pers. comm), indicate that oviposition sites are located at, or in close proximity to, calling sites. At the commencement of this study, breeding sites were initially chosen from calling sites located during the 1995 population survey (see Chapter 3). This survey was restricted predominantly to sub-alpine elevations (above approximately 1300 m) due to the then current knowledge of the species distribution (following Malone 1985a). During the population survey in 1996, breeding sites of P. frosti were discovered in montane habitats as low as 960 m on the south-western escarpment of the Baw Baw Plateau (Chapter 3). It was therefore necessary to sample additional sites from breeding habitat at montane elevations. These were selected from male calling positions located during 1997 and 1998 population surveys. A total of 73 breeding sites distributed across the known distribution of P. frosti was selected for examination (Fig. 4.1). Each was marked during population surveys (Chapter 3) with flagging tape, and its position accurately mapped.

The selection of random sites for sampling was initially based on a random, stratified design whereby sites were chosen from a 100 x 100 m grid overlayed on a 1:25000 map of the Baw Baw Plateau (Balkau 1987). Existing knowledge about the breeding habitat requirements and distribution of the species (Malone 1985a) was used to stratify selection of random sites. Only sites containing breeding habitat within the range of habitat characteristics previously described were eligible for selection. This stratification confined the selection of random sites to predominantly unforested habitats on the Baw Baw Plateau, consisting of wet alpine heath and bogs, bog with intersecting pools of water and open grassland habitats (modified vegetation within the Mt Baw Baw Alpine Resort). To ensure an equal spread of sites across the distribution of P. frosti, the species distribution identified by Malone (1985a) was stratified into three regions: (1) central western region, bounded by the Alpine Walking Track, Tanjil River West Branch and headwaters of the Tyers River West Branch and Ellery Creek catchments; (2) south-eastern region, bounded by the Alpine Walking Track and Whitelaw Creek; and (3) north-western region, bounded by the Alpine Walking Track, Tanjil River West Branch and Whitelaw Creek. Overlayed 100 x 100 m grid cells were numbered sequentially, and 37 random sites chosen from a set of generated random numbers. For each randomly-chosen grid cell, the location of each site was more accurately positioned by selecting the intersection of the north and eastern boundary.

The discovery of P. frosti in gully systems at lower, montane elevations (described above) resulted in the need for further random sites to be sampled from the newly recorded montane distribution of the species. Sites were selected with the same grid-cell approach described above, except that the distribution from which they were chosen was confined to gully systems with known populations of P. frosti, as of 1998. Due to the presence of a eucalypt canopy cover, and subsequent lack of

86 knowledge about understorey habitat characteristics in many gully systems, all gully systems from which P. frosti had been recorded at montane elevations were considered to have potential breeding habitat along their length, but not beyond the area comprising riparian vegetation. Additional 1:25000 mapping of vegetation and hydrology (Roberts 1996) was used to assist the selection of random sites from montane elevations. Including random sites chosen from sub-alpine elevations, a total of 51 random sites was sampled across the known distribution of P. frosti (Fig. 4.1).

4.2.3 Habitat Attributes and Measurement

Knowledge about the ecology, distribution and sensitivity of P. frosti (Malone 1985a, b), and consideration of other environmental factors notable for being important to general amphibian biology (see Duellman and Trueb 1994), were used select variables to best describe and quantify characteristics of habitat recorded at breeding and random sites. Ecological variables used to describe habitat included floristic, landform, hydrological, soil, and macro and micro-climatic factors.

4.2.3.1 Floristics

The floristic composition of 1 x 1 m and 10 x 10 m scale quadrats was recorded for different vegetation height classes following Peters (1983): 0 - 0.1, 0.1 - 0.3, 0.3 - 1, 1 - 2.5, 2.5 - 5, 5 - 8, 8 - 15, 15 - 27, 27 - 41 and 41 - 55 m. The cover abundance of each species, in each height class, was visually estimated using a simplified version of the Braun-Blanquet scale (Braun-Blanquet 1928). Total vegetation cover in each height class was visually estimated as the proportion of ground that would be covered by the vertical projection of foliage and branches (Walker and Tunstall 1981) within the following cover groups: 1 < 5% cover, 2 = 5 - 50% cover, 3 > 50% cover. Because not all flora could be identified to species level due to the absence of reproductive or diagnostic structures, taxa were amalgamated if they could not be identified accurately to species level (e.g., Hydrocotyle spp. = H. hirta or H. algida). In some cases it was possible to identify species at the sub-species level. This process derived a list of flora that was classified to either sub-species, species or generic levels, for analysis. Nomenclature used for vascular plants followed NRE (2001).

4.2.3.2 Structural Variables

A number of structural habitat attributes were also quantified using the same cover abundance scores as for flora: (a) exposed rock outcrop, (b) litter, (c) log, (d) bare ground, (e) ground/earth (for 1 x 1 m quadrats only), (f) standing water, (g) surface seepage, (h) running water, (i) Epacridaceae/Myrtaceae stems, (j) Leptospermum above ground roots/stumps, (k) road/bitumen,

87 and (l) other artificial structures. Because data from 1 x 1 m quadrats was collected to provide more detail, the collection of cover abundance data from the lower height classes was undertaken in a different manner to that of 10 x 10 m quadrats. The lowest substrate point in 1 x 1 m quadrats was taken to be the point of 'zero height' within the profile, whilst the general slope of the land surface in 10 x 10 m quadrats was considered to be the point of ‘zero height’ throughout the whole quadrat. Data collected in this manner from 1 x 1 m quadrats produced an enhanced 3-dimensional picture of the land surface and drainage characteristics. The variable ‘ground/earth’ was subsequently incorporated into profile descriptions to account for the estimated cover of apparent ground stratum within each height class.

Other biophysical variables quantified within 1 x 1 m and 10 x 10 m quadrats, and their method of measurement, are detailed in Table 4.1. These include a number of variables whose cover abundance was also estimated during the collection of floristic data.

4.2.4 Index of Breeding Habitat Importance

The number of calling males recorded at breeding sites within 10 x 10 m quadrats during the year of sampling was used as an index of breeding habitat preference. Because of the potential for variability in counts of calling males due to temporal influences (see Chapter 5), counts from each site were first standardised to minimise this variability (see Chapter 3 for methods). Unfortunately it was not possible to correct for seasonal variability at sites from which no calling males were recorded. The number of calling males recorded from random sites was not used as an index of breeding habitat importance because it was considered that the absence of breeding activity at a random site did not necessarily demonstrate that the habitat was less preferred for breeding purposes, given the limited number of censuses for calling males made at each random site during the study.

4.2.5 Soil Sampling Limitations

The collection of soil characteristics at each site was restricted by a number of factors. Field texture determination of soil was difficult due to the saturated nature of many profiles, as they are normally made on unsaturated samples (McDonald et al. 1990). The presence of ‘floating’ rock within the soil profile at some sites prevented collection of information about the characteristics of sub-soil horizons. Data from soil horizons two and three were therefore omitted from some analyses (see 4.2.6). It was also recognised that the depth of saturation of the soil greatly influenced by recent rainfall and run-off events. To minimise this potential bias, data collection was confined to periods of stable and consistent weather.

88 4.2.6 Data Analysis

Ordination was considered to be the most suitable procedure for summarising the large quantity of ecological data obtained, and for identifying environmental gradients that best discriminated breeding habitat attributes. Non-metric multidimensional scaling (NMDS) (Kruskal 1964) was used in preference to other multivariant analysis procedures involving ordination because of its robustness, effectiveness over linear techniques and suitability for gradient analysis (see Kenkel and Orloci 1986; Minchin 1987). The procedure constructs a configuration of sampling points in a pre-arranged number of spatial dimensions such that the distance between all site pairs are in rank order with their dissimilarities. A stress function is used to measure the level of agreement between the interpoint distances and dissimilarity values as adjustments are made to maximise their rank-order (Jongman et al. 1987). As agreement between interpoint distances and dissimilarity values improves, stress is decreased, with a perfect rank-order agreement deriving a stress of zero.

4.2.6.1 Ordination of Floristic and Structural Data

NMDS was performed on floristic and structural data collected from 1 x 1 m and 10 x 10 m scale quadrats at breeding and random sites. In total, 161 vascular and non vascular taxa, or groups of taxa, were quantified during analysis (Appendix 4.2). Integer cover score classes for taxa and structural attributes in each height class (pseudo-attributes) were first converted to percentage cover values using the midpoints of the three cover classes (1 = 0.025; 2 = 0.275; 3 = 0.750), multiplied by the total percentage cover in each height class. For example, if a pseudo-attribute in the 2 - 2.5 m height class had a cover score of 2, and an estimate for total vegetation cover of 60%, then the percentage cover for that pseudo-attribute was 0.275 x 60 = 16.5%. Percentage cover values were then standardised to unit maxima (SMAX) by dividing all cover values for all height pseudo-attributes, over all sites, by the maximum cover value attained by any of the height pseudo- attributes belonging to a taxon or structural attribute. All taxa and structural attributes that occurred in at least one quadrat were retained for analysis.

At 1 x 1 m and 10 x 10 m scales, the dissimilarity between all pairs of quadrats was determined using the Bray-Curtis dissimilarity coefficient (Bray and Curtis 1957). However, the algorithm used to derive each coefficient was arithmetically modified to allow for dissimilarities between pairs of quadrats to be derived from cover values present in adjacent quadrat height classes, with a weighting scheme used to determined the height and extent to which two height-class pairs should be considered different (see 4.2.6.2 below). This modified dissimilarity coefficient is referred to in this study as the ‘height-difference-weighted Bray-Curtis’ (h-d-w BC) (Minchin unpublished). It was used in preference to the standard Bray-Curtis coefficient (where total covers are examined) because it allowed for the interrelationships between pseudo-attributes in different height classes, from different quadrats, to be accounted for in the analysis, thereby providing greater capacity to

89 compare quadrat pairs in multidimensional space. The h-d-w BC dissimilarity (Djk,) between two sites j and k was computed as:

s  h h  2∑∑∑ wlm min()X ijl , X ikm  i===111 l m  D jk =1− ()T j +Tk

where s is the number of species, h is the number of height classes and Xijl is the abundance of species i, height class l in site j. Tj and Tk are 'totals' of sites j and k. The 'total' of a site j was defined as:

s  h h  Tj = ∑∑∑ wlmmin()X ijl , X ijm  i===111 l m 

The wlm are weights that depend on the difference between the two height classes (l and m) at which abundances are being compared. These can be generated using any function that takes the value 1.0, when l = m, and decreases with increasing difference between height classes. In this study, an additional analysis was undertaken to determine the most appropriate weighting scheme to be adopted (see 4.2.6.2 below).

Because the distribution of height classes used in the collection of floristic and structural data was not equal (i.e., being closer together as one nears the ground), it was not possible to apply a weighting function without either blurring the more detailed information present in the lower height classes, or analysing information in the upper height classes as if they were separate species. Each height class was subsequently analysed as an integer class, as follows: 0 - 0.1 = 1, 0.1 - 0.3 = 2, 0.3 - 1 = 3, 1 - 2.5 = 4, etc. NMDS was performed on 1 x 1 m and 10 x 10 m quadrat dissimilarity matrixes in one to six dimensions, using ten random starting configurations.

Analysis of similarities (ANOSIM) (Clarke 1993), was used to test the null hypothesis that the average rank of dissimilarities did not differ between breeding and random sites at 1 x 1 m and 10 x 10 m scales, using 10,000 random permutations. This analysis was further stratified for sites from sub-alpine elevations (> 1270 m) and montane elevations (≤ 1270 m) due to their vastly different physiography. For data sets where a significant difference was recorded, a step-wise variant of ANOSIM (Minchin unpublished), modified to work in association with the h-d-w BC, was also used to determine which taxa or structural attributes best discriminated between random and breeding sites. This method is based on the computation of the ANOSIM statistic - R, which is sensitive to the omission of species that are most beneficial to the groups investigated. A maximum value of R is eventually achieved for a smaller data subset that best discriminates between the groups in question, after which R declines as further species are removed. To work

90 with the h-d-w BC, the procedure was modified to omit all height pseudo-attributes for a given attribute, with only the lowest height-class pseudo-attribute being reported for retained attributes. It was considered that to retain individual height pseudo-attributes would alter the height profile pattern for that attribute. Each analysis was run with 1000 permutations.

4.2.6.2 Selection of a Weighting Scheme for Comparing Quadrat Height Class Pairs

The ANOSIM R statistic was also used to select a suitable weighting scheme for comparing height class pairs when deriving the h-d-w BC for ordinations. As it was the primary focus of this study to determine if there was a difference between breeding and random sites, a weighting scheme that maximised the separation between the two, as indicated by a maximum ANOSIM R statistic, was chosen to be the most suitable. This process was undertaken by examining the relationship between quadrat height classes and change in the ANOSIM R statistic using an exponential decay function with a 0.5 weight. The final weighting scheme chosen for each data set was therefore the height class spacing for which attributes within and between quadrat height classes would be considered half the same (referred to as ‘deltahalf’). For example, for a single height-class spacing, the exponential decay function would be 0.5 for a difference of one height class, 0.25 for two height classes, 0.125 for three height classes, 0.0625 for four height classes, etc. The weighting function used is given by:

b hl −hm wlm = e

where hl and hm are the heights represented by classes l and m and b is a negative user-defined constant. To produce weights that decline to 0.5 (the decay weight chosen in this study) at a height difference of a, the constant b was set at:

ln 0.5 b = a

4.2.6.3 Ordination of Biophysical Data

NMDS was also performed on each data set comprising of only biophysical data (see Table 4.1). Since most biophysical variables were interval data, the relationship between pairs of sites was examined using the Gower Metric dissimilarity coefficient (Gower and Legendre 1986). Non- interval variables, comprising LANDFORM (land surface morphology, surface curvature and aspect), SOIL (texture of horizon 1, 2, 3), and HYDROLOGY (sub-catchment drainage direction) were recoded to interval format. For surface morphology and surface curvature, separate binary variables were created for each possible value of each variable, comprising open depression,

91 uniform and uneven surface morphology, and convex, concave, flat and terrace surface curvature. For soil texture, binary variables were created to capture the most important textural information in a simpler form. These included a composite ‘organic’ texture class (comprising any texture class with 'organic', 'peat' or 'Sphagnum moss' in its name), as well as separate texture classes of ‘sand’, ‘loam’, ‘clay-loam’, ‘peat’ and ‘Sphagnum moss’. Variables containing true north aspect measurements were recoded to ‘degrees from north’ after being corrected for the difference between true and magnetic north. All ordinations were conducted in one to six dimensions, using ten random starting configurations.

Analysis of similarities (ANOSIM) was also used to test the null hypothesis that the average rank of dissimilarities derived from the biophysical data set did not differ significantly between breeding and random sites, using 10,000 random permutations. As for the analysis of floristic and structural data, this was conducted for all sites, and separately for sub-alpine and montane sites. Biophysical variables collected at all scales were included together in the analysis. The step-wise variant of ANOSIM (described above) was also conducted on the Gower Metric matrices to determine which variables contributed most to the observed difference between breeding and random sites at all, sub-alpine and montane elevations. Because of the constraint on collecting information from sub-soil horizons (see limitations above), and the fact that missing values had the potential to distort the analysis, soil information from horizons two and three for each site was omitted from the analysis.

4.2.6.4 Vector Fitting

Vector fitting was used as an analytical technique to investigate for correlations with variables in any direction across derived ordinations (see Kantvilas and Minchin 1989). This method locates the direction across an ordination along which the co-ordinates of a site (or quadrat) are most highly correlated with the values of the variable being examined. In this analysis, there was no attempt to reduce the number of variables prior to examination because of the studies exploratory nature, and that it may be more difficult to interpret relationships within the ordinations having removed some of the variables. As a result, all variables measured were examined, but only significant variables were presented in the ordination plots, comprising either those with the highest correlation, or those considered to be most informative. Other significant variables are presented in the relevant tables containing of the results from each ordination (see Appendix 4.1). In both 2-dimensional and 3-dimensional ordinations, the length of vectors for each variable was constructed in proportion to their correlation coefficient to aid in the interpretation. For vectors in 3-dimensional ordinations, it is the length in 3-dimensional space that is proportional to the correlation, not the apparent 2-dimensional length. The significance of derived correlation coefficients was tested by randomly permuting values (n = 1000) of each variable among the sites examined.

92 4.2.6.5 Analysis Programs

Data manipulation, vector fitting and interpretation was conducted using the DECODA program (Minchin 1991) in conjunction with the program KYST (Kruskal et al. 1973) to perform NMDS.

4.3 Results

4.3.1 Weighting Scheme for Height Class Comparisons

For floristic and structural data collected at the 10 x 10 m quadrat scale, the deltahalf value used in the weighting scheme for all quadrats was 1.5, and 3.0 for sub-alpine quadrats only (Fig. 4.2a, b). The graph for all quadrats shows that the amount of change in ANOSIM R was relatively small, indicating that a change in deltahalf did not make a large difference to the degree of separation between breeding and random sites. The graph for sub-alpine quadrats showed that there was an initial asymptotic increase in the separation of breeding and random sites until deltahalf reached approximately 3.0. For floristic and structural data collected at the 1 x 1 m quadrat scale, the relationship between ANOSIM R and quadrat height classes was different. The deltahalf value used in the weighting scheme for all quadrats (0.1) and sub-alpine quadrats (0.2) was considerably less than that selected for the 10 x 10 m quadrats, with discrimination between breeding and random sites becoming progressively worse as deltahalf was increased (Fig. 4.2c, d).

4.3.2 Ordination of Floristic and Structural Data from 10 x 10 m Scale Sites

4.3.2.1 All Sites

A 2-dimensional NMDS ordination (minimum stress = 0.175 from one of ten random starts) provided an adequate summary of all 124 10 x 10 m scale sites (Fig. 4.3). A 3-dimensional NMDS ordination (minimum stress = 0.138 from ten of ten random starts) did not display any new interpretable information about the distribution of breeding and random sites. Examination of the distribution of sites showed that random sites were more separated from breeding sites on the left side of the ordination (ANOSIM R = 0.12, p < 0.001), and that this separation was confined predominantly to sites from sub-alpine elevations (> 1270 m) (see below), and not at montane elevations (960 – 1270 m) (ANOSIM R = -0.01, p = 0.52).

Investigation for correlations with variables across the ordination showed that the distribution of random sites was associated with vectors that followed the plane running right to left and

93 approximately parallel to axis 1, including increasing altitude, cover of vegetation 0 – 1.8 m (%CV 0 – 1.8) and linear distance to the nearest locality of cool temperate mixed forest (LDNCTMF) (Fig. 4.3; Appendix 4.1, Table 1). The vector representing the number of calling males recorded in 10 x 10 m breeding sites was closely related to gradients that increased from left to right of the ordination, including cover of log/woody debris (%CL/WD) and distance to the nearest locality of sub-alpine wet heathland (LDNSWH). A number of other vectors were also diametrically opposed to the separation of random sites on the left side of the ordination, including an increase in cover of surface seepage (%CSEEP), total vegetation cover in the 27 – 41 m height class (TVC 27 – 41), distance to nearest locality of sub-alpine woodland (LDNSW) and more southerly aspects (ASPECT) towards the upper-right of the ordination, and an increase in cover of litter (%CLITTER) and total vegetation cover in the 2.5 – 5 and 5 – 8 m height classes towards the bottom-right of the ordination. A gradient relating to slope (RELIEF) and exposed rock height (AERH and MERH) increased towards the bottom of the ordination, running approximately parallel to axis 2, whilst uniform land surface (MORPH-F) and cover of standing water (%CSW) increase towards the upper-left of the ordination.

4.3.2.2 Sub-alpine Sites

A 2-dimensional NMDS ordination (minimum stress = 0.203 from ten of ten random starts) also provided an adequate summary of the 85 sub-alpine 10 x 10 m scale sites (Fig. 4.4). The extent to which random sites were separated from breeding sites in the sub-alpine data set was found to be even more significant than that recorded for all 10 x 10 m sites (ANOSIM R = 0.22, p < 0.0001), with separation occurring on the left and bottom sides of the ordination. A minimum stress of 0.154 from six of ten random starts was recorded in three dimensions, but no additional information could be interpreted.

Investigation for correlations with variables across the ordination showed that the environmental gradients described in the ordination comprising of all 10 x 10 m sites were also present, with a number of vectors relating more to variables at montane elevations being absent (% cover of logs/woody debris, % cover of litter, % cover of surface seepage and total cover of vegetation in the 27- 41 m height class), and those relating to sub-alpine elevations (total vegetation cover in the 0.3 – 1 m height class and distance to the nearest locality of montane riparian thicket) becoming significant (Fig. 4.4; Appendix 4.1, Table 2). Vectors that increased in the direction of random sites were uniform land surface (MORPH-F), cover of standing water (%CSW), cover of vegetation 0 – 1.8 m (%CV 0 – 1.8), distance to the nearest locality of montane riparian thicket (LDNMRT), altitude and total cover of vegetation in the 0.3 – 1 m and 0.1 – 0.3 m height class (TVC 0.3 – 1 and 0.1 – 0.3). Vectors that increased in the direction of breeding sites were more southerly aspects (ASPECT), distance to the nearest locality of sub-alpine wet heathland (LDNSWH), total vegetation cover in the 5 – 8 m, 2.5 – 5 m and 1 – 2.5 m height classes (TVC 5 – 8, 2.5 – 5 and 1 – 2.5), slope (RELIEF) and exposed rock height (AERH and MERH). The

94 vector representing the number of calling males recorded in 10 x 10 m quadrats was closely related to increases in total vegetation cover in the 2.5 – 5 m height class and distance to the nearest locality of sub-alpine wet heathland.

The best discrimination between random and breeding sites at sub-alpine elevations occurred with the retention of 41 from 145 floristic and structural attributes (maximum ANOSIM R = 0.45). The attributes that characterised breeding sites were those that reflected moist and shaded environments (Blechnum penna-marina, Coprosma perpusilla ssp. perpusilla, Epilobium spp., Hierochloe redolens, Huperzia australiana, liverwort (other), Luzula spp, moss and lichen, Olearia phlogopappa and Uncinia spp), seepage habitat on slopes with topographic protection (Leucopogon gelidus and Podocarpus lawrencei) and hydrological and ground surface features often associated with oviposition (surface seepage, log) (see lower portion of 10 x 10 m matrix presented in Appendix 4.3). Attributes that best characterised random sites reflected environments that were poorly drained and flat (Carex gaudichaudiana and Gonocarpus micranthus ssp. micranthus) or better drained with greater exposure (Epacris petrophila and Poa hiemata) (see upper portion of 10 x 10 m matrix presented in Appendix 4.3).

4.3.4 Ordination of Floristic and Structural Data from 1 x 1 m Scale Sites

4.3.4.1 All Sites

A 3-dimensional NMDS ordination (minimum stress = 0.188 from seven of ten random starts) was considered to provide a better summary of all 124 1 x 1 m scale sites (Fig. 4.5a, b). Inspection of the distribution of sites showed that there was a separation of random from breeding sites on the bottom and left side of the ordination displaying axes 1 and 3 (Fig. 4.5a), and on the top and left side of the ordination displaying axes 1 and 2 (Fig. 4.5b) (R = 0.16, p < 0.0001). Similar to 10 x 10 m sites, it was also apparent from the ordinations that random sites were more separated from breeding sites at sub-alpine elevations (> 1270 m) (see below) and not at montane elevations (ANOSIM R = -0.08, p = 0.95).

Investigation for correlations with variables across the ordinations showed that on the ordination displaying axes 1 and 3, vectors that increased in the direction of random sites were rock depth (ROCK DEPTH), soil saturation depth (SSD), ground surface-male calling site temperature differential (GS-BC TD), depth of soil horizon 2 (H2 DEPTH), sphagnum moss texture in soil horizon 1 (H1 SPHAGMOSS), altitude and the presence of peat in soil horizon 1 (H1 PEAT) (Fig. 4.5a; Appendix 4.1, Table 3). Vectors that increased in the direction of breeding sites were average exposed rock height (AERH), total vegetation cover in the 0.1 – 0.3 m and 0.3 – 1 m height classes (TVC 0.1 – 0.3 and 0.3 – 1), slope (RELIEF), conductivity and total vegetation cover in the 2.5 - 5

95 m height class (TVC 2.5 – 5). On the ordination displaying axes 1 and 2, vectors that increased in the direction of random sites were the presence of peat in soil horizon 2 (H2 PEAT), whilst those that increased in the direction of breeding sites included total vegetation cover in the 27 - 41 m height class (TVC 27 – 41), loam texture in soil horizon 1 (H1 LOAM), sand texture in soil horizon 2 (H2 SAND) and number of logs/woody debris 5 – 10 cm diameter (NO.L 5 – 10) (Fig. 4.5b). The vector representing the number of calling males recorded in 10 x 10 m quadrats was most closely associated with increasing number of logs/woody debris 5 – 10 cm diameter (NO.L 5 – 10) and sand texture in soil horizon 2 (H2 SAND).

4.3.4.2 Sub-alpine Sites

A 3-dimensional NMDS ordination (minimum stress = 0.185 from eight of ten random starts) was also considered to provide a better summary of the 85 sub-alpine 1 x 1 m scale sites (Fig. 4.6a, b). As for 10 x 10 m sub-alpine sites, there was also a better separation of random from breeding sites when examining only sub-alpine sites (R = 0.27, p < 0.0001). Inspection of the distribution of sites shows that there was a clear separation of random from breeding sites on the bottom and left sides of the ordination displaying axes 1 and 3 (Fig. 4.6a), and on the top, bottom, and left side of the ordination displaying axes 1 and 2 (Fig. 4.6b). Investigation for correlations with variables across the ordinations (Fig. 4.8) showed that most of the environmental gradients described in the ordination comprising of all 1 x 1 m sites were also present, with the exception of a number of variables relating more to habitat at montane elevations (conductivity, H1 LOAM and TVC 27 – 41) which were not recorded as significant (see Appendix 4.1, Table 4). Male calling site temperature (BC Temp) was an additional variable found to be correlated with an ordination axis, increasing in the direction of random sites at the top of the ordination displaying axes 1 and 2 (Fig. 4.6b). The vector representing increasing slope (RELIEF) was most closely associated with the vector representing number of calling males recorded in 10 x 10 m quadrats.

The best discrimination between 1 x 1 m random and breeding sites from sub-alpine elevations occurred with the retention of 28 from 106 habitat attributes (maximum ANOSIM R = 0.30). Similar to 10 x 10 m sub-alpine sites, the attributes that characterised breeding sites typically reflected habitats that were moist and shaded (Baeckea latifolia, Coprosma nitida, Isolepis spp., liverwort, Luzula spp., moss and lichen and Wittsteinia vacciniacea), topographically protected (exposed rock and Leptospermum grandifolium) and provided structural features often associated with oviposition (surface seepage and bare ground) (see lower portion 1 x 1 matrix presented in Appendix 4.3). Attributes that characterised random sites typically reflected habitats that were poorly drained or flat (Carpha spp., Empodisma minus and Sphagnum cristatum) or better drained (Olearia algida and Poa hiemata) (see upper portion of 1 x 1 matrix presented in Appendix 4.3).

96 4.3.5 Ordination of Biophysical Data

4.3.5.1 All Sites

Whilst a 3-dimensional NMDS ordination of all sites (minimum stress = 0.168 from ten of ten random starts) was considered to display further variation among sites with respect to the measured biophysical variables, a 2-dimensional ordination (minimum stress = 0.222 from six of ten random starts) was considered to provide a satisfactory summary of sites with respect to the distribution of breeding and random sites (Fig. 4.7). Similar to the ordinations of floristic and structural data from all 10 x 10 m and 1 x 1 m sites, there was a significant difference between the distribution of breeding and random sites (ANOSIM R = 0.21, p < 0.0001), with random sites being increasingly separated from breeding sites in the top half of the ordination. However, unlike the ordinations of floristic and structural data, where random sites were separated from breeding sites at only sub-alpine elevations, separation was apparent at both sub-alpine and montane elevations (see below for montane elevation). The separation of random sites in the top half of the ordination was associated with a number of variables, including vectors representing uniform land surface (MORPH-F10), peat in soil horizon 2 (H2 PEAT1), sphagnum moss in soil horizon 1 (H1 SPHAGMOSS1), and increasing soil saturation depth (SSD1), rock depth (ROCK DEPTH1), male calling site temperature (BC TEMP1) and depth of soil horizon 2 (H2 DEPTH1) (Fig. 4.7; Appendix 4.1, Table 5). Vectors representing cover of log/woody debris (%CL/WD10), relief (RELIEF10), total vegetation cover in the 0.3 – 1 m, 1 – 2.5 m, 5 – 8 m, 8 – 15 m and 27 – 41 m height class (TVC 0.3 – 110, 1 – 2.510, 5 – 810, 8 – 1510, 27 – 4110), cover of litter (%CLITTER10), average exposed rock height in 1 x 1 m quadrat (AERH1), uneven land surface (MORPH-S10), loam texture in soil horizon 1 (H1 LOAM1), CONDUCTIVITY1, sand texture in soil horizon 2 (H2 SAND1), exposed rock volume (ERV10), southerly aspect (ASPECT10), pH (PH1) and cover of surface seepage (%CSEEP10) increased in the direction of breeding sites. The vector representing number of calling males recorded in 10 x 10 m quadrats was associated with increasing surface seepage cover (%CSEEP10), relief (RELIEF10), uneven land surface (MORPH-S10) and sand texture in soil horizon 2 (H2 SAND1).

4.3.5.2 Sub-alpine Sites

A 3-dimensional NMDS ordination (minimum stress = 0.161 from ten of ten random starts) was considered to provide a better summary of sub-alpine sites (Fig. 4.8a, b). Separation of random from breeding sites was better than that recorded for all sites, and montane sites (below), with random sites distributed on the left side of ordinations displaying both axes 1 and 3 and 1 and 2. Vectors that increased in the direction of random sites on the ordination displaying axes 1 and 3 were male calling site temperature (BC TEMP1), soil saturation depth (SSD1) and ground surface-

97 male calling site temperature differential (GS-BC TD1), whilst those that increased in the direction of breeding sites were more southerly aspects (ASPECT10) and surface curvature-terrace (SC-T10) (Fig. 4.8a; Appendix 4.1, Table 6). Along the plane running approximately parallel to axis 3 was a climatic gradient relating to increased relative humidity (RH) at the top of the ordination and increased dry bulb temperature (TD) at the bottom. The direction of the vector representing increased dry bulb temperature also appeared to be in the direction of random sites, whilst the vector for relative humidity was more in the direction of breeding sites.

On the ordination displaying axes 1 and 2, vectors that increased in the direction of random sites were the presence of sphagnum moss in soil horizon 1 (H1 SPHAGMOSS1), rock depth (ROCK DEPTH1), depth of soil horizon 2 (H2 DEPTH1), peat texture in soil horizon 2 (H2 PEAT1), uniform land surface (MORPH-F10), surface curvature-flat (SC-F10) and cover of standing water (%CSW10), whilst those that increased more in the direction of breeding sites were total vegetation cover in the 0.3 – 1 and 1 – 2.5 m height classes (TVC 0.3 – 110 and 1 – 2.510), maximum exposed rock height in both 1 x 1 m and 10 x 10 m quadrats (MERH1 and MERH10), uneven land surface (MORPH-S10), sand texture in soil horizon 2 (H2 SAND1), relief in 10 x 10 m quadrat (RELIEF10), pH (PH1), total vegetation cover in the 2.5 – 5 and 8 – 15 m height classes (TVC 2.5 – 510 and 8 – 1510), peat in soil horizon 1 (H1 PEAT1) and distance to the nearest location containing sub-alpine wet heathland (LDNSWHS) (Fig. 4.8b). The vector representing increased altitude was positioned approximately between random and breeding sites, but tended to increase more in the direction of random sites.

The best discrimination between random and breeding sites from sub-alpine elevations occurred by retaining seven of the 66 structural attributes (maximum ANOSIM R = 0.49). Attributes that best characterised breeding sites were increasing slope in both 1 x 1 m and 10 x 10 m quadrats (RELIEF1 and RELIEF10), whilst those that characterised random sites were increasing rock depth (ROCK DEPTH1), soil saturation depth (SSD1), distance to cool temperate mixed forest (LDNCTMFS), and soils with a clay loam texture in soil horizon 1 (H1 CLAYLOAM1) and a yellow-brown soil in horizon 1 (H1 CHROMA1).

4.3.5.3 Montane Sites

A 2-dimensional NMDS ordination (minimum stress = 0.244 from ten of ten random starts) was considered to provide a satisfactory summary of montane sites (Fig. 4.9). A minimum stress of 0.163 from ten of ten random starts was recorded in three dimensions, but no additional information could be interpreted. Although significant, separation of random from breeding sites was not as pronounced as is was for the ordination of all sites and sub-alpine sites (ANOSIM R = 0.31, p < 0.0001), with greater separation between random and breeding sites appearing to occur towards the lower half of the ordination. Vectors that increased in the general direction of random sites were altitude, PH1, % cover of flowing water (%CFW10), open depression land surface

98 (MORPH-O10), average exposed rock height (AERH10), exposed rock volume (ERV10), light yellowish-brown soil horizon 1 (H1 CHROMA10 and VALUE10), light coloured soil (with variable colour) horizon 2 (H2 VALUE10) and clay loam texture in soil horizon 2 (H2 CLAYLOAM1), whilst those that increased in the general direction of breeding sites were CONDUCTIVITY1, uneven land surface (MORPH-S10), organic soil horizon 1 (H1 ORGANIC1), sand texture in soil horizon 2 (H2 SAND1), male calling site temperature (BC TEMP1) and total vegetation cover in the 0 – 0.1 m, 1 – 2.5 m, 5 – 8 m, and 27 – 41 m height classes (TVC 0.3 - 110, 1 – 2.510, 5 - 810, and 27 – 4110) (Fig. 4.9; Appendix 4.1, Table 7).

The best discrimination between random and breeding sites from montane elevations occurred by retaining eight of the 68 structural attributes (maximum ANOSIM R = 0.53). Attributes that best characterised breeding sites were uneven land surface (MORPH-S10), increasing length of drainage systems within 250 m radius of breeding site (DD250S) and increased total vegetation cover in the 8 – 15 m, 15 – 27 m and 41 – 55 m height classes (TVC 8 – 1510, 15 – 2710 - and 41 – 5510), whilst those that characterised random sites were increasing soil saturation depth (SSD1), wet bulb temperature (TW) and exposed rock cover (%ER10).

4.4 Discussion

4.4.1 Preferred Breeding Habitat Attributes

The results of this study show that selection of breeding sites by P. frosti was not random, and this result was related to specific macro and micro-habitat attributes and associated environmental gradients. Analysis of the combined data set comprising floristic and structural information (10 x 10 m and 1 x 1 m sites) detected a significant difference between random and breeding sites from sub-alpine elevations (> 1270 m), but not at montane elevations (≤ 1270 m), whilst analysis of the biophysical data set recorded significant differences between random and breeding sites at both elevations. This finding suggests that the biophysical structure of breeding habitat may be more important in determining the distribution of P. frosti breeding sites compared to floristics, although the separation of random and breeding sites from sub-alpine elevations suggests that floristics is also a reasonable predictor of breeding sites at higher elevation. Inspection of landform and drainage gradients present at sub-alpine elevations (see description of geomorphology and vegetation types in Chapter 2) shows that there are a greater diversity of breeding habitats (floristic communities) available compared to that which is present at montane elevations. The greater range of floristic and biophysical habitat attributes found to best discriminate between breeding and random sites from sub-alpine elevations also support this explanation. By comparison, landform on the south-western escarpment of Baw Baw Plateau at montane elevations is more uniform, and subsequently has a less diverse range of floristic communities.

99 Environmental gradients that best explained the distribution of breeding all sites from sub-alpine and montane elevations were decreasing altitude, wetter and cooler climatic conditions, greater topographic protection, steeper and uneven landform and wetter substratum conditions at a macro- habitat scale, and greater topographic protection, steeper landform, increasing cover of vegetation and woody debris, water chemistry (increasing conductivity and decreasing acidity) and soil structure (loam texture in soil horizon 1 (montane habitats) or peat soil horizon 1 (sub-alpine habitats) and sand texture in soil horizon 2 at a micro-habitat scale. At a macro-habitat scale, those that best explained the distribution of random sites included increasing altitude, reduced topographic protection and more uniform land surface. The presence of peat and sphagnum moss soil horizon 1, peat in soil horizon 2, increasing rock depth and depth of soil horizon 2, increasing soil saturation depth, increasing temperature at male calling sites and a larger temperature differential between the surface of the ground and male calling sites best explained the distribution of random sites at a micro-habitat scale.

The analysis of only sub-alpine sites identified similar environmental gradients to the analysis of all sites. Differences that were recorded related to the absence of gradients that were more indicative of habitat at montane elevations (e.g., cover of logs and woody debris, litter and total vegetation cover in the 27 – 41 m height class), and the addition of attributes that were more indicative of habitat at sub-alpine elevations (increasing cover of standing water and more uniform and flatter land surface. Similarly, the analysis of montane sites also identified the presence of gradients that were present in the ordination of all sites and sub-alpine sites, but also additional gradients not present in the other analyses, including vectors that increased in the direction of breeding sites (male calling site temperature), and those that increased in the direction of random sites (rock outcrop height and volume, cover of flowing water and ‘open depression’ land surface). Increasing cover and height of rock outcrop were typically associated with breeding sites in the analysis of all sites and sub-alpine sites, but apparently not at montane elevations, where rock outcrop was more associated with open gully habitats (open depression) comprising of fast flowing water (secondary drainage).

In contrast with the results at sub-alpine elevations, the vector representing temperature of male calling sites increased more so in the direction of breeding sites at montane elevations, suggesting that low temperature may be a constraint to breeding site selection. This result is interesting, given that low temperature was also recorded as factor limiting calling activity and movement by P. frosti at montane elevations, and that average ambient and substratum temperature during the breeding season were recorded to be lower at a montane elevation breeding site compared with a high elevation, sub-alpine breeding site (see Chapters 5 and 6). These observations suggest that P. frosti selects breeding sites within a narrow climatic range, being constrained by temperature conditions that can be too hot at sub-alpine elevations and too cold at montane elevations. This confinement to a narrow climatic range is also supported indirectly by population estimates

100 derived in Chapter 3, where the greatest density of calling males was estimated to occur at mid- elevation (1300 – 1400 m) between these two climatic extremes.

Drainage system density (DD250) was also recorded as one of eight variables that best discriminated for breeding sites over random sites at montane elevations, but not at sub-alpine elevations, or during the analysis of all sites. This result suggests that montane breeding sites are more likely to occur at localities with an increased density of drainage lines present. The non- association of DD250 at sub-alpine elevations suggests that there are other factors influencing the distribution of breeding sites at this elevation. A possible explanation for this is that a portion of drainage lines at sub-alpine elevations are also located in drier habitats that are unsuitable for breeding by P. frosti, such as sub-alpine shrubland and sub-alpine woodland. Increasing wet bulb temperature (TW) also discriminated random sites from montane elevations, but not at sub-alpine elevations, or in the analysis of all sites. This variable provides a measurement of temperature within an atmosphere that is 100% saturated, with random sites having higher temperatures recorded from them than breeding sites.

A number of habitat variables collected at macro and micro scales were also found to be associated the abundance of calling males recorded in 10 x 10 m quadrats. In the examination of all sites, increasing cover of log and woody debris, surface seepage, relief, distance to the nearest locality of sub-alpine wet heathland and uneven land surface were more associated with larger numbers of calling males at a macro-habitat scale, whilst sand texture in soil horizon 2 and increasing number of logs and woody debris 5 – 10 cm diameter were more associated with larger numbers of calling males at a micro-habitat scale. For sub-alpine sites only, there was a greater association with numbers of calling males with increasing total vegetation cover in the 2.5 – 5 m height class, maximum rock outcrop height and distance to the nearest locality of sub-alpine wet heathland at a macro-habitat scale, and increasing relief at a micro-habitat scale. A smaller sample size examined from montane elevations may have precluded detection of an association between number of calling males and ordination axes.

4.4.2 Historical Use of Breeding Habitat

Malone (1985a) provides the most detailed description of the breeding macro-habitat of P. frosti with which the results of this study can be compared. Although Malone’s observations were confined predominantly to elevations above 1300 m, his results indicate a clear preference by the species for unforested areas where 97.6% of calling males were recorded, with the remaining 2.4% being recorded from habitats with a sparse understorey of Nothofagus cunninghamii and/or Leptospermum grandifolium. Within unforested areas, he recorded the majority of calling males from wet alpine health-bog ecotone (75%), followed by wet alpine heath (22.5%), bog (1.6%) and grassland (modified) areas within the Baw Baw Alpine Resort (0.9%). By comparison, this study’s analysis of 10 x 10 m sites from sub-alpine elevations (almost identical to the area sampled by

101 Malone) suggests that many of the attributes of wet alpine health-bog ecotone, where Malone recorded most calling males, are associated with the distribution of random sites rather than breeding sites (e.g., increasing cover of standing water (pools), total vegetation cover in the 0 – 0.3 m and 0.3 – 1 m height classes and more uniform land surface). In contrast to this, habitat attributes associated with a sparse understorey of N. cunninghamii and/or L. grandifolium (where Malone only recorded 2.4% of calling males) appear more associated with the distribution of breeding sites in this study (increasing, total vegetation cover in the 1 – 2.5 m, 2.5 – 5 m and 5 – 8 m height classes, relief, rock-outcrop height and more southerly aspects). These contrasting observations suggest that populations of P. frosti observed by Malone (1985a) from sub-alpine elevations, when the density of the species was much higher than at present, were less selective in their preference for breeding habitat compared to extant population which appears to be using a smaller, subset of breeding habitat. This smaller, subset of breeding habitat represents a change from predominantly more open sub-alpine wet heathland communities in 1983 and 1984 to more topographically protected, cool, moist communities in this study. This pattern of breeding habitat preference is further supported in this study by the larger population estimates and densities recorded in habitats at lower elevation (Chapter 3) where these communities predominate.

As for macro-habitat, Malone (1985a) also provides the most detailed description of breeding micro-habitat (oviposition localities) with which to compare the results from this study. From elevations mostly above 1300 m, he recorded 79% of egg masses in natural cavities of vegetation that act as catchments for water travelling down slope, whilst the remaining 21% were recorded from beneath logs, rocks, and building materials from mostly disturbed habitats within the Mt Baw Baw Alpine Resort. The flora he recorded most often in association with these breeding cavities included Sphagnum sp., Astelia alpina var. novae-hollandiae, Empodisma minus, Epacris paludosa, Richea continentis and Carex spp. Malone also noted that due to the variability in species composition of breeding sites, and that some were associated with non-vegetative material, that structural characteristics were probably more important in determining suitable breeding habitat for the species. The results of this study’s analysis of micro-habitat support Malone’s observation about breeding habitat structure, with attributes relating to soil structure, water chemistry, drainage, micro-climate, surface cover and topographic protection identified as the most significant. Records of a number of breeding sites in modified or disturbed habitats from both sub-alpine and montane elevations (Malone 1985a; M. Littlejohn pers. comm.) also suggest that habitat structure is more influential. As for macro-habitat, this study also showed that floristic attributes of micro-habitat are also reasonable predictors of breeding sites at higher, sub-alpine elevation.

Both slope (relief) and rock (exposed rock outcrop) were identified in this study, and the study by Malone (1985a), as important micro-habitat attributes at breeding sites. However, Malone also recorded breeding habitat attributes that are in contrast with the attributes recorded in this study as characterising the location of breeding sites. Two species commonly encountered at oviposition

102 sites by Malone (E. minus and Sphagnum sp.) were recorded in this study as attributes that best discriminated for random sites, whilst the remaining species recorded by Malone were found not to discriminate between either random or breeding sites. Both E. minus and Sphagnum sp. are species typically associated with poorer drained and flatter micro-habitats. In this study, flora that best discriminated for breeding micro-habitats included those that reflected sites that were topographically protected (shaded) and moist, and often associated with the presence of surface seepages (water movement on slopes), including Baeckea latifolia, Coprosma nitida, Isolepis spp., liverwort, Luzula spp., moss and lichen, Wittsteinia vacciniacea and Leptospermum grandifolium. Although both studies recognise structural habitat attributes as being more important in the selection of breeding sites, flatter, more poorly drained micro-habitats, where Malone recorded a significant portion of oviposition sites, appear less likely to be used by the extant population of P. frosti.

Values of pH at breeding sites were mostly on the acid side of neutral (mean = 5.7, range = 4.9 – 7.2), with more neutral pH values being correlated with vectors pointing in the general direction of breeding sites, and more acid values pointing in the general direction of random sites. The flatter, more poorly drained sites described above appear to be linked to more acidic habitats. It is also these habitats that are subjected to hotter substratum temperatures, as indicated by the higher temperatures recorded at male calling sites. Environmental pH is an abiotic factor that may influence both abundance and distribution of species (Moore and Klerks 1998). In anurans, for example, sub-lethal and lethal effects in larvae have been attributed to low values of pH (McDonald et al. 1984), although complex interactions between various parameters associated with pH can complicate interpretation of pH tolerance in amphibians (Beebee 1996). Moore and Klerks (1998) showed that anuran larvae exposed to environments with low pH can lead to significant disruption of ion balance with a decreased active intake and an increased passive loss of critical ions. Further to this, temperature was also shown to act synergistically with decreased pH, such that increased temperature resulted in the regulatory ability of larvae to retain sodium to be exceeded. Other studies have also demonstrated amphibian sensitivity to low pH values (e.g., Sugalski and Claussen 1997; Vatnick et al. 1999). Notwithstanding the importance of biotic factors in determining amphibian distributions (Griffiths et al. 1993), factors limiting the current suitability of habitats for breeding by P. frosti may therefore relate to a combination of factors relating to habitat structure and water/soil properties.

4.4.3 Stable, Persistent and Moist Breeding Habitat

By demonstrating that P. frosti requires a stable, persistent, moist environment for breeding, this study supports the findings of Malone (1985a) that embryonic and larval developmental in the species is potentially sensitive to habitat modification or disturbance (see 4.1). Malone noted that vegetation appeared to be responsible for breeding cavities retaining their reservoir of water for longer periods in undisturbed habitats compared with the disturbed habitats that he examined. He

103 also suggested that whilst the foam nest appears to offer some protection for embryos from desiccation, the larval stage of P. frosti, given their confinement to small cavities for a relatively lengthy period (5 – 10 weeks), was clearly dependent on the persistence of a moist environment.

Although sub-alpine habitats at higher elevation typically comprise of dense vegetation cover < 1.8 m height, they generally have no overstorey, and only sparse midstorey cover. This vegetation structure may predispose sub-alpine habitats to be more vulnerable to both higher temperatures, due to decreased solar insulation, and periods of very low temperature, due to their higher elevation, particularly following habitat modification or disturbance. The situation at lower, montane elevations, where habitats appear to offer increased protection through the presence of greater overstorey and midstorey cover (cover of eucalypts and rainforest species) as well as a greater variety of surface cover attributes (large logs, rocks and vegetation), contrasts with habitats at sub-alpine elevation. A number of terrestrial salamanders appear similar to P. frosti with regards to their potential sensitivity to habitat modification or disturbance. Welsh and Lind (1996) showed that Rhyacotriton variegatus occupies a narrow range of physical and micro-climatic conditions in northwestern California, being associated with cold, clear headwater streams with loose coarse substrates in humid forest habitats with large conifers, abundant moss and dense canopy closure. They also point out that these attributes are characteristic of lotic systems found primarily in closed-canopy, complex-structured, older forests where the deep litter and abundant downed woody debris equate to greater terrestrial and aquatic micro-habitat complexity. The narrow range of moisture and temperature requirements by three relictual amphibians in southwestern Oregon and northwestern California examined by Welsh (1990) also appear similar to P. frosti. Because almost all of P. frosti’s habitat was burnt by a wildfire in 1939, its even age precludes assessment of a relationship with forest age. However, the significant association found between larger numbers of calling males and increasing cover of log and woody debris in this study suggests that habitats comprising older-aged forest, where woody debris is a character attribute (e.g., Scotts 1991; Woodgate et al. 1994), may be preferred by the species. Other studies have also demonstrated the importance of particular macro and micro-habitat in similar environments occupied by terrestrial amphibians (e.g., Ovaska 1991; Welsh and Lind 1995; Grover 1998; Marnell 1998; Wilkins and Peterson 2000; Davis 2002).

For P. frosti, sub-alpine habitats appear to provide considerably reduced habitat complexity compared with montane habitats. This reduced habitat complexity is due essentially to the absence of large eucalypts (Eucalyptus delegatensis ssp. delegatensis, E. nitens, E. glaucesens and E. regnans), a mostly absent midstorey, and absent or reduced cover of large logs and woody debris. The decline of the population from sub-alpine elevations (see Chapter 3) may be related directly or indirectly to the reduced ability of breeding habitats at sub-alpine elevation to retain moisture and provide adequate protection during periods of adverse or changing climate.

104 4.4.4 Related Taxa

Vegetation communities below 1400 m elevation appear to include breeding habitat attributes that are preferred by extant breeding aggregations of P. frosti. These communities include, cool temperate rainforest (central highlands montane scrub cool temperate rainforest, central highlands montane riparian cool temperate rainforest), cool temperate mixed forest (central highlands cool temperate mixed forest), montane riparian thicket (synonymous in part with central highlands montane scrub cool temperate rainforest) and montane wet forest ecological vegetation classes (see LCC 1991; Peel 1999, Table 2.1). Sub-alpine wet heathland also occurs below 1400 m, however, at this altitude it tends to merge with cool temperate rainforest (central highlands montane scrub cool temperate rainforest) or cool temperate mixed forest. Structurally, these rainforest vegetation communities are very similar to breeding vegetation types utilised by other members of the genus Philoria, which all have very similar life history strategies to P. frosti. These include montane sub-tropical rainforest (P. kundagungan), warm temperate and sub-tropical rainforest (P. loveridgei) and cool temperate rainforest, sub-tropical rainforest and wet sclerophyll forest (P. sphagnicolus) (Ingram and Corben 1975; Anstis 1981, 2002; De Bavay 1993; Hines et al. 1999).

Oviposition sites described for P. kundagungan, P. loveridgei and P. sphagnicolus are also generally similar to micro-habitats described for P. frosti in this study and previous studies (Littlejohn 1963; Malone 1985a), with each noted for using either water-filled cavities or burrows constructed from a variety of vegetative or non-vegetative substrates (Anstis 2002). A number of structural attributes found to be preferred by P. frosti in this study, however, stand out as being particularly similar to the other species: attributes associated with soil structure, hydrology, relief and substratum barriers (e.g., rocks, logs and vegetation roots). Soil profiles comprising of a peat, or organic material in horizon 1, over a sandy or gravely-textured horizon 2, were frequently encountered at breeding sites of P. frosti. Underground seepages on sloping terrain were typically associated with a coarse soil horizon 2, the distribution of which is likely to be related to processes associated with the weathering of the granitic parent material. Often located in association with soil horizon 2 were also granite rocks positioned independently within the profile. Most of the calling sites of P. frosti examined in this study were located on the base of this coarse-textured horizon 2, particularly at mid and high elevation where the granitic parent material appears closer to the ground surface. Substratum barriers (measured indirectly by cover of surface attributes) also provided physical obstructions for small reservoirs of water to establish. Oviposition descriptions for P. loveridgei and P. sphagnicolus appear similar to P. frosti, and include, burrows within free- draining moist soil, often with gravel, up to 15 cm below the surface (P. loveridgei), and burrows under rocks or logs, at the base of clumps of sphagnum moss, in rocky, sloping crevices where water drains through all sites, often over a deeper rock base (P. sphagnicolus) (Anstis 2002).

105 The macro and micro-habitat similarities between species of the genus Philoria suggest that the core breeding habitat of P. frosti may have always been the more protected, cooler, wetter montane rainforest gullies on the south-western escarpment of the Baw Baw Plateau, and that higher elevation, sub-alpine habitats, particularly on the north-eastern side, may only be occupied by populations during periods of more favourable climate. Support for this hypothesis is given by the greater densities of P. frosti recorded on the south-western side of the plateau compared with the north-eastern side taht were during population surveys conducted in 1983 and 1984 (Malone 1985a), and during this study between 1993 and 2002 (Chapter 3). A modelling study of orographically-forced postfrontal rain in the vicinity of the Baw Baw Plateau (Abbs and Jensen 1993) showed that at times of frontal passage, which typically arrive from the south-west, there is a complex interaction between the front (with its inherent clouds and precipitation) and the existing plateau generated clouds. This interaction results in a cap cloud above Mt Baw Baw and convective clouds on the western escarpment of the plateau. These convective clouds are particularly high in liquid water content, and may persist for very long periods (Abbs and Jensen 1993). Population estimates, historic and current patterns of distribution, and the pattern of climate described above, appear to support the argument that habitats located on the south-western escarpment of the Baw Baw Plateau represent the core area for breeding in P. frosti. ______This chapter has demonstrated that specific habitat attributes are preferred by P. frosti for breeding purposes, at a macro and micro-habitat scale. Biophysical habitat attributes appear more relevant to the selection of breeding habitat by the species compared with floristic attributes. Breeding habitats located within central highlands montane scrub cool temperate rainforest and central highlands cool temperate mixed forest communities on the south-western escarpment of the Baw Baw Plateau appear to represent the core breeding habitat of the extant population. The results also suggest that historical populations observed prior to this study, when densities were much higher, were less selective in their preference for breeding habitat compared to the extant population following the decline of the species. The use of a smaller subset of breeding habitat by the extant population is represented by a change from the use of predominantly open, heathland habitat at higher sub-alpine elevations, to climatically stable, persistent habitats at lower montane elevations. Preferred breeding habitat appears to be constrained to a narrow climatic range, being restricted by temperature conditions that can be too hot within some habitats at sub-alpine elevation and too cold within some habitats at montane elevation. Breeding macro and micro-habitats documented for P. frosti in this study are similar to attributes of breeding habitat reported for other members of the genus Philoria.

106 Table 4.1. List of environmental variables quantified at breeding and random sites, and details of their field measurement.

‘Scale’ refers to size of quadrat or area (1 x 1 m, 10 x 10 m and sub-catchment) examined.

Variable Ordination Scale (m) Details of Measurement Code LAND SURFACE STRUCTURE % Cover of Exposed Rock Outcrop %CER 10 x 10 % cover of exposed rock outcrop, estimated from line transect established between quadrat diagonal, with minimum considered length = 0.05 m Maximum Exposed Rock Height MERH 1 x 1, 10 x 10 estimated as maximum vertical height from lowest to highest point of exposed rock outcrop (m) Average Exposed Rock Height AERH 1 x 1, 10 x 10 estimated as the average vertical height from lowest to highest point of all outcropping rock (m) Exposed Rock Volume ERV 10 x 10 calculated as the % cover of exposed rock outcrop (m2)), multiplied by the average rock outcrop height Number of Exposed Rocks NO.ER 10 x 10 estimated as number of individual exposed rocks > 0.05 m diametre Number of logs 5-10 cm diameter NO.L 5 - 10 1 x 1 Number of logs 0.05 – 0.10 m diametre Number of logs 11-40 cm diameter NO.L 11 - 40 1 x 1 Number of logs 0.11 – 0.40 m diametre Number of logs >40 cm diameter NO.L > 40 1 x 1 Number of logs > 0.40 m diametre % Cover of Log/woody debris %CL/WD 10 x 10 % cover estimated from line transect established between quadrat diagonal, with minimum considered length = 0.05 m % Cover of Vegetation %CV 0 – 1.8 10 x 10 % cover 0 – 1.8 m estimated from line transect established between quadrat diagonal, with minimum considered length = 0.05 m % Cover of Bare Ground %CBG 10 x 10 % cover estimated from line transect established between quadrat diagonal, with minimum considered length = 0.05 m % Cover of Litter %CLITTER 10 x 10 % cover estimated from line transect established between quadrat diagonal, with minimum considered length = 0.05 m % Cover of Standing Water %CSW 10 x 10 % cover estimated from line transect established between quadrat diagonal, with minimum considered length = 0.05 m % Cover of Flowing Water %CFW 10 x 10 % cover estimated from line transect established between quadrat diagonal, with minimum considered length = 0.05 m % Cover of Surface Seepage %CSEEP 10 x 10 % cover estimated from line transect established between quadrat diagonal, with minimum considered length = 0.05 m VEGETATION Linear Distance To Nearest Locality of LDNSWH Sub-catchment Measured as linear distance from male calling site, or random site, to nearest known Sub-alpine Wet Heathland locality of sub-alpine wet heathland (Vegetation information from 1:25000 Baw Baw Plateau Map, Roberts 1996, and local knowledge). (no slope correction)

107 Variable Ordination Scale (m) Details of Measurement Code Linear Distance To Nearest Locality of LDNCTR Sub-catchment Measured as linear distance from male calling site, or random site, to nearest known Montane Riparian Thicket locality of montane riparian thicket (Vegetation information from 1:25000 Baw Baw Plateau Map, Roberts 1996, and local knowledge). (no slope correction) Linear Distance To Nearest Locality of LDNSW Sub-catchment Measured as linear distance from male calling site, or random site, to nearest known Sub-alpine Woodland locality of sub-alpine woodland (Vegetation information from 1:25000 Baw Baw Plateau Map, Roberts 1996, and local knowledge). (no slope correction) Linear Distance To Nearest Locality of LDNCTMF Sub-catchment Measured as linear distance from male calling site, or random site, to nearest known Cool Temperate Mixed Forest locality of cool temperate mixed forest (Vegetation information from 1:25000 Baw Baw Plateau Map, Roberts 1996, and local knowledge). (no slope correction) LANDFORM Altitude ALTITUDE 1 x 1, 10 x 10 metres above sea level Relief RELIEF 1 x 1, 10 x 10 For 1 x 1 m quadrats; slope measured in degrees from highest to lowest point in quadrat, and categorised as:1 = 0-0.5, 2 = 6-10, 3 = 11-15, 4 = 16-20, 5 = 21-25, 6 = 26-30, 7 =3 1- 35, 8 = 36-40, 9 = 41-45, 10 > 45 degrees, for 10 x 10 m quadrats; slope measured in degrees from highest to lowest substrate surface in quadrat, over the width of the quadrat, using a clinometer. Aspect ASPECT 1 x 1, 10 x 10 dominant slope direction of substrate surface measured in degrees from magnetic north using a compass, and converted to true north. Morphology MORPH 10 x 10 morphology of general land surface shape in quadrat and immediate area, categorised as V = open depression, F = uniform/flat, S = uneven/sloping (Following McDonald et al. (1990). Morphology-upslope MORPH-UP 10 x 10 element upslope from quadrat relative to relief, and categorised as: gentler slope = -1, the same slope = 0, steeper slope = 1. Morphology-downslope MORPH-DO 10 x 10 element downslope from quadrat relative to relief, and categorised as: gentler slope = -1, the same slope = 0, steeper slope = 1. Surface Curvature SC 10 x 10 curvature of surface within quadrat, incorporating upslope and downslope morphological elements, and categorised as: convex, concave, flat and terrace. HYDROLOGY Sub-Catchment Drainage Direction SDD Sub-catchment slope direction of sub-catchment measured in degrees from magnetic north using a compass, and converted to true north. Drainage Density250 DD250 Sub-catchment Sum total length of drainage systems (identified from 1:25000 hydrology layer, Roberts 1996) located within a 250 m radius of male calling site, or random site, using a planimeter. Soil Saturation Depth SSD 1 x 1 Estimated depth that soil profile became fully saturation, measured from within 2 m of male calling site, or in middle of 1 x1 quadrat for random sites, and categorised as: 1 = 0 – 0.10, 2 = 0.11 – 0.30, 3 = 0.31 – 0.60, 4 > 0.60 m. WATER CHEMISTRY pH PH 1 x 1 measured from water sample taken from male calling site, or nearest available water source for random sites, during the breeding season using a MultLine P4 Pocket Meter (Merick, 207 Colchester Road, Kilsyth, Victoria, 3137).

108 Variable Ordination Scale (m) Details of Measurement Code Conductivity CONDUCTIVITY 1 x 1 measured in us/cm from water sample taken from male calling site, or nearest available water source for random sites, during the breeding season using a MultLine P4 Pocket Meter (Merick, 207 Colchester Road, Kilsyth, Victoria, 3137). SOIL Depth of Soil Horizon 1, 2, 3 H DEPTH 10 x 10 measured from soil sample taken from within 2 m of male calling site, or in middle of 1 x 1 quadrat for random sites, and categorised as: 1 = 0 – 0.05, 2 = 0.06 – 0.10, 3 = 0.11 – 0.20, 4 = 0.21 – 0.30, 5 = 0.31 – 0.40, 6 = 0.41 – 0.50, 7 = 0.51 – 0.60, 8 > 0.60 m. Texture of Soil Horizon 1, 2, 3 H TEXTURE 10 x 10 measured from soil sample taken from within 2 m of male calling site, or in middle of 1 x1 quadrat for random sites, and categorised as: OS = organic sand, CS = coarse sand, LS = loamy sand, OCS = organic coarse sand, OLS = organic loamy sand, S = sand, CSL = coarse sandy loam, OSL = organic sandy loam, L = loam, OL = organic loam, SL = sandy loam, SCL = sandy clay loam, CL = clay loam, MP = muck peat, P = peat, GP = gravely peat, SP = sandy peat, SMP = sandy muck peat, SM = sphagnum moss, HP = humic peat (modified from McDonald et al. 1990). Hue of Soil Horizon 1, 2, 3 H HUE 10 x 10 estimated soil colour using the Munsell Colour System. Value of Soil Horizon 1, 2, 3 H VALUE 10 x 10 estimated soil colour using the Munsell Colour System. Chroma of Soil Horizon 1, 2, 3 H CHROMA 10 x 10 estimated soil colour using the Munsell Colour System. Rock Depth ROCK DEPTH 10 x 10 Estimated distance from surface to parent material in soil profile, measured from within 2 m of male calling site, or in middle of 1 x 1 quadrat for random sites, and categorised as: 1 = 0 – 0.05, 2 = 0.06 – 0.10, 3 = 0.11 – 0.20, 4 = 0.21 – 0.30, 5 = 0.31 – 0.40, 6 = 0.41 – 0.50, 7 = 0.51 – 0.60, 8 > 0.60 m. MACRO AND MICRO-CLIMATIC Wet Bulb Temperature TW 1 x 1, 10 x 10 Wet bulb temperature (Brannan Whirling Hygrometer (oC), Prospectors, Seven Hills, NSW) measured approximately 1.5 m above male calling site, or middle of 1 x 1 quadrat for random sites. Dry Bulb Temperature TD 1 x 1, 10 x 10 Dry bulb temperature (Brannan Whirling Hygrometer (oC), Prospectors, Seven Hills, NSW) measured approximately 1.5 m above male calling site, or above middle of 1 x 1 quadrat for random sites. Relative Humidity RH 1 x 1, 10 x 10 Relative Humidity (%) derived from wet and dry bulb temperature measured approximately 1.5 m above male calling site, or middle of 1 x 1 quadrat for random sites. Ground Surface Temperature GST 1 x 1 Temperature measurement (LCD Insertion Thermometer, range = -20 - 70, + 1oC, RS Components) of substrate directly above male calling site, or in middle of 1 x 1 quadrat for random sites, during breeding season. Male Calling Site (breeding cavity) BC TEMP 1 x 1 Temperature measurement (LCD Insertion Thermometer, range = -20 - 70, + 1oC, RS Temperature Components) of male calling site, or in middle of 1 x 1 quadrat for random sites at a depth of 0.15 m, during breeding season. Dry Bulb Temperature/Male Calling Site TD-BC TD 1 x 1 Difference between dry bulb temperature and male calling site temperature, or temperature Temperature Differential taken at 0.15 m depth for random sites.

109 Variable Ordination Scale (m) Details of Measurement Code Ground Surface Temperature/Male GS-BC TD 1 x 1 Difference between substrate surface temperature and male calling site temperature, or Calling Site Temperature Differential temperature taken at 0.15 m depth for random sites. Dry Bulb Temperature/Weather Station TD-TDW D 1 x 1, 10 x 10 Difference between dry bulb temperature and male calling site temperature, or Temperature Differential temperature taken at 0.15 m depth for random sites, and ambient temperature recorded at 1470 m a.s.l. in the Mt Baw Baw Alpine Resort (data logger, Tain Electronics Pty. Ltd., 10 Rowen Court, Box Hill North, 3129, Victoria, stored in a Stevenson Screen). Relative Humidity/Weather Station RH-RHW D 1 x 1, 10 x 10 Difference between relative humidity and male calling site temperature, or temperature Relative Humidity Differential taken at 0.15 m depth for random sites, and relative humidity recorded at 1470 m a.s.l. in the Mt Baw Baw Alpine Resort (data logger, Tain Electronics Pty. Ltd., 10 Rowen Court, Box Hill North, 3129, Victoria, stored in a Stevenson Screen).

110 Map Code Site Name Map Code Site Name 1 Access Rd 1 21a 46 Frangipanni Saddle 2r 2 Access Rd 1 21b 47 Frangipanni Saddle 3r 3 Access Rd 1 47 48 Frangipanni Saddle 4r 4 Access Rd 1 48 49 Frangipanni Saddle 5r N 5 Access Rd 1 49 50 Frangipanni Saddle 63 6 Access Rd 2 50 51 Frangipanni Saddle 64 7 Access Rd 3 56r 52 Hope Creek 52 8 Access Rd 3 57 53 Hope Creek 53 9 Access Rd 3 58 54 Hope Creek 80r 10 Access Rd 3 85r 55 Hope Creek left branch 56 11 Barnies Creek 84r 56 Jeep Track 25r 12 Baragwanath Flat 11 57 Long Creek 65 12 00 13 Baragwanath Flat 12 58 Long Creek 66 14 Baragwanath Flat 13 59 Long Creek 67 15 Baragwanath Flat 28 60 Long Creek side branch 69 110 16 Baragwanath Flat 85 61 Latrobe Plain 62 17 Buddies Rd Extension 87r 62 Latrobe Plain 69 # 18 Creek Corner 2 16 63 Latrobe Plain 69r 116 19 Creek Corner 23 64 Latrobe Plain 7 # 119 20 Creek Corner 2 35 65 The Morass 13r 200 105 117 1 50 21 Creek Corner 2 36 66 The Morass 14 # 51 # 22 Creek Corner 2 37 67 The Morass 14r 49 ## 106 23 Creek Corner 2 43r 68 The Morass 1r #

# 107 24 Creek Corner 2 47 69 The Morass 21r 47# # # 25 Creek Corner 4 70 The Morass 24 # #

# $ 26 Creek Corner 5 71 The Morass 25 46 # 48 Mt Whitelaw 1 0 1 0 115 00 1 27 Currawong Flat 17 72 The Morass 26 # # # 122 # 28 Currawong Flat 17r 73 The Morass 27 99 45 120 00 29 Currawong Flat 18 74 The Morass 29 4 # 123 1 # 30 Currawong Flat 19 75 The Morass 30 # 100 108 112 31 Currawong Flat 20 76 The Morass 31 111 32 Currawong Flat 23r 77 The Morass 32 # 109 33 Currawong Flat 24r 78 The Morass 33 121 # # 9 # 113 0 34 Currawong Flat 27r 79 The Morass 34 # 0 35 Currawong Flat 38 80 The Morass 4r 77 114 124 36 Charity Creek 59 81 Mustering Flat 14r 1300 78 # 81 37 Charity Creek 60 82 Mustering Flat 15r 76 # 75 118 38 Charity Creek 61 83 Mustering Flat 8r # 83 # 39 Charity Creek 82r 84 Neulyne Plain 6 79 # 74 82 # # 40 Charity Creek 83r 85 Pudding Basin 8 # 80 500 # 68 1 14 41 Chairlift 15 86 South Cascade 34r 00 # 29 30 # # 42 Chairlift 2 87 South Cascade 35r # # 73 # 93 43 Chairlift 48 88 Sewage Pond 54 67 # 65 44 Freeman Flat 17r 89 Sewage Pond 55 71 # 70 # 31 1500 32 # Baw Baw National Park # # 45 Frangipanni Saddle 1r 90 Sewage Pond 81r # 33 # 72 # # 69 91 Tullicouty 29r # # 28 $ 66 # # 44 Mt St Gwinear 22 34 92 Tullicouty 39r 26 25 Mt St Phillack 15 00$ 19 21 24 35 # 20 15 1 16 # 93 Tanjil Plain 1 23 500 27 # # 94 # 1 94 Tanjil Plain 10 # #### # 18 # # #5 56 ### 00 # 14 Contour height (metres a.s.l.) # # 1 95 Tanjil Plain 9 36 102 # 50 ## 0 91 37 $ # 5 1 2 12 Mt Baw Baw 96 Tyers River 31r # 103 # 13 # 38 # 1 ## 4 85 # 95 50 92 97 Tyers River 32r 39 # # # # 3 # 0 90 # 6 # 40 98 Tyers River 67r ## 7 150 88 # # 0 900 41 101 104 96 99 Tanjil River West Branch 68 # # 52 ## # # 1 43 # 97 Stream/creek/seepage/soak 89 53 30 100 Tanjil River West Branch 88r # # 0 # 8 # 62 54 42 84 64 101 Village Flat 2 22 # # # 0 9 # 40 102 Village Flat 2 23 10 63 61 # 861 # 103 Village Flat 2 32r # 55 104 Village Flat 46r 98 # 87 105 Whitelaw 11r 11 # 106 Whitelaw 12r 2 0 2 KilometersKilometres 107 Whitelaw 14r 1000 108 Whitelaw 18r 109 Whitelaw 27r 59 110 Whitelaw 2r 60 1 20 111 Whitelaw 30r # 0 # # # 112 Whitelaw 34r # 113 Whitelaw 37r Fig. 4.1. Distribution of breeding sites (red) and random sites (black) sampled between 114 Whitelaw 38r 57 17 00 1996 and 1999, Baw Baw Plateau and adjacent escarpment. Individual quadrats are 15 115 Whitelaw 39r 14 0 identified by number and name in adjacent table. 58 00 1400 0 $ 116 Whitelaw 3r 15 Mt Erica 117 Whitelaw 40 11 118 Whitelaw 40r 00 119 Whitelaw 41 11 120 Whitelaw 42 00 121 Whitelaw 43 122 Whitelaw 44 123 Whitelaw 45 124 Whitelaw Creek 46 Fig. 4.1. Distribution of breeding sites (red) and random sites (black) sampled between 1996 and 1999, Baw Baw Plateau and adjacent escarpment.

Individual quadrats are identified by number and name in adjacent table.

111

10x10 m - All Sites

0.118

0.117

0.116 (a) 0.115

0.114

0.113 00.511.5 22.53 3.5 4

10x10 m - Sub-alpine Sites

0.220

0.215

0.210 (b)

0.205

0.200

0.195 012345 678910

1 x 1 m - All Sites

0.164

0.162 NOSIM R Value 0.160 A

0.158 (c) 0.156

0.154

0.152

0.150

0.148

0.146 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5

1 x 1 m - Sub-alpine Sites

0.275

0.270

0.265

0.260 (d)

0.255

0.250

0.245

0.240 0 1 2345678910

Height Difference for 0.5 Weight

Fig. 4.2. Relationship between quadrat height class spacing and ANOSIM R statistic for floristic and structural data collected at 10 x 10 m and 1 x 1 m quadrat scales for all sites (a and c), and separately for only sub-alpine (> 1270 m) sites (b and d). Height classes for which a maximum R value was recorded were chosen as ‘deltahalf’ values to be applied to height class weighting schemes for each data set.

113 MORPH-F ASPECT %CSEEP LDNSW %CSW TVC 27-41 LDNSWH

NO.MALES %CL/WD %CV 0-1.8 Axis 2 ALTITUDE %CLITTER LDNCTMF TVC 2.5-5 RELIEF TVC 5-8 TVC 1-2.5 AERH TVC 0.3-1 MERH

Axis 1 Fig. 4.3. Distribution of all 10 x 10 m scale sites within a 2-dimensional NMDS ordination using floristic and structural data, including vectors of maximum correlation for biophysical

variables. z = breeding sites, { = random sites. See Table 4.1 for identification of variable labels.

ASPECT

LDNSWH MORPH-F NO. MALES %CSW TVC 2.5-5 TVC 5-8 TVC 8-15

Axis 2 %CV 0-1.8 RELIEF LDNMRT ALTITUDE MERH AERH

TVC 1-2.5 TVC 0.1-0.3 TVC 0.3-1

Axis 1

Fig. 4.4. Distribution of sub-alpine 10 x 10 m scale sites within a 2-dimensional NMDS ordination using floristic and structural data, including vectors of maximum correlation for

biophysical variables. z = breeding sites, { = random sites. See Table 4.1 for identification of variable labels.

114 (a) TVC 0.3-1 TVC 0.1-0.3 AERH RELIEF

NO. MALES ALTITUDE H1 PEAT TVC 27-41

H1 LOAM H2 PEAT PH H2 SAND Axis 3

CONDUCTIVITY NO.L 5-10 H1 SPHAGMOSS SSD GS-BC TD H2 DEPTH TVC 2.5-5 ROCK DEPTH (b)

Axis 1

(b)

H1 PEAT SSD TVC 27-41 H2 PEAT TVC 0.1-0.3 & 0.3-1 H2 DEPTH H1 LOAM GS-BC TD CONDUCTIVITY ROCK DEPTH PH ALTITUDE TVC 2.5-5 Axis 2 H1 SPAGMOSS

RELIEF AERH NO.MALES H2 SAND NO.L 5-10

Axis 1

Fig. 4.5. Distribution of all 1 x 1 m scale sites within a 3-dimensional NMDS ordination using floristic and structural data, including vectors of maximum correlation for biophysical

variables for axes 1 vs 3 (a) and axes 1 vs 2 (b). z = breeding sites, { = random sites. See Table 4.1 for identification of variable labels.

115 (a)

TVC 0.3-1 AERH TVC 0.1-0.3 ALTITUDE RELIEF

H1 PEAT NO.MALES TVC 1-2.5 H2 PEAT

H2 SAND Axis 3

PH H1 SPAGMOSS BC TEMP GS-BC TD NO.L 5-10 ROCK DEPTH TVC 2.5-5

SSD TVC 5-8

Axis 1

(b)

H1 SPHAGMOSS BC TEMP

ROCK DEPTH TVC 5-8 NO.L 5-10 NO.MALES RELIEF AERH & TVC 2.5-5 PH H2 PEAT H2 SAND TVC 0.1-0.3 Axis 2 GS-BC TD SSD TVC 1-2.5 TVC 0.3-1 ALTITUDE

H1 PEAT

Axis 1

Fig. 4.6. Distribution of sub-alpine 1 x 1 m scale sites within a 3-dimensional NMDS ordination using floristic and structural data, including vectors of maximum correlation for

biophysical variables for axes 1 vs 3 (a) and axes 1 vs 2 (b). z = breeding sites, { = random sites. See Table 4.1 for identification of variable labels.

116 ROCK DEPTH1 H2 DEPTH1 H1 SPAGMOSS1 MORPH-F10 1 1 1 10 H1 LOAM H2 PEAT SSD TVC 8-15 10 ASPECT10 %LITTER BC TEMP1 1 PH %CLWD10 TVC 5-810 CONDUCTIVITY1 TVC 27-4110 %CV 0-1.810 Axis 2 %CSEEP10 10 ALTITUDE NO.MALES H2 SAND1 10 10 1 ERV RELIEF H1 PEAT AERH1 MORPH-S10 10 TVC 0.3-1 10 TVC 1-2.5

Axis 1

Fig. 4.7. Distribution of all sites within a 2-dimensional NMDS ordination using only biophysical

data, including vectors of maximum correlation for variables. z = breeding sites, { = random sites. See Table 4.1 for identification of variable labels. Numbers in superscript following label indicate scale of collection: 10 = 10 x 10; 1 = 1 x 1; S = sub-catchment.

117 (a)

RH

ASPECT10 1 H1 SPAGMOSS 1 BC TEMP SC-T10 10 10 MORPH-F TVC 8-15 10 MORPH-S 1 %CSW10 H2 SAND MERH10 TVC 2.5-510 H2 DEPTH1 RELIEF10 LDNWSHS 1 Axis 3 ROCK DEPTH 10 MERH1 10 SC-F 1 NO.MALES SSD1 H1 PEAT1 PH H2 PEAT1 10 GS-BC TD1 TVC 1-2.5 TVC 0.3-110 TD

Axis 1

(b) (b)

1 ALTITUDE H1 PEAT TVC 0.3-110 RH MERH1

10 MORPH-F10 SC-F TVC 1-2.510 & MORPH-S10 %CSW10 H2 SAND1 1 SSD RELIEF10 H2 PEAT1 MERH10 NO.MALES10 1 H2 DEPTH PH1 1 SC-T10

Axis 2 ROCK DEPTH BC TEMP1 H1 SPAGMOSS1 10 1 TVC 2.5-5 GS-BC TD TD 10 ASPECT10 TVC 8-15 LDNSWHS

Axis 1

Fig. 4.8. Distribution of sub-alpine sites within a 3-dimensional NMDS ordination using only biophysical data, with vectors of maximum correlation for variables. (a) axes 1 v 3 and

(b) axes 1 v 2. z = breeding sites, { = random sites. See Table 4.1 for identification of variable labels. Numbers in superscript following label indicate scale of collection: 10 = 10 x 10; 1 = 1 x 1; S = sub-catchment.

118 CONDUCTIVITY1 1 H1 ORGANIC MORPH-S10 H2 SAND1 BC TEMP1 TVC 27-4110 TVC 0.3-110 TVC 5-810 TVC 1-2.510 TVC 0.1-0.310

H2 VALUE1 H2 CLAYLOAM1 Axis 2 10 %CLITTER H1 VALUE1 1 %CFW10 ALTITUDE H1 CHROMA 10 MORPH-O10 ERV AERH10 PH1

Axis 1

Fig. 4.9. Distribution of montane sites within a 2-dimensional NMDS ordination using only biophysical data, including vectors of maximum correlation for variables. z = breeding

sites, { = random sites. See Table 4.1 for identification of variable labels. Numbers in superscript following label indicate scale of collection: 10 = 10 x 10; 1 = 1 x 1; S = sub- catchment.

119

Chapter 5

CALLING BEHAVIOUR: ANNUAL, SEASONAL AND TEMPORAL VARIATION, AND THE INFLUENCE OF ENVIRONMENTAL FACTORS

5.1 Introduction

The number of studies undertaken on vocal communication and mating systems in anurans has increased rapidly in recent times, emphasising their usefulness in analysing behaviour and establishing evolutionary and phylogenetic relationships (Bush 1997; Penna 1997). Because of the importance of vocal activity to anuran reproductive success (Wells 1988), an understanding of the ecological and distributional limits within which breeding activity takes place could be considered fundamental to amphibian conservation, as well as studies investigating population declines. For example, monitoring the breeding patterns of amphibians with long-term data sets is considered important to more fully understand how we can mange threatened populations (Blaustein et al. 2001). Mating systems can also have a large effect on the way a population responds to natural and human-induced habitat changes as well as determining limits of distribution, so that an understanding of them is also important for management (Anthony and Blumstein 2000).

Studies on anuran communication show that there are a variety of vocalisations that have a specific function. Male calls have been shown to attract females, defend territories and determine intermale spacing (e.g., Littlejohn 1977; Wells 1977b, 1998; Gerhardt 1988; Ryan 1988, 1991; Brenowitz and Rose 1999; Bastos and Haddad 2002; Hermans et al. 2002); to be associated with courtship (e.g., Pengilley 1971; Wells 1977a; Townsend and Stewart 1986; Michael 1996; Bush 1997) and other short-range encounters (e.g., Tyler 1989; Bush 1997); and to occur in response to environmental changes (Tyler 1989). A range of biotic and abiotic factors has been shown to influence anuran vocal activity. Examples of these include light, water availability and temperature (Jaeger 1981; Gayou 1984; Navas 1996a, b; Woolbright 1985; Radwan and Schneider 1988; Salvador and Carrascal 1990; Navas and Bevier 2001; Hatano et al. 2002), as well as energy expenditure, physical condition, size and acoustic stimuli by conspecifics (Taigen and Wells 1985; Wells 1986, 1988; Wells and Taigen 1989; Docherty et al. 1995).

An interesting aspect of anuran vocal communication is the behaviour of species that occupy high- altitude, cold environments. Behaviours such as increased diurnal activity, basking, and associations with thermally buffered habitat have been reported as thermoregulatory activities that may result in more moderate activity temperatures being attained in high-elevation amphibians

121 (e.g., Pearson and Bradford 1976; Carey 1978; Bider and Morrison 1981; Sinsch 1989; Navas 1996b). Other behaviours, such as adopting vocal activity that is energetically less demanding, have also been demonstrated (Navas 1996a). Compared to other ectotherms, the ability of amphibians to maintain a preferred thermal environment for particular activities is further compromised by the need to maintain a hydrological balance (e.g., Brattstrom 1979), making amphibians sensitive to thermal extremes such as low temperature. In addition to behavioural adaptations, it is also likely that substantial modification of the thermal physiology of amphibians may enhance their capacity for vocal activity, particularly at low temperatures (e.g., Navas 1996a, b).

Anuran mating systems are highly variable, ranging from 'explosive' to 'prolonged' breeders (sensu Wells 1977a). For species with an explosive mating system, the breeding season is usually short, and females are synchronous in their sexual receptivity, compared to species with a prolonged mating system, where the breeding season is longer, and females breed asynchronously (Murphy 1999). As the male advertisement call in anurans is normally used to attract females and/or defend territories, it is suggested that increased calling activity rates at particular times may relate to periods when receptive females are available (Murphy 1999). This calling pattern is evident for species with prolonged breeding seasons when males cannot afford to expend energy calling all the time (Docherty et al. 1995).

The advertisement call of P. frosti has been described as a short ‘clunk’, repeated in sequences of up to 30 calls (Littlejohn 1963). The modulation of the call has been described as being irregular, being broken up into 3-7 pulses, and because of its irregularity a pulse repetition rate has not be determined (Littlejohn 1963). Previous studies on P. frosti indicate that it may fit into the category of a ‘prolonged’ breeder, with breeding occurring in late spring (November) and lasting for 2 - 3 weeks (Littlejohn 1963; Malone 1985a), whilst calling activity has been recorded over a period of approximately 8 - 9 weeks, commencing as early as the third week in October and continuing to as late as 24 December (Malone 1985a). Malone (1985b) noted that during the three weeks over which he observed oviposition in P. frosti, a peak in calling activity was also recorded. High levels of calling activity have also been recorded throughout the day, except when air temperatures fall below 5 oC (Malone 1985a). These observations on breeding in P. frosti are likely to refer predominantly to activity in sub-alpine habitat, as most observations on breeding aggregations by Malone’s (1985a) were confined to elevations above 1300 m in wet alpine heaths and wet alpine heath-bog ecotones.

The recent decline of P. frosti from higher altitude, sub-alpine habitat, and the persistence of a larger population in previously unsurveyed, montane habitat (see Chapter 3), suggests that, if related to environmental factors such as climate or habitat structure, there could be some measurable differences in breeding activity patterns and environmental conditions between the two habitat types. Other studies on amphibians have previously reported of the relationships between

122 timing of breeding and climate (Beebee 1995). The effects of climate change to breeding patterns of amphibians may subsequently result in significant changes to population structure, or population decline for sensitive species (Blaustein et al. 2001). In this study, calling activity of P. frosti was monitored over several breeding seasons to: (1) examine annual, seasonal and temporal patterns in breeding activity; (2) examine the climatic envelope within which calling activity takes place; (3) examine the influence of weather on rates of calling activity and participation; (4) further quantify call variation within the species; and (5) investigate for differences in calling activity and climate between sub-alpine and montane habitat types.

5.2 Methods

5.2.1 Automatic Recording Units

Calling activity was recorded during 1994 - 1999 breeding seasons using an automatic tape- recording unit placed at known breeding sites. Each unit consisted of a battery-powered portable Panasonic tape recorder (Model No. RQ-2102), with extendable microphone, attached to a battery- powered programmable Click timer (App. No. 090125), all placed within a waterproof ammunition box (Fig. 5.1). The extendable microphone was passed through an outlet in the ammunition box within a 15-mm diameter plastic tube for protection. Recording distance was estimated to be up to 50 m for P. frosti calls, but this was influenced by prevailing weather conditions (e.g., wind and rainfall) and micro-topography of calling sites. Most recording units were placed within a range of 10 m from known individual calling sites. Given the aggregative nature, and relatively low number of male P. frosti at breeding localities (see Chapter 3), this method of recording vocal activity was considered suitable for this species.

Annual, seasonal and temporal variation in calling activity by P. frosti was examined using two recording regimes programmed into individual units: (1) a 24-h regime encompassing six time periods (01:00, 07:00, 10:00, 13:00, 15:00 and 20:00 h Eastern Standard Time); and (2) a 24-h regime encompassing one time period only (13:00 h EST). An arbitrary recording interval of 5 min was considered sufficient to quantify calling activity at each time period. The recording regime encompassing six time periods was considered sufficient for examining diel patterns of calling activity, whilst the 13:00-h time period was used to examine calling activity within and between breeding seasons. The time periods of 07:00, 10:00, 13:00 and 15:00 h were selected to examine diurnal variation in calling activity, and 01:00 and 20:00 h for crepuscular and nocturnal calling activity. The selection of 13:00 h as the single recording period was based on the observation by Malone (1985a) that high levels of calling activity by P. frosti occurred throughout the day.

123 Diel patterns of calling activity were examined only during the 1994 and 1998 breeding seasons, with five and two recording units, respectively, whilst within and between season variation was examined between 1994 and 1999 with six units in 1994 and 1995, and two in years 1996 - 1999. Table 5.1 details the location, habitat, recording interval regime, year, number of calling males and number of recording units used to monitor calling activity.

5.2.2 Selection of Breeding Site

Knowledge about the permanency and number of frogs at known breeding sites, as recorded during annual survey monitoring (see Chapter 3), was used to select suitable breeding sites for placement of recording units in each year. Emphasis was placed on those breeding sites that were annually active, and those that contained the greatest number of calling males. In most cases this enabled monitoring of an aggregation of calling males to be monitored at each breeding site. Recording units were also located at breeding sites at both sub-alpine and montane elevations to examine for potential differences.

5.2.3 Quantification of Acoustic Data

Calling activity data were recorded on TDK cassettes (D120, IECI.TYPE I, Normal Position). For the 13:00-h recording regime, cassettes were removed and replaced every 12 days, whilst a four- day removal-replacement period was adopted for the recording regime encompassing six time periods. Data from the cassettes were manually transcribed onto data sheets using a Sony Professional Walkman (WM-D6C) and head phones. Calling activity was then quantified for each recording unit in the following manner: 1. the number of frogs calling during each time period; 2. the number of calls/frog/min during each time period; 3. the number of pulses/call/frog/min during each time period, calculated as the total number of pulses divided by the total of calls for each time period; 4. the number of courtship/amplexus/territorial calls/frog/min (referred to as 'growling' or 'growls') during each time period - a distinctive, softer note occurring in a sequence or as an independent note within a series of pulses (not quantified in 3); and 5. the intensity or speed at which a sequence of pulses was elicited within a call, estimated as one of four speeds: slow, medium, fast, very fast.

Calling intensity was measured to quantify a potential source of calling variation not quantified by the number of pulses/call/frog/min. Whilst quantifying frog calling rates, it was recognised that it was possible to record the same average number of pulses/call/min, but with a different pulse speed.

124 Quantification of calling activity and frog participation from each recording period was undertaken only when activity could be clearly distinguished. Undiscernible calling activity, due to either background noise or multiple numbers of males calling, was not quantified.

5.2.4 Acquisition of Weather and Micro-climatic Data

A battery powered data logger (Tain Electronics Pty. Ltd., 10 Rowen Court, Box Hill North, 3129, Victoria), stored in a Stevenson Screen and located at Village Flat (1470 m above sea level) in the Mt Baw Baw Alpine Resort, was used to monitor weather conditions. Ambient temperature (oC), relative humidity (%), solar radiation (W/m2, visible spectrum-700 nm) and rainfall (mm) were recorded at hourly intervals to examine the influence of weather on frog calling activity and participation. Due to the failure of the relative humidity probe between August 1997 and May 1998, relative humidity data were obtained from the Victorian Bureau of Meteorology for this period. These data were obtained from a weather station located in close proximity (approximately 400 m) to the data logger used in this study. In the absence of long-term climate data from the Baw Baw Plateau, data recorded from a weather station located at the south-eastern end of the plateau (Erica; 400 m a.s.l., Victorian Bureau of Meteorology) was used as a surrogate to examine long- term patterns in precipitation in relation to the breeding season of P. frosti.

From the 6 October to the 11 November 1999, data loggers (Tain Electronics) were located at a montane (1010 m) and sub-alpine (1485 m) breeding site to examine for potential thermal differences. Each was programmed to record ambient and substratum temperature at hourly intervals. For substratum temperature, the temperature probe was placed 150 mm below the ground surface to approximate the depth of a calling site, and positioned close to an aggregation of calling males. Ambient temperature probes were positioned 2.5 m above the ground and away from direct sunlight.

5.2.5 Data Analysis

5.2.5.1 Annual and Seasonal Calling Activity

Rates of annual and seasonal calling activity were examined by averaging call data recorded at 13:00 h on each day, from each recording unit, in each season. In 1994 from 8 - 15 December, only one recording unit was recording activity at 13:00 h, with the other four units recording at 15:00 h. To increase the sample size of the 1994 data set, data collected at 15:00 h were utilised as a surrogate for 13:00 h for this period. Each data set was reduced to include the period

125 encompassing the commencement and completion of calling activity. Distance-weighted least squares smoothing (tension = 0.5) was used to plot the relationship between time (date) and mean daily calling activity for each season. Kolmogorov-Smirnov two-sample tests were used to compare the distribution (mean, standard deviation and shape) of data from each season.

Because not all recording units were operating on each day during some breeding seasons, to statistically test for patterns in calling rates over the duration of each season it was necessary to truncate some data sets to encompass only those days when each recording unit was operating. This manipulation ensured an equal contribution by each respective unit over the duration of each season. For each season, this confined the analysis to: 24 – 25 and 28 - 30 November, and 1 - 16 December 1994; 26 - 31 October, 1 - 4, 6 - 12, and 14 - 30 November, and 1 - 19 December 1995; 22 October - 11 November 1996; 1 - 24 November 1997; 6 October - 5 November, 8 - 27 and 29 November and 2 December 1998; and 7 September - 3 November 1999. A preliminary examination of the data showed that they were not normally distributed due to the high frequency of cases with no calling activity during each season. A Friedman test was subsequently used to test the hypothesis that there was no systematic pattern in daily calling rates and intensity over the duration of each season. Mean calling rates and intensity derived from each recording unit were used as replicates in the analysis. This allowed for five and six replicates for 1994 and 1995 seasons, respectively, and two replicates for years 1996 - 1999.

Participation by calling males over the duration of each season was examined by deriving the mean daily proportion of males calling for each day of each season. For each breeding season, this was calculated by dividing the sum of the number of calling males recorded on each day at each recording unit by the sum of the largest number of calling males recorded at each recording site during a season. The Kolmogorov-Smirnov two-sample test was used to compare the distribution of participation data from each season.

5.2.5.2 Diel Calling Activity

The Friedman test was also used to examine for patterns in calling activity and calling intensity across six time periods (01:00, 07:00, 10:00, 13:00, 15:00 and 20:00 h) recorded from 1994 and 1998 breeding seasons. To exclude potential seasonal variation in calling activity, the analysis was confined to a group of days encompassing the peak of calling activity in each breeding season, and when each recording unit was operating. Calling activity peaks for each season were arbitrarily selected as a group of 11 days containing the highest call rates, as visually estimated from smoothing curves depicting seasonal calling rates in each year.

To ensure an equal contribution from recording units, the analysis of the 1998 data set was confined to nine days (16 - 24 October) and two sites (Tanjil Plain and Access Road 2), whilst the analysis of the 1994 data was confined to seven days (19, 20, 23 - 27 November) and four sites

126 (Access Road 1, Baragwanath Flat, Tanjil Plain and Neulyne Plain), although the unit at Tanjil Plain contributed to the analysis only on days 23 - 25 November. Inclusion of Tanjil Plain in the analysis, although only for three days, was considered important to increase sample size because of the relatively large number of males at the breeding site. Records of multiple calling activity from the same time period were averaged for each day from each recording unit. The Friedman test was also used to examine patterns in male participation across the six time periods. For each time period from each recording unit, combined, the proportion of calling males participating in calling activity for the selected days was used as a replicate in the analysis.

5.2.5.3 Variation in Calling Activity Between Males

Levene’s test for unequal variances (Levene 1960) was used to statistically examine the daily variance in male calling rates over the duration of each breeding season and at different time periods. Box plots were used to examining trends in the data for the specified groups. Seasonal patterns were examined by confining the analysis to the 13:00-h time period for each breeding season, thus removing potential daily variation. For the 1994 breeding season, calling activity recorded at 15:00 h were also included in the examination of seasonal patterns to increase sample size (see above). Diel patterns were examined by analysing the six time periods during the identified calling activity peak of each season. Calling variances derived separately for sub-alpine and montane habitat types were compared by visually examining box plots.

5.2.5.4 Frog Participation and Associations between Calling Rate and Calling Intensity

One-way ANOVA and the Kruskal-Wallis test were used to examine the influence of numbers of calling frogs (frog participation) on call rates > 0 (calls/min and pulses/call/min), growling rates and calling intensity. Averages derived from individual recording units in each year were used as the unit of replication for analysis, with data from sub-alpine and montane habitat being pooled for analysis. For call-rate associations, the potential influence of growling activity on call rate was excluded by including only cases where growling rates were considered to be low (≤ 1/min). To improve data distribution and variances, mean calls/min and pulses/call/min were square-root and natural-log transformed, respectively. In total, up to five calling males were assessed for their influence on calling activity. The Kruskal-Wallis test was used to examine the influence of male participation on growl rates and call intensity after transformations failed to normalise the data.

127 5.2.5.5 Associations Between Calling Activity and Weather

The relationship between weather and calling activity was investigated using data retrieved from the weather station located in the Mt Baw Baw Alpine Resort. Because of the large number of cases in the data where calling activity was zero, analysis of the relationship between weather and calling activity was undertaken in two steps: (1) correlation analysis between weather data and (a) calling activity data > 0 and (b) participation data; and (2) an analysis of calling activity and weather data using logistic regression analysis, with the response variable coded as either calling or not calling. Because the covariates for this analysis were either positively or negatively correlated with each other, a univariant analysis of each was conducted. Weather variables examined during the analysis included ambient temperature, relative humidity, solar radiation, total hourly rainfall, total daily rainfall and total rainfall from the previous three and five days.

Correlation Analysis

Seasonal and daily variation in call rates in each breeding season was first removed from the analysis by utilising data recorded only at 13:00 h during the identified calling activity peak of each breeding season. The potential influence of growling activity on calling rates was also removed from the analysis by including cases only where growling rates were considered low (≤ 1/min). This manipulation removed cases with a mix of both calls and growls, and when growling was the dominant activity. For analysis of call rates, mean calling data > 0 from each recording unit in each season was used as the unit of replication. Spearman rank correlation coefficients (probability values were obtained using look-up tables) are presented when variables could not be transformed to normality; otherwise Pearson correlation coefficients are presented.

The influence of weather on males participating in either calling or growling activity was examined by deriving the daily proportion of males calling for each recording unit. Proportions were calculated as already described, with the analysis being confined to recording units that had greater than one frog present. Recording units with only one frog present were considered to bias results, with a possible participation outcome restricted to 0% or 100%. This restriction excluded recording units from Access Road 1(b), Baragwanath Flat and The Morass in 1994, and Neulyne Plain in 1995. Analyses were conducted separately for sub-alpine and montane habitat types. Spearman rank correlation coefficients were also used to examine the relationship between weather and male participation.

Due to the presence of some missing data, all correlation coefficients were derived using listwise deletion.

128 Logistic Regression Analysis

Manipulation of the data set to only include data from the calling activity peak of each season resulted in the number of cases with no calling activity being too few for analysis of the six time periods separately, as well as precluding the grouping of each covariate to investigate potentially different responses to different levels of weather. To examine the influence of weather on the probability of recording P. frosti calling (coded as '1') or not calling (coded as '0'), logistic regression analysis was conducted on diurnally recorded data collected at 10:00, 13:00 and 15:00 h during the calling activity peak of each season. The pooling of data from all three time periods provided a sufficient sample size whilst confining the analysis to the time of day when high levels of calling activity are known to occur. Data recorded at 13:00 h were utilised from all years, whilst data from 10:00 and 15:00 h were available only from the 1994 and 1998 seasons. The analysis was conducted separately for calling and growling activity, and for sub-alpine and montane habitat types. Each covariate was considered continuous and independent, and the logit within each was presumed to be linear. To exclude the occurrence of zero cells, each covariate was collapsed into the following categories for analysis: ambient temperature (< 3, 3 - 4.9, 5 - 6.9, 7 - 8.9, 9 - 10.9, 11 - 12.9, 13 - 14.9, 15 - 16.9, 17 - 18.9, > 18.9 oC); relative humidity (< 40, 40 – 49.9, 50 - 59.9, 60 - 69.9, 70 - 79.9, 80 - 89.9, 90-100 %); solar radiation (0 - 199.9, 200 - 399.9, 400 - 599.9, 600 - 799.9, 800 - 999.9, > 999.9 W/m2); total rainfall (hourly) (0, > 0 mm), total rainfall (daily) (0, 0.1 - 4.9, 5 - 9.9, 10 - 19.9, > 19.9 mm); total rainfall (previous three days) (0, 0.1 - 9.9, 10 - 19.9, 20 - 39.9, > 39.9 mm); and total rainfall (previous five days) (0 - 0.9, 1.0 - 9.9, 10 - 19.9, 20 - 39.9, 40 - 59.9 > 59.9 mm). Logits were predicted from category means for each covariate and the response variable.

5.2.5.6 Thermal Dynamics of Micro-habitats

After recording calling activity for breeding seasons 1994 – 1998, it was recognised that data recorded from the weather station located at sub-alpine elevation may not be entirely appropriate for examining calling activity trends at lower, montane elevations. It was therefore of interest in this study to investigate and compare the thermal conditions present at a sub-alpine and montane breeding site over the duration of a breeding season (see measurement details in section 5.2.4). The relationship between ambient and substratum temperature recorded at hourly intervals at each breeding site was compared by examining regression slopes derived for each separate time period in each habitat. The significance of the ANCOVA interaction term (a * x * y) in the model (y = constant + a+x+y + a * x + a * x * y) was used as the test statistic for this comparison (homogeneity of slopes test).

Due to the lack of independence associated with repeated hourly temperature measurements, a comparison of regression slopes derived for separate hourly time periods within each habitat type could not be conducted. Regression slopes were subsequently compared by visual examination of

129 hourly time plots only. For a statistical comparison of the different temperature types from each habitat, the lack of independence associated with repeated hourly temperature measurements was removed from the analysis by arbitrarily using data recorded from a single time period only (13:00 h). One-way ANOVA and regression was then used to compare daily variation in ambient and substratum temperature recorded for the 13:00-h time period in each habitat. Ambient and substratum temperatures from both habitats were transformed (natural log) to improve data distribution and variances prior to analysis. As there were numerous cases where temperature measurements were ≤ 0 oC, to include all cases in the analysis following a natural log transformation, 5 oC was added to each temperature measurement prior to transformation (i.e., natural log + 5).

5.2.5.7 Duration of Calling Activity

Spearman rank correlation coefficients were used to examine the association between breeding- season duration (the interval encompassing commencement and completion of calling activity in each season) and weather recorded over the same period for years 1994 - 1999. The weather variables examined were mean breeding-season temperature, relative humidity, solar radiation and total rainfall. Because of potential climatic differences between sub-alpine and montane habitat, and that weather data were retrieved from a site at sub-alpine elevation, analysis was confined to data on calling activity from sub-alpine habitat. Smoothing curves depicting the seasonal trend in mean calling activity (calls/min) were used to estimate commencement dates for years 1994 and 1998 due to the absence of this information. Dates from remaining years were identified directly from data recorded from tape-recording units.

5.2.5.8 Weather during Commencement, Completion and Peak of Calling Activity

Because of the significant association between temperature and solar radiation and rainfall and relative humidity, only temperature and rainfall were selected to examine the relationships between weather and the commencement, completion and peak of calling activity in each breeding season. Included in the analysis of temperature conditions were data recorded prior to and after completion of each breeding season. These conditions were considered to have a possible influence on the commencement and completion of calling activity. The period encompassing 1 September to 31 December was chosen for examination in each season, except for 1994 and 1995 when the recording of temperature commenced only on 12 and 5 October, respectively. Linear regression analysis was used to identify the trend representing mean daily ambient temperature conditions prevailing over the duration of 1994 - 1999 breeding seasons at sub-alpine elevations. Dates identified for the commencement, completion and peak of calling activity for each breeding season at sub-alpine locations were then used to estimate prevailing mean ambient temperature

130 conditions at each respective time period using the derived regression formulae. The same 11 days chosen to represent the calling activity peak of each breeding season, above, was also used in this analysis. One-way ANOVA was used to compare mean daily ambient temperatures derived from the calling-activity peak of each season. Tukey’s pairwise mean comparisons were used to compare breeding season group means. A comparison was also made of the difference between ambient and predicted substratum temperatures during 1994 - 1999 breeding seasons. Substratum temperature was predicted from the derived relationship between mean daily ambient and substratum temperature recorded at a sub-alpine breeding site in 1999 (see section 5.2.5.6).

Visual examination of fortnightly rainfall totals was also undertaken to examine if there was any relationship between rainfall and the commencement and completion of breeding in seasons 1994 - 1999. As above, this examination was confined to data collected in sub-alpine habitat. Monthly precipitation records from Erica were also averaged across years 1932 - 1999 to inspect annual rainfall patterns with respect to the timing of breeding seasons of P. frosti.

5.3 Results

5.3.1 Call Structure and Variance

Calling activity by P. frosti was recorded from a total of 61 males over the duration of the study. An average of 87.2 ± 1.1 % of males present across all breeding sites had their calling activity quantified. The remainder were excluded due to background noise and/or undiscernible calling activity. Two discrete call types were recorded: (1) a call, made up of a number of consecutive notes (pulses), with a distinct break between each call; and (2) a distinct, softer note occurring in either a sequence or as an independent note within a series of pulses (growling). For all breeding seasons examined, the highest levels of calling activity recorded were 15 calls/min (mean = 3.5 ± 0.06), 99 pulses/call/min (5.5 ± 0.19), and 19 growls/min (0.79 ± 0.09). Over a 5-min recording interval, the highest number of consecutive pulses recorded was 160, the highest number consecutive calls was 75, and the highest number of consecutive growls was 93. These maximum rates of calling activity were retrieved from the seasonal peak of calling activity from all breeding seasons.

The variance in daily calling rates differed significantly over the duration of each breeding season (Table 5.2), with larger variances occurring as both call and pulse rates increased during the peak of calling activity compared to smaller variances at the beginning and end of each season (see box plots in Appendix 5.1). This trend was generally present in both sub-alpine and montane habitat types, although there was a pattern of larger variances in sub-alpine habitat for each season. For

131 growling activity, variances were also significantly larger during the peak in calling activity. There were no trends in variance evident for growling activity in different habitat types, although there was generally a lack of multiple records on the same day to allow for an assessment to be made.

The variance in daily calling rates also differed significantly over the six time periods examined in 1994 and 1998 (Table 5.3). However, in 1994, larger variances were associated with time periods when call rates were greatest (10:00, 13:00 and 15:00 h), whilst in 1998, variances were smaller during the period when call rates were greatest (Appendix 5.1). As for seasonal patterns, variances were generally larger in sub-alpine habitat compared to montane habitat across the six time periods. The variance in growling activity was also found to be significantly different across the six time periods examined, with increased variation at 10:00, 13:00 and 15:00-h time periods in 1998, whilst a variable pattern in variance was found across the six time periods in 1994. A lack of multiple records of growling activity over the six time periods precluded examination of variances between habitat types.

5.3.2 Annual and Seasonal Calling Activity, Duration and Male Participation

Mean calling activity rates and duration of breeding seasons 1994 - 1999 are shown in Figure 5.2. In 1994 and 1998, the commencement of the breeding season was not recorded, due to an unseasonal snowfall that prevented the placement of recording units at suitable breeding sites in 1994 (the first season of the study), and an unexpected early start to the breeding season in 1998. The call-rate trajectories for these years, as depicted by their smoothing curves, suggest that in 1994 calling activity may have commenced sometime in late October or early November, and in mid to late September for 1998. Subsequently, 4 November and 25 September were used as estimated commencement dates for 1994 and 1998, respectively.

Seasonal call-rate distributions (calls/min) were variable among years, with 1995 being significantly different to 1996 and 1997, and 1999 being different to 1996, whereas other comparisons were not significantly different (Table 5.4). Comparison of 1994 and 1998 distributions with other years was omitted due to their unknown starting periods. Using the estimated commencement times for 1994 and 1998, the duration of calling activity over all seasons ranged between 39 and 75 days (5.6 - 10.7 weeks), starting as early as 7 September and finishing as late as 19 December. This period encompassed spring and early summer (Fig. 5.3).

A distinct node of calling activity was recorded in most breeding seasons, with the number of calls/min reaching a peak during approximately the middle of each season, before and after which calling activity was less. The peak in calling activity for each season was estimated to be: 19 - 29 November in 1994; 9 - 19 November in 1995; 5 - 15 November in 1996; 3 - 13 November in 1997;

132 14 - 24 October in 1998; and 5 - 15 October in 1999 (Fig. 5.3). Including incomplete data sets from 1994 and 1998, this activity pattern was significant for all years except for 1996, when call rates were more uniform over the duration of the season (Table 5.5). Like call rates, the number of pulses/call/min also peaked during the middle of each season, except in 1996 (Fig. 5.2). Growling rates did not differ significantly over the duration of each season, although during 1994, 1995, 1997 and 1999, the smoothing curves suggest a tendency for higher growling rates during the period when call rates peaked (Fig. 5.2).

The identified calling-activity peak in each season was associated with an increase in the number of frogs participating in calling activity, before and after which frog participation was less (Fig. 5.4). The distribution of participation data for 1995 differed significantly from all other years, with no differences present for other year comparisons (Table 5.6). Comparison of 1994 and 1998 distributions with other years was omitted due to their unknown starting periods. Maximum frog participation recorded during the calling activity peak ranged from 44% in 1994 to 78% in 1999, as identified by the smoothing curves for each season (Fig. 5.4).

No relationship was found between numbers of calling frogs and the mean number of calls/min, but there was a significant effect of male participation on the mean the number of pulses/call/min (Table 5.7). Figure 5.5 suggests that the number of pulses/call/min increases as male participation increased, although the mean for five calling males was less than that for three and four calling males. Male participation was also found to significantly affect growling rates and calling intensity, with growling activity decreasing as participation increased (Fig. 5.6), but no pattern was evident for the relationship between participation and calling intensity.

Calling intensity was more variable than call rate over the duration of most breeding seasons. Although only significant for years 1994 and 1999, calling intensity tended to peak during the activity peak of each season, described above (Fig. 5.7, Table 5.5). Calling intensity was also found to be positively correlated with calling rates (Table 5.8).

5.3.3 Diel Calling Activity and Male Participation

Calling rate in both 1994 and 1998 was higher during the day (07:00, 10:00, 13:00 and 15:00 h) compared with dusk and nocturnal hours (20:00 and 01:00) (Fig. 5.8). This trend was highly significant for 1998 and approaching significance for the 1994 data set (Table 5.9). The number of pulses/call/min was also higher during the diurnal hours in both 1994 and 1998 (Fig. 5.9), but this was only significant for the 1998 data. No trend was apparent for growling activity across the six time periods. The number of frogs participating in calling activity was also significantly different over the six time periods in both years (Table 5.9), with higher participation occurring during the diurnal hours (7:00, 10:00, 13:00 and 15:00) compared with 20:00 and 01:00 h (Fig. 5.10). Calling

133 intensity also differed significantly over the six time periods, being similar to call rates and participation for 1994 and 1998 (Fig. 5.11). Mean calling intensity also tended to be similar in pattern to prevailing mean temperature for both seasons, with higher call intensities being associated with higher temperatures.

5.3.4 Calling Activity, Weather and Micro-climate

In sub-alpine habitat, ambient temperature and solar radiation were the only weather variables recorded as having a significant association with calling activity (mean calls/min > 0) (Table 5.10), with calling rate increasing as temperature and solar radiation increased (Fig. 5.12). Weather was not significantly associated with any other form of calling activity (growls/min and pulses/call/min) or frog participation in sub-alpine or montane habitat (Table 5.10). In contrast to this, calling intensity was significantly associated with most weather variables in both sub-alpine and montane habitat (Table 5.8). In sub-alpine habitat, there was a positive association with temperature and solar radiation and a negative association with relative humidity and hourly rainfall, whilst in montane habitat, there was positive association with temperature and negative association with relative humidity and rainfall from the previous three and five days (Table 5.8). Figure 5.13 shows the relationship between calling intensity and temperature and relative humidity variables in both habitat types.

In sub-alpine habitat, ambient temperature was a significant predictor of calling activity (excluding growling activity), with the probability of occurrence decreasing as ambient temperature increased (Fig. 5.14, Table 5.11). However, ambient temperature was not a significant predictor of calling activity in montane habitat. Both ambient temperature and relative humidity were significant predictors of growling activity in montane habitat, but not in sub-alpine habitat. Growling activity was more likely to occur as ambient temperature increased (Fig. 5.15) and as relative humidity decreased (Fig. 5.16). Neither solar radiation or the rainfall variables were predictors of calling or growling activity, although in sub-alpine habitat, a negative relationship with total rainfall from the previous five days approached significance (p = 0.053).

Although there were some visual differences in the regression slopes derived individually for hourly ambient and substratum temperatures measurements in each habitat (Fig. 5.17), a statistical comparison of slopes was not significant (Table 5.12). This result suggests a similar relationship between ambient and substratum temperature in both sub-alpine and montane habitat for different time intervals. The regression plots also show some slight differences between the time intervals at which the regression slopes change in each habitat type. The rate of increase in ambient temperature over that of substratum temperature (heating phase) appears to commence earlier in montane habitat than in sub-alpine habitat, but also commences to decrease (cooling phase) earlier in montane habitat than in sub-alpine. In montane habitat, the slopes appear similar to that of sub-

134 alpine habitat for the period between 22:00 - 03:00 h, but at approximately 06:00 h the slope becomes distinctively flatter. A flatter regression slope was maintained in montane habitat between 07:00 - 09:00 h. At 10:00 h, the slope in montane habitat became steeper again whilst the slopes for sub-alpine habitat continued to flatten up until 14:00 h, after which it became steeper again.

Statistical examination of ambient and substratum temperatures recorded at 13:00 h from sub- alpine and montane habitat sites showed that although being positively correlated, they differed significantly from each other both from within and between habitat types (Table 5.12; Fig. 5.18). Ambient temperature was significantly higher than substratum temperature in both habitats, whilst both substratum and ambient temperature in sub-alpine habitat was significantly higher than in montane habitat (Fig. 5.19).

Examination of the relationship between weather and duration of the breeding season suggests that average weather conditions prevailing before and during a breeding season influence the period over which calling activity occurs. Longer breeding seasons were positively associated with higher total rainfall (rs = 0.94) and relative humidity (rs = 0.37), and negatively associated with warmer

average temperatures (rs = -0.66) and higher levels of solar radiation (rs = -0.54) (Fig. 5.20).

There was a significant relationship between date and mean daily ambient temperature for the spring-summer period examined in each breeding season (Table 5.13), with a gradual increase in temperature occurring from start to finish (Fig. 5.21). Comparison of temperature conditions prevailing during the commencement of each breeding season (measured at the intercept of regression lines for each season) showed that variation in mean daily ambient temperature ranged from 2.9 to 5.6 oC (mean = 4.49 ± 0.43) over all seasons (Table 5.14). Figure 5.21 shows that mean daily ambient temperature may influence the commencement date of calling activity, with breeding seasons having flatter regression slopes appearing to start earlier than those with steeper slopes. Using the derived equation for predicting substratum temperature from ambient temperature (mean daily substratum temperature = 0.156*mean daily ambient temperature + 4.025; r2 = 0.52, n = 33, p < 0.005), mean daily substratum temperature ranged from 4.5 to 4.9 oC (4.67 ± 0.06). Over all seasons, temperatures associated with the completion of calling activity in each breeding season ranged from 5.4 to 9.5 oC (7.32 ± 0.58) for ambient temperature and 4.9 to 5.5 oC (5.11 ± 0.09) for substratum temperature (Table 5.14). From commencement to completion of each breeding season examined, there was a mean daily ambient and substratum temperature increase of 2.83 oC and 0.44 oC, respectively.

Inspection of call-rate (calls/min) distributions from both sub-alpine and montane habitat in each breeding season (see box plots in Appendix 5.1) indicates no clear difference in the period encompassing the calling-activity peak. The dates defining the calling-activity peak for each season identified above, were subsequently used to examine temperature conditions prevailing during this period. Analysis of mean daily ambient temperature during the activity peak of each

135 season showed that there was a significant difference between years (F = 375.72, df = 5, n = 66, p < 0.0005). A comparison of group means showed that all years were significantly different from each other (p < 0.0005), except for the 1996, 1998 and 1999 comparisons (Fig. 5.22). The group means ranged from 4.3 to 7.7 oC (5.32 ± 0.55), with those for 1994, 1995 and 1997 being higher than those for 1996, 1998 and 1999. By comparison, predicted mean daily substratum temperatures for the calling-activity peak of each season ranged from 4.7 to 5.2 oC (4.86 ± 0.09).

Figure 5.23 shows the relationship between total fortnightly rainfall and the commencement and completion dates of each breeding season in sub-alpine habitat. Graphs for each season indicate that the commencement of calling activity may be associated with an increase in fortnightly rainfall whilst completion might be associated with a decrease. Irrespective of the actual rainfall totals from each fortnight, the period over which calling activity occurred in each season appears to be generally associated with a pulse of rainfall activity.

A comparison of mean monthly rainfall data collected at Erica shows that each of the months from September - December received the greatest amount of rainfall, averaging greater than 100 mm. These four months coincided with vocal activity recorded from P. frosti over the breeding seasons examined in this study (see Fig. 5.24).

5.4 Discussion

5.4.1 Call Structure

The structure of the call of P. frosti, described as occurring in sequences of up to 30 calls (Littlejohn 1963) (quantified as pulses in this study), is within the range of pulse sequences recorded during this study. However, during the calling-activity peak, consecutive pulse sequences of up to 160 were recorded over a 5-min interval, and even longer sequences could have been recorded given a longer recording interval. The number of consecutive calls by a frog in this study also did not seem to be limited by time, with up to 75 consecutive calls being recorded from an individual frog over 5 min. Littlejohn (1963) recorded his data over a short period (25 - 26 November, 1960 and 11 November, 1961) from five calling males situated in a burrow at 8 oC. In this study, calls were recorded over the duration of entire breeding seasons, and from a variety of different weather conditions. Although the call of P. frosti has been described as irregular, complex and variable (Littlejohn 1963, 1987), there has been no formal recognition of the distinctive, softer, prolonged call recorded and described in this study as 'growling' (see below).

136 5.4.2 Annual and Seasonal Patterns of Calling Activity

Annual and seasonal differences in calling activity of P. frosti were detected over the six breeding seasons examined. The 75 days (10.7 weeks) over which calling activity was recorded in 1999 was slightly longer than the 8 - 9 weeks recorded by Malone (1985a) in 1983 and 1984, whilst the 5.6 weeks recorded in 1997 is considerably less than that observed by Malone. Calling activity commenced as early as 7 September in 1999, which was markedly earlier than that recorded by Malone (the third week in October), and continued at late as 19 December in 1995, this being similar to 24 December recorded by Malone. Although not recorded during this study, there are anecdotal reports of calling activity of P. frosti continuing into January (J. Coventry pers. comm.). Irregular, sporadic calling activity has also been recorded from the species after rainfall during February and March (G. Hollis pers. obs.). This pattern of calling activity suggests that breeding by P. frosti falls into the category of a 'prolonged breeder' (sensu Wells 1977a), where males of the species call to attract females and defend territories, and females breed asynchronously. In comparison with others members of the genus Philoria, calling activity has been recorded between August and February for P. kundagungan (Ingram and Corben 1975), late spring to January for P. loveridgei (Anstis 2002) and spring and summer for P. sphagnicolus (Anstis 1981; De Bavay 1993).

The influence of weather on anuran breeding activity is well documented (e.g., Wagner 1989; Sullivan 1992; Beebee 1995; Duellman 1995; Bertoluci 1998; Sullivan and Fernandez 1999; Marsh 2000; Bertoluci and Rodrigues 2002). In this study, weather was also shown to influence the breeding season of P. frosti, with longer seasons of calling activity being positively correlated with higher rainfall during the breeding season, and negatively correlated with cooler average temperature over the duration of each breeding season (Fig. 5.19). The reduced number of weeks over which calling activity occurred in 1997 (5.6 weeks) is thought to be related to the extremely dry conditions that prevailed at the time of breeding. This observation is supported by the unusual El Niño-Southern Oscillation (ENSO) features recorded at the time, which along with ENSO features recorded in 1982-83, are considered unprecedented in current records, and are included in seven of the most severe events that have occurred in the past 500 years (Carey and Alexander 2003).

The months over which calling activity of P. frosti has been recorded (September - December), both in this study and in previous studies (Littlejohn 1963; Malone 1985a, b), also coincides with the four wettest months of the year recorded from Erica (Fig. 5.24). As well as precipitation, average temperature also appeared to influence commencement of the breeding season, with earlier commencement of calling activity being associated with warmer early spring average temperatures (as occurred in 1998 and 1999) (Fig. 5.21). Considering that calling activity by P. frosti was recorded to occur significantly earlier in this study than previously by others (e.g., Littlejohn 1963; Malone 1985a), the apparent relationship between earlier commencement of

137 breeding activity and warmer weather supports a broad pattern emerging from other studies that some temperate-zone anurans are showing a trend towards breeding earlier in response to warmer climate (e.g., Blaustein et al. 2001). This study also showed that in sub-alpine habitat, the climatic window within which calling activity of P. frosti was recorded is very narrow, with average ambient and substratum temperature increasing by only 2.83 and 0.44 oC, receptively, from commencement to completion. This climatic window is likely to be even narrower in montane habitat due to its more uniform, buffered thermal range (see below). Other factors, such as the persistence of snow cover following the winter, may also influence commencement of calling activity by delaying attainment of temperature preferred for calling activity and the melting of snow to provide water at oviposition sites (e.g., Corn 2003).

The distinct calling-activity peak recorded in most breeding seasons examined was also supported by the observation of Malone (1985a, b) who noted that calling activity appeared to peak during the period over which recently-deposited egg masses were recorded. In this study, it was also during this period that most growling activity appeared to occur, although the result was not statistically significant. This call type could be associated with courtship, amplexus or defence of territories. During his intensive searching for egg masses of P. frosti, Malone (1985a) observed the presence of numerous silent males, and more than one male was frequently found at a calling site, suggesting that male interactions were common and that the growling call could be defensive in nature. In contrast to this, several field inspections of males participating in growling in this study found them to be accompanied by one or more females, and only the occasional satellite male being located close by, suggesting that growling may be associated with courtship or amplexus. In anurans that breed terrestrially, courtship calls are likely to be wide-spread, particularly where the male leads the female to the nest site (Wells 1988). Advertisement, territorial defence, courtship and amplexus calls have also been recorded for other anurans (e.g., Michael 1996; Penna 1997; Ovaska and Caldbeck 1997a, 1997b, 1999; Bourne et al. 2001). Verification of the functional significance of this call type by P. frosti could be obtained through call play-back experiments using selected call stimuli (e.g., Tárano 2002).

Unfortunately, the distribution and variance of data collected in this study did not allow for conventional, parametric, multi-variate analyses to be used to assess the effect of multiple variables on calling activity at one time. Nevertheless, independent examination of individual variables has provided insight into some aspects of the breeding activity of the species. The absence of a relationship between temperature conditions and the calling-activity peak recorded in each season suggests that the timing of this peak is not pre-determined by the onset of particular weather conditions. Although weather conditions have some influence on calling activity, social interactions between conspecifics (such as awareness by males of arriving females) appear more likely to be responsible for the peak in male calling activity and participation recorded in each season. In other anuran studies, temporal and spectral calling-activity cues, such as increases in calling rates, louder and longer calls, and calls with added notes have also been suggested as a

138 likely response to female stimulus (Straughan 1975; Halliday and Tejedo 1995; Navas 1996b; Bevier 1997; Brenowitz and Rose 1999), whilst acoustic stimuli from neighbouring males may also result in increased calling rates (Wells 1977a, 1988). Participation by calling males has also been shown to be stimulated by the arrival of receptive females (Murphy 1999).

Variances in calling activity were generally greater during the calling-activity peak of most breeding seasons for P. frosti (see box plots in Appendix 5.1). According to Docherty et al. (1995), an increase in call-rate variance is particularly evident for species with prolonged breeding seasons where males can not afford to expend energy calling all the time. Because anuran vocalisation has been shown to be energetically expensive (Pough et al. 1992), it is possible that P. frosti may be also adopting this strategy. However, the association between larger call variances and peak activity might also suggest conspecific interactions, such as the temporal availability of receptive females or intermale spacing. Although not presented in this study, calling rates and intensity also varied between breeding sites at which activity was recorded in the same breeding season. Such variation also suggests the influence of site-specific characteristics. For example, in an analysis of calling activity by Cophixalus ornatus across different sites, Brooke et al. (2000) determined that 35% of the variation in activity was due to factors common to all sites (weather, moon illumination, large scale social facilitation), but that 64.2% was due to site-specific factors such as small-scale social facilitation and micro-environmental physical factors that did not vary consistently over sites.

The inconsistent association between calling-male participation and calling-activity levels supports the view that other factors may have obscured a potentially clearer result. Participation of calling males was shown to have a significant influence on the number of pulses/call/min, but not on calls/min. This observation suggests that the number of pulses within each call increased as male participation increased, although the group with maximum male participation (5 males) did not record the highest pulse rate (Fig. 5.5). This group, however, contained only a small number of replicates (12) for analysis, which may have influenced the result. Social facilitation is also considered by others to effect calling participation by anurans on a day-to-day basis (Brooke et al. 2000). The lack of trend associated with the effect of male participation on calling intensity suggests that weather has a greater influence on calling intensity (see Fig. 5.13). The significant effect of male participation on growling rates suggests an interaction between calling males, although the resulting trend was for a decrease in growling rates as male participation increased. If growling rates were indicative of male territorial defence then one might have expected them to increase with greater male participation, although Wells (1988) argues that inter-male spacing is more important in influencing male interactions. Other studies have shown that anuran advertisement and encounter calls regulate intermale spacing within breeding choruses (Brenowitz and Rose 1999), and that the temporal features of calls, such as pulse repetition rate, are important for recognition in some species (Straughan 1975; Gerhardt 1988).

139 Although not quantified during this study, it was noticed that at breeding sites where several males were participating in calling activity at the same time, some would frequently alternate their pulse sequence to avoid overlap with the pulse sequence of neighbours, whilst others maintained independent pulse sequences. It was also apparent that males alternating their pulse sequence with a neighbour would compete with each other by increasing their calling intensity and call rate. These observations indicate that vocal interactions may have occurred between males in close proximity. The volume of recorded calls from males alternating their pulse sequence was also observed to be similar, suggesting calling positions in close proximity. Opposing temporal call patterns have also been observed in other anurans in response to broadcast stimuli or neighbouring conspecific calls (e.g., Penna 1997; Bosch and Márquez 2001; Lüddecke 2002).

5.4.3 Patterns of Diel Calling Activity

The higher calling rates, participation and call intensity recorded in this study during diurnal time periods (07:00, 10:00, 13:00 and 15:00 h), compared with the nocturnal and crepuscular time periods (01:00 and 20:00 h), was also broadly supported by Malone (1985a), who noted high levels of calling activity throughout the day during population surveys that he conducted. This suggests a preference by P. frosti for day-time chorusing. Although only two nocturnal time periods were examined, the results clearly show that calling activity at this time was considerably less than during diurnal time periods. This diel pattern of calling activity is similar to the diel pattern recorded for mean temperature (Fig. 5.11), particularly calling intensity which was significantly correlated with mean temperature.

A temporal shift towards diurnal calling activity could be interpreted as an adaptation by P. frosti, given its high-altitude, cold environment, to select preferred temperatures for calling or breeding activity. That calling activity by P. frosti was also recorded at 20:00 and 01:00 h suggests that either the climatic conditions were suitable on the particular days that it was recorded, or that other factors may have stimulated male calling activity regardless of temperature. Similar to the lack of association recorded between call rates and weather over each breeding season, the poor association between call rates and weather over the six time periods also suggests that interacting factors may have influenced calling activity at these times. The uneven variance recorded for calling activity across the six time periods also tends to support this conclusion. Other factors that may have influenced calling activity include the condition or energy reserves of the individual (e.g., Wells et al. 1995) or other circadian rhythms (e.g., Robertson 1976). Interpopulational shifts towards increased diurnality have been recorded for some amphibians at high altitude or latitude (e.g., Carey 1978; Bider and Morrison 1981), but not others (e.g., Navas 1996a). For most amphibians, it is recognised that the compromise between maintaining a thermal and hydric balance makes them considerably vulnerable to extreme thermal environments (Carey 1978; Brattstrom 1979). Preference by P. frosti for diurnal calling activity suggests a reduced hydric

140 constraint at this time, with the breeding season of the species coinciding with the wettest four months of the year, and selection of a greater range of thermal conditions during the day. A similar observation regarding hydric balance was made for the stream-dwelling frog, swinhoana, where seasonal activity was found to be more associated with air temperature than with precipitation due to readily available water (Kam and Chen 2000).

5.4.4 Habitat Variability

Given the associations recorded between calling activity and weather, and the recorded differences in daily temperature between sub-alpine and montane breeding sites, it is possible that calling activity may vary between sub-alpine and montane habitats. Although no consistent pattern in the commencement, completion and calling-activity peak of each season could be attributed to habitat type, a comparison of calling activity over the duration of each season suggested generally higher call rates and variances in sub-alpine habitat compared to montane (see box plots in Appendix 5.1). Comparison of ambient and substratum temperature recorded from sub-alpine and montane sites in 1999 showed that there were significant differences, both within and between temperature types and habitat types. Interestingly, both ambient and substratum temperatures from the lower altitude montane site were significantly lower than temperatures from the higher altitude, sub- alpine site (Fig. 5.18). In addition to this, the ambient temperature range recorded from sub-alpine habitat was substantially greater than the more uniform range recorded from montane habitat, with warmer and colder temperatures being recorded (Fig. 5.19). Substratum temperatures from both habitats, although colder in montane habitat, had low thermal variability, being buffered considerably when compared to external ambient temperatures.

The observed differences in micro-climate between sub-alpine and montane habitat can be interpreted as being due to differences in vegetation structure and their associated thermal dynamics. Although at lower altitude, the tall eucalypt overstorey and thick shrub understorey at the montane site, appeared to provide additional buffering from thermal extremes compared with the sub-alpine site, comprising of a thick, low understorey of heath and a largely absent overstorey (see chapter 4 for more detailed descriptions). The higher seasonal call rates and variances recorded from sub-alpine habitat could therefore be due to the higher daily temperatures and greater thermal range available to the species compared to montane habitat, where smaller variances and lower call rates were recorded. This pattern suggests that some breeding sites in montane habitat may have a reduced selection of thermal niches compared to sub-alpine habitat. Castellano et al. (1999) proposed that between population differences of acoustic properties may be the combined effect of three different properties: (1) differences in environmental thermal conditions; (2) differences in the relationships between micro-habitat and thermal conditions of calling males; and (3) differences in the relationships between thermal conditions and acoustic

141 properties. These properties could contribute to some of the calling variation recorded for P. frosti in different habitats.

This study also showed that higher temperatures prevailing in sub-alpine habitat may limit calling activity by P. frosti (Fig. 5.14), but not in montane habitat. In contrast, higher temperature was not found to limit the occurrence of growling activity in sub-alpine habitat, but lower temperatures and higher humidity were found to limit the occurrence of growling activity in montane habitat (Figs 5.15 and 5.16). There appears to be no obvious explanation for these observations, other than that the activities associated with these different vocalisations may possibly be different (i.e., advertisement call compared with courtship or amplexus call). For example, if growling activity by P. frosti is associated with courtship or amplexus, as opposed to advertisement or defence, then the lower probability of growling activity occurring at lower temperatures in montane habitat could imply reduced interactions between calling males and receptive females at low temperatures.

The significant relationship recorded between calling intensity and weather in both sub-alpine and montane habitat is interesting (Fig. 5.13), given that a significant association was recorded only between weather and call rate in sub-alpine habitat (Fig. 5.12), and call rates (calls/min and pulses/call/min) were significantly associated with calling intensity in both habitat types (Table 5.8). The lack of association between weather and calling rate in montane habitat could be due to the lower thermal variability recorded there, as noted above. However, it may also suggest that calling in sub-alpine habitat during extreme, cold temperatures may be energetically expensive for P. frosti, resulting in a reduction in calling rate, whilst maintaining the same calling intensity. Unfortunately, it is not possible from the results of this study to determine if calling activity responses by P. frosti from sub-alpine and montane elevations is different for cold temperatures of similar magnitude. However, differences in the call characteristics of other anurans occupying different elevations have been viewed as an adaptation to cold temperatures (e.g., Lüddecke and Sánchez 2002)

Observed differences in the diel pattern of call rates and male participation between sub-alpine and montane habitat types is also likely to be as a result of differences in micro-climatic conditions due to vegetation structure and associated thermal dynamics. In both 1994 and 1998, the diel pattern of calling activity (calls/min) in both habitats appeared to coincide with the heating phase recorded for the relationship between ambient and substratum temperature in both habitats (Fig. 5.17). In sub-alpine habitat, call rates were generally highest between 10:00 and 15:00 h, compared with montane habitat when they were highest between 07:00 and 13:00 h (see box plots in Appendix 5.1). Diel patterns of calling activity appear to be associated with the period when increases in ambient temperature are maximised relative to increases in substratum temperature. Relative differences in calling rates between sub-alpine and montane habitat in 1994 and 1998 could also be related to prevailing weather conditions during each season. Given that call rates were shown to be correlated with temperature in sub-alpine habitat, the higher calling rates recorded there in 1994

142 may simply have been due to the warmer weather recorded at the time, whilst the lower calling rates recorded in sub-alpine habitat in 1998 may have been due to cooler temperatures (see Fig. 5.21). ______This chapter has documented annual, seasonal and temporal variation in calling activity by P. frosti. It has also further described and quantified the call structure of the species, as well as alluding to its possible function. Recorded seasonal and temporal differences in calling rates and participation emphasises the need for survey monitoring of the species to account for variation in detectability of calling males. Macro and micro-climatic variation, due to differences in habitat structure and elevation, was shown to influence the duration, timing, intensity, rate and variance of calling behaviour by the species. The period encompassing calling activity was also shown to occur within a relatively narrow climatic window, indicating that reproductive activity by the species may be particularly sensitive to factors such as climate change. The recording of calling activity significantly earlier than previously recorded may be an example of the breeding response of P. frosti to warming temperatures during late winter and early spring during the later part of the 1990s, and may be linked to the decline of the species from sub-alpine habitats. There are clearly other factors not addressed in this study that also influence calling behaviour, such as social interactions. Further research is required in the area of acoustic communication in P. frosti to understand some of the observed patterns and associations that could not be explained in this study.

143 Table 5.1. Details of sites used to monitor calling activity of Philoria frosti on the Baw Baw Plateau, 1994 - 1999.

* denotes calling activity recorded at 15:00 h for the 13:00 h regime (see methods).

Year Site Altitude Habitat Recording Date (Diel Regime) Recording Date (13:00 h Regime) No. of (m a.s.l.) Type Calling Males 1994 Access Road 1 1250 Montane 17-27 Nov, 4-8 Dec., 18-21 Dec. 17-20 Nov., 23 Nov.-7 Dec., *8-15 Dec., 16 Dec., 18-21 Dec. 3 1994 Access Road 1(b) 1240 Montane 5-8 Dec., 17-21 Dec 28 Nov.-7 Dec., *8-15 Dec., 16-21 Dec. 1 1994 Baragwanath Flat 1490 Sub-alpine 18-28 Nov., 4-8 Dec., 16-21 Dec. 18-20 Nov., 22 Nov.-7 Dec., *8-15 Dec., 16-21 Dec. 1 1994 Neulyne Plain 1500 Sub-alpine 19-28 Nov., 4-8 Dec., 16-21 Dec. 19-21 Nov., 23 Nov.-7 Dec., *8-15 Dec., 16-21 Dec. 2 1994 Tanjil Plain 1505 Sub-alpine 22-25 Nov., 4-8 Dec., 18-21 Dec. 22-25 Nov., 28 Nov.-7 Dec., *8-15 Dec., 16 Dec., 18-21 Dec. 4 1994 The Morass 1340 Montane - 24 Nov.-21 Dec. 1 1995 Access Road 1 1250 Montane - 26 Oct.-31 Dec. 3 1995 Baragwanath Flat 1490 Sub-alpine - 26 Oct.-31 Dec. 2 1995 Neulyne Plain 1500 Sub-alpine - 26 Oct.-31 Dec. 1 1995 Tanjil Plain 1505 Sub-alpine - 26 Oct.-31 Dec. 3 1995 The Morass 1340 Montane - 26 Oct.-31 Dec. 2 1995 The Morass 2 1340 Montane - 26 Oct.-31 Dec. 3 1996 Access Road 1 1250 Montane - 21 Oct.-31 Dec. 2 1996 Tanjil Plain 1505 Sub-alpine - 22 Oct.-31 Dec. 5 1997 Access Road 2 1090 Montane - 1 Nov.-14 Dec. 4 1997 Tanjil Plain 1505 Sub-alpine - 8 Oct.-1 Dec. 4 1998 Access Road 2 1090 Montane 16 Oct.-20 Nov. 6 Oct., 13-15 Oct., 21 Nov.-5 Dec. 6 1998 Tanjil Plain 1505 Sub-alpine 16 Oct.-5 Nov., 8-20 Nov. 6-15 Oct., 21 Nov.-5 Dec. 5 1999 Access Road 2 1090 Montane - 31 Aug.-3 Nov., 11 Nov.-4 Dec. 5 1999 Tanjil Plain 1505 Sub-alpine - 31 Aug.-4 Dec. 4

144 Table 5.2. Results of Levene’s test for unequal variances on daily calling and growling rates recorded at 13:00 h over the duration of 1994 - 1999 breeding seasons.

N.B. activity recorded at 15:00 h in 1994 has been used as a surrogate for 13:00 h for the period 8 - 15 December (see methods).

1994 1995 1996 1997 1998 1999 Calls/min. x Date F = 7.54, df = 29, F = 3.52, df = 54, F = 1.63, df = 47, F = 1.97, df = 31, F = 1.76, df = 57, F = 2.33, df = 73, p < 0.0005 p < 0.0005 p < 0.05 p < 0.05 p < 0.005 p < 0.0005 Pulses/Call/min. x Date F = 5.12, df = 27, F = 2.19, df = 52, F = 4.82, df = 45, F = 6.35, df = 26, F = 7.23, df = 51, F = 5.56, df = 62, p < 0.0005 p < 0.0005 p < 0.0005 p < 0.0005 p < 0.0005 p < 0.0005

Growls/min. x Date F = 2.74, df = 29, F = 4.34, df = 54, F = 8.78, df = 47, F = 4.07, df = 31, F = 7.61, df = 57, F = 3.67, df = 73, p < 0.0005 p < 0.0005 p < 0.0005 p < 0.0005 p < 0.0005 p < 0.0005

Table 5.3. Results of Levene’s test for unequal variances on daily calling and growling rates recorded over six time periods (01:00, 07:00, 10:00, 13:00, 15:00 and 20:00 h) during the calling-activity peak of 1994 and 1998 breeding seasons.

1994 1998 Calls/min. x Time F = 4.40, df = 5, F = 4.40, df = 5, p < 0.005 p < 0.005 Pulses/Call/min. x Time F = 2.29, df = 5, F = 4.94, df = 5, p = 0.05 p < 0.0005 Growls/min. x Time F = 3.90, df = 5, F = 8.71, df = 5, p < 0.005 p < 0.0005

145 Table 5.4. Results of the Kolmogorov-Smirnov two sample test on mean daily call rates (calls/min) for 1995 - 1999 breeding seasons.

Results for comparisons with 1994 and 1998 have been omitted due to their unknown starting periods. D = maximum difference for group pairs.

Year 1995 1996 1997 1995 - - - 1996 D = 0.40, p < 0.005 - - 1997 D = 0.30, p < 0.05 D = 0.18, p = 0.47 - 1999 D = 0.16, p = 0.39 D = 0.29, p < 0.05 D = 0.24, p = 0.12

146 Table 5.5. Results of the Friedman test on mean daily calling and growling rates and mean calling intensity recorded at 13:00 h over the duration of 1994 - 1999 breeding seasons.

Fr = Friedman test statistic, KC = Kendall coefficient of concordance. N.B. activity recorded at 15:00 h in 1994 has been used as a surrogate for 13:00 h for the period 8 - 15 December (see methods).

1994 1995 1996 1997 1998 1999 No. of Recording Units 5 6 2 2 2 2

Calls/min. x Date Fr = 39.7, df = 20, Fr = 72.9, df = 52, Fr = 46.1, df = 47, Fr = 37.4, df = 23, Fr = 64.9, df = 48, Fr = 102.7, df = 66, p < 0.01, KC = 0.40 p < 0.05, KC = 0.23 p = 0.51, KC = 0.49 p < 0.05, KC = 0.81 p = 0.05, KC = 0.68 p < 0.005, KC = 0.78

Growls/min. x Date Fr = 14.8, df = 20, Fr = 35.3, df = 52, Fr = 24.6, df = 47, Fr = 27.2, df = 23, Fr = 35.3, df = 48, Fr = 62.5, df = 66, p = 0.79, KC = 0.15 p = 0.96, KC = 0.11 p = 0.99, KC = 0.26 p = 0.25, KC = 0.59 p = 0.91, KC = 0.37 p = 0.60, KC = 0.47

Pulses/Call/min. x Date Fr = 43.8, df = 20, Fr = 75.6, df = 52, Fr = 52.6, df = 47, Fr = 36.8, df = 23, Fr = 68.4, df = 48, Fr = 99.3, df = 66, p < 0.01, KC = 0.44 p < 0.05, KC = 0.24 p = 0.27, KC = 0.56 p < 0.05, KC = 0.80 p < 0.05, KC = 0.71 p < 0.01, KC = 0.75

Calling intensity x Date Fr = 31.4, df = 20, Fr = 68.2, df = 52, Fr = 48.5, df = 47, Fr = 32.4, df = 23, Fr = 59.6, df = 48, Fr = 99.3, df = 66, p = 0.05, KC = 0.31 p = 0.06, KC = 0.22 p = 0.41, KC = 0.52 p = 0.09, KC = 0.70 p = 0.12, KC = 0.62 p < 0.01, KC = 0.75

147 Table 5.6. Results of the Kolmogorov-Smirnov two sample test on calling participation by males for 1995 - 1999 breeding seasons.

Results for comparisons with 1994 and 1998 have been omitted due to their unknown starting periods. D = maximum difference for group pairs.

Year 1995 1996 1997 1995 - - - 1996 D = 0.34, p < 0.01 - - 1997 D = 0.41, p < 0.005 D = 0.18, p = 0.46 - 1999 D = 0.31, p < 0.01 D = 0.22, p = 0.12 D = 0.16, p = 0.58

Table 5.7. Results of one-way ANOVA and Kruskal-Wallis test on the relationship between male participation and mean calling and growling rates, and mean calling intensity, recorded at 13:00 h during the calling-activity peak of breeding seasons 1994 - 1999.

'SQR' and 'Ln' prefix = square root and natural log transformation, respectively.

Dependent Variable Factor/Grouping F/ Kruskal-Wallis test df n p Variable statistic SQRcalls/min No. of Calling Males F = 1.48 4 138 = 0.21 Lnpulses/call/min No. of Calling Males F = 2.46 4 138 < 0.05 Calling intensity No. of Calling Males K-W = 15.29 4 142 < 0.005 Growls/min No. of Calling Males K-W = 12.70 4 83 < 0.05

148 Table 5.8. Results of Spearman rank correlation analysis between mean calling intensity, recorded at 13:00 h during the calling-activity peak in 1994 - 1999 breeding seasons, and mean calling rates and weather.

Coefficients are presented separately for sub-alpine and montane habitat types. ‘n.s’ denotes non-significant p value (α = 0.05).

Variable Sub-alpine Montane calls/min rs = 0.68 n = 244 p < 0.0005 rs = 0.63 n = 198 p < 0.0005 pulses/call/min rs = 0.57 n = 244 p < 0.0005 rs = 0.60 n = 198 p < 0.0005

Ambient Temperature rs = 0.37 n = 244 p < 0.0005 rs = 0.42 n = 198 p < 0.0005

Relative Humidity rs = -0.31 n = 226 p < 0.0005 rs = -0.26 n = 178 p < 0.0005

Solar Radiation rs = 0.21 n = 244 p < 0.001 rs = 0.10 n = 197 n.s

Total Rainfall (hourly) rs = -0.15 n = 244 p < 0.05 rs = -0.02 n = 198 n.s

Total Rainfall (daily) rs = -0.11 n = 244 n.s rs = -0.03 n = 198 n.s

Total Rainfall (previous 3 days) rs = -0.01 n = 244 n.s rs = -0.14 n = 198 p = 0.05

Total Rainfall (previous 5 days) rs = 0.11 n = 244 n.s rs = -0.15 n = 198 p = 0.05

149 Table 5.9. Results of the Friedman test on calling and growling rates, calling intensity and participation over six time periods (01:00, 07:00, 10:00, 13:00, 15:00 and 20:00 h) during the calling-activity peak of 1994 and 1998 breeding seasons.

Fr = Friedman test statistic, KC = Kendall coefficient of concordance.

1994 1998 Days/Localities 7/4 9/2

Calls/min. x Time Fr = 10.3, p = 0.07, df = Fr = 24.5, p < 0.0005, df = 5, n = 22, KC = 0.09 5, n = 18, KC = 0.27

Pulses/Call/min. x Time Fr = 10.4, p = 0.07, df = Fr = 34.8, p < 0.0005, df = 5, n = 24, KC = 0.09 5, n = 22, KC = 0.39

Growls/min. x Time Fr = 1.8, p = 0.88, df = 5, Fr = 6.4, p = 0.27, df = 5, n = 24, KC = 0.87 n = 18, KC = 0.27

Proportion (%) of males calling x Time Fr = 11.2, p < 0.05, df = Fr = 32.4, p < 0.0005, df = 5, n = 7, KC = 0.32 5, n = 9, KC = 0.72

Calling intensity x Time Fr = 12.1, p < 0.05, df = Fr = 37.3, p < 0.0005, df = 5, n = 24, KC = 0.10 5, n = 18, KC = 0.41

150 Table 5.10. Results of correlation analysis between weather and mean calling-activity rates (call and growl rates > 0) and frog participation (proportion of calling males), recorded at 13:00 h during the calling- activity peak in 1994 - 1999 breeding seasons.

Spearman rank coefficients (rs) are presented for variables with distributions that could not be normalised. Pearson coefficients (r) are presented for variables with normal distributions. Analyses are provided for sub-alpine (S) and montane (M) habitat separately. Correlations were estimated using listwise deletion. * denotes calling-activity variable square-root transformed. The number of observations (n) associated with Pearson coefficients is provided separately. 'n.s' denotes non- significant p value (α = 0.05).

Mean no. of calls/min Mean no. of growls/min Mean no. of Proportion of males pulses/call/min calling S M S M S M S M n = 75 n = 50 n = 39 n = 34 n = 75 n = 50 n = 81 n = 99

Total rs = -0.20, rs = 0.09, rs = 0.28, rs = -0.12, rs = -0.13, rs = 0.18, rs = 0.02, rs = 0.13, Rainfall n.s n.s n.s n.s n.s n.s n.s n.s (hourly)

Total rs = -0.18, rs = 0.09, rs = 0.05, rs = -0.16, rs = -0.14, rs = 0.03, rs = 0.11, rs = -0.19, Rainfall n.s n.s n.s n.s n.s n.s n.s n.s (daily)

Total rs = -0.11, rs = 0.04, rs = 0.01, rs = -0.24, rs = -0.05, rs = 0.02, rs = 0.002, rs = -0.19, Rainfall n.s n.s n.s n.s n.s n.s n.s n.s (previous 3 days)

Total rs = -0.03, rs = -0.23, rs = 0.11, rs = -0.27, rs = 0.04, rs = -0.04, rs = 0.03, rs = -0.18, Rainfall n.s n.s n.s n.s n.s n.s n.s n.s (previous 5 days)

Solar *r = 0.27, *r = rs = -0.20, rs = 0.03, rs = -0.13, rs = -0.01, rs = -0.01, rs = -0.02, Radiation n = 84, 0.003, n = n.s n.s n.s n.s n.s n.s p.< 0.05 61, n.s

Ambient *r = 0.32, *r = 0.16, rs = 0.003, rs = 0.32, r = -0.13, r = 0.09, n rs = 0.07, rs = 0.15, Temperature n = 84, p n = 61, n.s n.s n.s n = 84, n.s = 61, n.s n.s n.s < 0.01

Relative rs = -0.19, rs = -0.09, rs = 0.22, rs = -0.15, rs = 0.05, rs = -0.22, rs = 0.06, rs = -0.06, Humidity n.s n.s n.s n.s n.s n.s n.s n.s

151 Table 5.11. Results of univariant logistic regression analysis on weather and calling activity (calls and growls) recorded during the peak of calling activity in 1994 - 1999 breeding seasons, in sub-alpine and montane habitat.

AT = ambient-temperature category, RH = relative-humidity category, SR = solar-radiation category, TRh = total hourly rainfall, TRd = total daily rainfall, TRp3d = total rainfall from previous three days, , TRp5d = total rainfall from previous five days, Ref. = reference (no calling), Resp. = response (calling), Est. = estimated coefficient, SE = standard error, t ratio = standardised coefficient, G = likelihood ratio statistic, bracketed values = constant, * denotes significant p values.

Call Variable Habitat Ref./Resp. Est. SE t ratio G df χ2 p value Type Call AT Sub-alpine 168/27 -0.091(2.728) 0.041(0.486) -2.248(5.613) 5.441 1 0.020* Montane 166/40 -0.025(1.661) 0.034(0.377) -0.731(4.406) 0.540 1 0.462 RH Sub-alpine 168/27 0.014(0.739) 0.010(0.808) 1.361(0.915) 1.796 1 0.180 Montane 166/40 -0.005(1.801) 0.009(0.761) -0.514(2.367) 0.269 1 0.604 SR Sub-alpine 168/27 0.001(1.331) 0.001(0.452) 1.185(2.946) 1.433 1 0.231 Montane 166/40 0.001(1.069) 0.001(0.392) 0.985(2.729) 0.977 1 0.323 TRh Sub-alpine 168/27 0.835(1.779) 1.026(0.212) 0.814(8.390) 0.833 1 0.361 Montane 166/40 0.813(1.360) 0.745(0.182) 1.092(7.476) 1.445 1 0.229 TRd Sub-alpine 168/27 -0.010(1.880) 0.026(0.249) -0.387(7.538) 0.145 1 0.704 Montane 166/40 -0.030(1.579) 0.022(0.216) -1.369(7.303) 1.771 1 0.183 TRp3d Sub-alpine 168/27 -0.009(1.990) 0.011(0.293) -0.828(6.787) 0.659 1 0.417 Montane 166/40 -0.010(1.578) 0.010(0.243) -0.975(6.492) 0.918 1 0.338 TRp5d Sub-alpine 168/27 0.005(1.666) 0.008(0.327) 0.617(5.086) 0.393 1 0.531 Montane 166/40 -0.004(1.548) 0.007(0.227) -0.602(5.597) 0.357 1 0.550 Growl AT Sub-alpine 59/136 0.006(-0.887) 0.028(0.296) 0.208(-2.993) 0.043 1 0.835 Montane 77/129 0.083(-1.330) 0.029(0.328) 2.865(-4.056) 8.745 1 0.003* RH Sub-alpine 59/136 0.015(-2.057) 0.009(0.740) 1.713(-2.778) 3.108 1 0.078 Montane 77/129 -0.021(1.097) 0.007(0.591) -2.793(1.856) 7.912 1 0.005* SR Sub-alpine 59/136 0.000(-0.951) 0.001(0.361) 0.358(-2.633) 0.128 1 0.720 Montane 77/129 0.000(-0.768) 0.001(0.337) 0.835(-2.281) 0.699 1 0.403 TRh Sub-alpine 59/136 0.455(-0.874) 0.536(0.164) 0.849(-5.344) 0.698 1 0.403 Montane 77/129 0.119(-0.528) 0.467(0.152) 0.255(-3.479) 0.065 1 0.799 TRd Sub-alpine 59/136 0.006(-0.867) 0.020(0.186) 0.317(-4.653) 0.099 1 0.753 Montane 77/129 -0.024(-0.406) 0.021(0.171) -1.133(-2.370) 1.354 1 0.245 TRp3d Sub-alpine 59/136 -0.003(-0.782) 0.009(0.213) -0.363(-3.663) 0.133 1 0.715 Montane 77/129 -0.011(-0.349) 0.009(0.193) -1.257(-1.803) 1.642 1 0.200 TRp5d Sub-alpine 59/136 -0.012(-0.455) 0.006(0.248) -1.876(-1.837) 3.756 1 0.053 Montane 77/129 -0.001(-0.496) 0.006(0.222) -0.115(-0.233) 0.013 1 0.908

152 Table 5.12. ANOVA, regression and ANCOVA (homogeneity of slopes test) analyses on ambient and substratum temperatures recorded at 13:00 h from sub- alpine and montane breeding sites between 6 October - 11 November, 1999.

All temperature variables were transformed (natural log + 5) to improve data distribution.

One-way ANOVA Dependent Variable Factor n MS df F p Sub-alpine Habitat Temperature Temperature Type (ambient/substratum) 70 1.361 1 13.384 < 0.0005 Montane Habitat Temperature Temperature Type (ambient/substratum) 69 2.218 1 44.560 < 0.0005 Substratum Temperature Habitat Type (sub-alpine/montane) 69 1.869 1 111.796 < 0.0005 Ambient Temperature Habitat Type (sub-alpine/montane) 69 1.071 1 7.860 < 0.01

ANCOVA (Homogeneity of Slopes Test) Dependent Variable Independent Variable n MS df F p Substratum Temperature Habitat Type*Ambient Temperature*Time Period 1637 0.015 23 1.317 = 0.14

Regression Analyses Dependent Variable Independent Variable r2 Coeff. MS d.f. F p Substratum Temperature (sub-alpine habitat) Ambient Temperature 0.40 2.015 (const) 0.066 34 6.378 < 0.05 (sub-alpine habitat) 0.101 (ambient) Substratum Temperature (montane habitat) Ambient Temperature 0.37 1.498 (const) 0.097 33 5.061 < 0.05 (montane habitat) 0.193 (ambient)

153 Table 5.13. Results of regression analysis for the relationship between mean daily ambient temperature and date for the period encompassing the 1994 - 1999 breeding seasons.

Regression coefficients are also provided for the relationship between predicted mean daily substratum temperature and date. Integer code for Date: 1 September = - 29; 1 October = 1; 1 November = 32, etc.

Year Independent variable Dependent Variable r2 Coeff. d.f. p 1994 Date (12/10-31/12) (1) Mean daily ambient temperature 0.49 1.644 (const); 0.110 (date) 80 < 0.0005 (2) Predicted mean daily substratum temperature 4.281 (const); 0.017 (date) 1995 Date (5/10-31/12) (1) Mean daily ambient temperature 0.28 3.064 (const); 0.045 (date) 87 < 0.01 (2) Predicted mean daily substratum temperature 4.503 (const); 0.007 (date) 1996 Date (1/9-31/12) (1) Mean daily ambient temperature 0.46 2.052 (const); 0.055 (date) 121 < 0.0005 (2) Predicted mean daily substratum temperature 4.345 (const); 0.009 (date) 1997 Date (1/9-31/12) (1) Mean daily ambient temperature 0.61 2.710 (const); 0.087 (date) 121 < 0.0005 (2) Predicted mean daily substratum temperature 4.448 (const); 0.014 (date) 1998 Date (1/9-31/12) (1) Mean daily ambient temperature 0.42 3.180 (const); 0.059 (date) 121 < 0.0005 (2) Predicted mean daily substratum temperature 4.521 (const); 0.009 (date) 1999 Date (1/9-31/12) (1) Mean daily ambient temperature 0.31 4.056 (const); 0.037(date) 121 < 0.005 (2) Predicted mean daily substratum temperature 4.658 (const); 0.006(date)

154 Table 5.14. Regression-derived mean daily ambient and substratum temperatures for different stages (calling-activity peak, calling commencement and calling completion) of 1994 - 1999 breeding seasons. Temperature estimates are for sub-alpine habitat only. * denotes commencement date predicted from smoothing curve (see methods).

1994 1995 1996 1997 1998 1999 Mean SE Range Breeding Season Start *4 November 26 October 22 October 24 October *25 September 23 September

Breeding Season Finish 11 December 19 December 8 December 21 November 2 December 8 November

Mean Daily Ambient Temperature (oC): Start 5.60 4.23 5.56 4.80 2.89 3.83 4.49 0.43 2.71

Mean Daily Ambient Temperature (oC): Finish 9.56 6.66 8.14 7.23 6.90 5.40 7.32 0.58 4.16

Mean Daily Substratum Temperature (oC): Start 4.89 4.69 4.54 4.78 4.48 4.63 4.67 0.06 0.41

Mean Daily Substratum Temperature (oC): Finish 5.51 5.06 4.97 5.18 5.09 4.86 5.11 0.09 0.65

Calling-activity peak 19 - 29 9 - 19 5 - 15 3 - 13 14 - 25 5 - 15 (11 days) November November November November October October

Mean Daily Ambient Temperature (oC): Calling-activity peak Average 7.69 ± 0.11 5.09 ± 0.05 4.31 ± 0.06 6.10 ± 0.09 4.30 ± 0.06 4.41 ± 0.03 5.32 0.55 3.39 + SE

Mean Daily Substratum Temperature (oC): Calling-activity 5.22 ± 0.02 4.82 ± 0.01 4.71 ± 0.01 4.99 ± 0.01 4.69 ± 0.01 4.71 ± 0.01 4.86 0.09 0.53 peak Average + SE

155 Fig. 5.1. Recording unit used to monitor calling activity by male Philoria frosti at breeding sites between 1994 and 1999, showing programable timer, cassette recorder and ammunition box.

156 Growls Pulses/call 12 Calls

10

8 1994 6

4

Mean calling activity/min 2

0 -30-20-10 0 10 20 30 40 50 60 70 80 90 Date (days from 1 October)

Growls Pulses/call 14 Calls

12

10

8 1995

6

4 Mean calling activity/min calling Mean 2

0 -30-20-10 0 10 20 30 40 50 60 70 80 90 Date (days from 1 October)

Growls Pulses/call 22 Calls 20 18 16 14 12 1996 10 8 6

Mean calling activity/min calling Mean 4 2 0 -30 -20 -10 0 10 20 30 40 50 60 70 80 90 Date (days from 1 October)

Fig. 5.2. Relationship between date and mean daily calling activity recorded at 13:00 h over the duration of 1994 - 1999 breeding seasons, as depicted by distance-weighted least squares smoothing. Date code: 0 = 1 October, 10 = 10 October, etc. N.B. commencement of 1994 and 1998 breeding seasons was not recorded, as indicated by incomplete smoothing curves; calling activity recorded at 15:00 h in 1994 has been used as a surrogate for 13:00 h for the period 8 - 15 December (see 5.2).

157 Fig. 5.2 cont’.

Growls Pulses/call 14 Calls

12

10

8 1997 6

4 Mean calling activity/min 2

0 -30-20-10 0 10 20 30 40 50 60 70 80 90 Date (days from 1 October)

Growls Pulses/call 20 Calls 18 16 14 12 10 1998 8 6

Mean calling activity/min 4 2 0 -30-20-10 0 10 20 30 40 50 60 70 80 90 Date (days from 1 October)

33.7 Growls Pulses/call 14 Calls

12

10

8 1999 6

4 Mean calling activity/min 2

0 -30-20-10 0 10 20 30 40 50 60 70 80 90 Date (days from 1 October)

158 Year

Grey = 1999 8 Black = 1998 Brown = 1997 7 Green = 1996 6 Blue = 1995 Red = 1994 5 4 3 2 1 Mean number of calls/min 0 31/8 10/9 20/9 30/9 10/10 20/10 30/10 9/11 19/11 29/11 9/12 19/12 29/12 Date (day/month)

Fig. 5.3. Comparison of distribution curves (distance-weighted least squares) for mean daily call rates (calls/min) recorded over the duration of 1994 - 1999 breeding seasons. N.B. commencement of 1994 and 1998 breeding seasons was not recorded, as indicated by their incomplete curves.

Year

Grey = 1999 100 Black = 1998 Brown = 1997 90 Green = 1996 80 Blue = 1995 70 Red = 1994 60 50 40 30 20 10 Mean proportion of males calling (%) calling males of Mean proportion 0 31/8 10/9 20/9 30/9 10/10 20/10 30/10 9/11 19/11 29/11 9/12 19/12 29/12 Date (day/month)

Fig. 5.4. Comparison of distribution curves for participation by calling males over the duration of 1994 - 1999 breeding seasons, as depicted by distance-weighted least squares smoothing. Participation data are expressed as the proportion of the largest number of calling males recorded at a breeding site. N.B. commencement of 1994 and 1998 breeding seasons was not recorded, as indicated by their incomplete smoothing curves.

159 0.6

0.4

0.2

0.0

-0.2 Ln mean number of pulses/call/min of number Ln mean -0.4 1 2 3 4 5 Number of calling males

Fig. 5.5. Relationship between number of calling males and call rate (pulses/call/min ± SE)

1.4

1.2

1.0

0.8

0.6

0.4

mean number of growls/min number mean 0.2

0.0 0 1 2 3 4 5 Number of calling males

Fig. 5.6. Relationship between number of calling males and mean growling rate (growls/min ± SE).

160 Year

Grey = 1999 Black = 1998 4 Brown = 1997 Green = 1996 Blue = 1995 Red = 1994 3

2 Calling intensity

1

0 31/8 10/9 20/9 30/9 10/10 20/10 30/10 9/11 19/11 29/11 9/12 19/12 29/12 Date (day/month)

Fig. 5.7. Relationship between date and calling intensity (1 = slow, 2 = medium, 3 = fast, 4 = very fast) recorded at 13:00 h over the duration of 1994 - 1999 breeding seasons, as depicted by distance-weighted least squares smoothing.

Year 5 1998 1994 4

3

2

1 Mean numberMean of calls/min

0 0 1 2 3 4 5 6 7 8 9 1011121314151617181920212223 Time (hours)

Fig. 5.8. Relationship between time and mean call rates (calls/min ± SE) recorded for six time periods (01:00, 07:00, 10:00, 13:00, 15:00, 20:00 h) during the calling-activity peak of 1994 and 1998 breeding seasons. 0 = 24:00, 1 = 01:00, 7 = 07:00 h, etc.

161 Year 2.0 1998 1994

1.5

1.0

0.5 Mean number of pulses/call/min of number Mean

0.00 1 2 3 4 5 6 7 8 9 1011121314151617181920212223 Time (hours)

Fig. 5.9. Relationship between time and mean call rates (pulses/call/min ± SE) recorded for six time periods (01:00, 07:00, 10:00, 13:00, 15:00, 20:00 h) during the calling-activity peak of 1994 and 1998 breeding seasons. 0 = 24:00, 1 = 01:00, 7 = 07:00 h, etc.

Year 1998 80 1994 70 60 50 40 30 20

Mean proportion of males calling (%) calling males of proportion Mean 10

00 1 2 3 4 5 6 7 8 9 1011121314151617181920212223 Time (hours)

Fig. 5.10. Relationship between time and mean calling participation (± SE) for six time periods (01:00, 07:00, 10:00, 13:00, 15:00, 20:00 h) during the calling-activity peak of 1994 and 1998 breeding seasons. Participation data are expressed as the proportion of the largest number of calling males recorded at a breeding site. 0 = 24:00, 1 = 01:00, 7 = 07:00 h, etc.

162 Year 1998 1994

4 15 (oC) temperature Mean

3 10

2 5

Mean calling intensityMean calling 1 0

00 1 2 3 4 5 6 7 8 9 1011121314151617181920212223-5 Time (hours)

Fig. 5.11. Mean temperature (± SE) (dotted lines) and mean calling intensity (± SE) (solid lines) (1 = slow, 2 = medium, 3 = fast, 4 = very fast) recorded from six time periods (01:00, 07:00, 10:00, 13:00, 15:00, 20:00 h) during the calling-activity peak of 1994 and 1998 breeding seasons. 0 = 24:00, 1 = 01:00, 7 = 07:00 h, etc.

Ambient temperature Solar Radiation

20 1100

C) 1000 o ) 15 2 900 800 10 700 600 5 500 400

mbient temperature ( 300

A 0

Solar Radiation (W/m 200 -5 100 0 1 2 3 4 0 1 2 3 4 Mean SQRcalls/min > 0 Mean SQRcalls/min > 0

Fig. 5.12. Relationship between mean call rate > 0 (calls/min - square root transformed) and weather (ambient temperature and solar radiation) in sub-alpine habitat, for activity recorded at 13:00 h during the calling-activity peak of 1994 - 1999 breeding seasons.

163 Sub-alpine habitat Montane habitat 25 25 C) C) o 20 o 20

15 15

10 10

5 5 mbient temperature ( temperature mbient A 0 ( temperature mbient A 0 -5 1 2 3 4 -5 Mean calling intensity 1 2 3 4 Mean calling intensity

Sub-alpine habitat Montane habitat 100 100 90 90 80 80 70 70 60 60 50 50

Relative Humidity (%) Humidity Relative 40 Relative Humidity (%) 40 30 30 20 20 1 2 3 4 1 2 3 4 Mean calling intensity Mean calling intensity

Fig. 5.13. Relationship between mean calling intensity (1 = slow, 2 = medium, 3 = fast, 4 = very fast), ambient temperature (oC) and relative humidity (%) in sub-alpine and montane habitat, as recorded at 13:00 h during the calling-activity peak of 1994 - 1999 breeding seasons.

164 Sub-alpine habitat

1.0 p = 0.02

0.9

0.8

0.7 Probability of calling 0.6

0.5 0 5 10 15 20 Mean category temperature (oC)

Fig. 5.14. Relationship between ambient temperature and the probability (with upper and lower bounds) of calling activity (excluding growling activity) occurring in sub-alpine habitat.

Montane habitat 0.8 p = 0.003 0.7

0.6

0.5

0.4

0.3 Probability of growling 0.2

0.1 0 5 10 15 20 Mean category temperature (oC)

Fig. 5.15. Relationship between ambient temperature and the probability (with upper and lower bounds) of growling activity occurring in montane habitat.

165 Montane habitat 0.8 p = 0.005 0.7

0.6

0.5

0.4 Probability of growling 0.3

0.2 30 40 50 60 70 80 90 100 Mean category relative humidity (%)

Fig. 5.16. Relationship between relative humidity and the probability (with upper and lower bounds) of growling activity occurring in montane habitat.

166 8 8 8 04:00 8 01:00 02:00 03:00 7 7 7 7 6 6 6 6 5 5 5 5 4 4 4 4 3 3 3 3 2 2 2 2 1 1 1 1 0 0 0 0 -1 -1 -1 -1 -5 0 5 10 15 20 25 -5 0 5 10 15 20 25 -5 0 5 10 15 20 25 -5 0 5 10 15 20 25

8 8 08:00 8 06:00 8 7 05:00 7 07:00 7 7 6 6 6 6 5 5 5 5 4 4 4 4 3 3 3 3 2 2 2 2 1 1 1 1 0 0 0 0 -1 -1 -1 -1 -5 0 5 10 15 20 25 -5 0 5 10 15 20 25 -5 0 5 10 15 20 25 -5 0 5 10 15 20 25 C) o

8 09:00 8 8 8 12:00 7 10:00 11:00 7 7 7 6 6 6 6 5 5 5 5 4 4 4 4 3 3 3 3 2 2 2 2 1 1 1 1 0 0 0 0 -1 -1 -1 -1 -5 0 5 10 15 20 25 -5 0 5 10 15 20 25 -5 0 5 10 15 20 25 -5 0 5 10 15 20 25

8 13:00 8 8 16:00 7 14:00 8 15:00 7 7 6 7 6 6 6 5 5 5 5 4 4 4 4 3 3 3 3 2 2 2 2 1 1 1 1 0 0 0 0 -1 -1 -1 -1 -5 0 5 10 15 20 25 -5 0 5 10 15 20 25 -5 0 5 10 15 20 25 -5 0 5 10 15 20 25 Substratum temperature (

8 17:00 8 18:00 8 19:00 8 20:00 7 7 7 7 6 6 6 6 5 5 5 5 4 4 4 4 3 3 3 3 2 2 2 2 1 1 1 1 0 0 0 0 -1 -1 -1 -1 -5 0 5 10 15 20 25 -5 0 5 10 15 20 25 -5 0 5 10 15 20 25 -5 0 5 10 15 20 25

8 21:00 8 24:00 7 8 22:00 8 23:00 7 6 7 7 6 6 6 5 5 5 5 4 4 4 4 3 3 3 3 2 2 2 2 1 1 1 1 0 0 0 0 -1 -1 -1 -1 -5 0 5 10 15 20 25 -5 0 5 10 15 20 25 -5 0 5 10 15 20 25 -5 0 5 10 15 20 25

Ambient temperature (oC)

Fig. 5.17. Diel relationship between ambient and substratum temperature recorded at a sub-alpine (red regression line) and montane (blue regression line) breeding site, 6 October - 11 November, 1999.

167 2.5 Sub-alpine 2.4 Montane

2.3

2.2

2.1

2.0

1.9

1.8

1.7 Ln substratum temperature +5 (oC) +5 temperature Ln substratum

1.6 1.4 1.6 1.8 2.0 2.2 2.4 2.6 2.8 3.0 3.2 3.4 3.6 Ln ambient temperature +5 (oC)

Fig. 5.18. Relationship between ambient and substratum temperature recorded at 13:00 h at a sub- alpine (hatched line) and montane (solid line) breeding site, 6 October - 11 November 1999.

25 am bient/sub-alpine substratum/sub-alpine 20 am bient/m ontane substratum /m ontane )

15

10

5 Temperature (oC

0

-5 7-Oct-99 9-Oct-99 11-Oct-99 13-Oct-99 15-Oct-99 17-Oct-99 19-Oct-99 21-Oct-99 23-Oct-99 25-Oct-99 27-Oct-99 29-Oct-99 31-Oct-99 2-Nov-99 4-Nov-99 6-Nov-99 8-Nov-99 10-Nov-99

D ate (day/m onth)

Fig. 5.19. Ambient and substratum temperature recorded at hourly intervals from a breeding site located in sub-alpine and montane habitat, 6 October - 11 November, 1999.

168 Rainfall Temperature 450 7.5 400 7.0 350 6.5 300 6.0 250 5.5 200 5.0 150 4.5 100 4.0

Total breeding season rainfall (mm) rainfall season Total breeding 50 20 30 40 50 60 70 3.5

Mean breeding season temperature (oC) temperature season breeding Mean 20 30 40 50 60 70 Breeding season duration (days) Breeding season duration (days) )

2 Solar Radiation Relative Humidity 250 95 240

230 90 220 210 200 85 190 180 80 170 160 75 150 20 30 40 50 60 70 Mean breeding season relative humidity (%) humidity relative season breeding Mean 20 30 40 50 60 70 Breeding season duration (days)

Mean breeding season solar radiation (W/m radiation solar season Mean breeding Breeding season duration (days)

Fig. 5.20. Relationship between breeding-season duration (length of calling-activity period) and mean breeding-season weather in sub-alpine habitat for years 1994 - 1999. 169 Commencement Completion Year Ambient GREY=1999 Year Ambient GREY=1999 25 BLACK=1998 1999 1999 BLACK=1998 BROWN=1997 25 1998 1998 BROWN=1997 20 GREEN=1996 1997 1997 20 1996 GREEN=1996 BLUE=1995 15 1996 BLUE=1995 RED=1994 15 1995 1995 RED=1994 1994 1994 10 10

5 5

0 0 Mean daily ambient temperature (oC) temperature ambient daily Mean -5 (oC) temperature ambient daily Mean -5 -30 -20-10 0 10 20 30 40 50 60 70 80 90 100 -30 -20 -10 0 10 20 30 40 50 60 70 80 90 100 Date Date

Year Substratum 1999 1999 GREY=1999 1998 Year Substratum 8.0 1998 BLACK=1998 8.0 1997 GREY=1999 1997 7.5 7.5 BROWN=1997 1996 1996 BLACK=1998 7.0 7.0 GREEN-1996 1995 BROWN=1997 6.5 1995 6.5 BLUE=1995 1994 6.0 1994 6.0 GREEN=1996 5.5 RED=1994 5.5 BLUE=1995 5.0 5.0 RED=1994 4.5 4.5 4.0 4.0 3.5 3.5 3.0 3.0 2.5 2.5

Mean daily substratum temperature (oC) 2.0 2.0 -30 -20 -10 0 10 20 30 40 50 60 70 80 90 100 -30 -20 -10 0 10 20 30 40 50 60 70 80 90100

Date Mean daily substratum Mean(oC) daily temperature Date

Fig. 5.21. Estimated mean daily ambient and substratum temperature for breeding-season commencement and completion in sub-alpine habitat for years 1994 - 1999, as predicted from regression formulae. Date code: -20 = 10 September, 0 = 30 September, 20 = 20 October, 40 = 9 November, 60 = 29 November, 80 = 19 December, 100 = 8 January.

170 Ambient

C) 9 o

8

7

6

5

Mean daily ambient temperature ( temperature ambient daily Mean 4 1994 1995 1996 1997 1998 1999 Year

C) Substratum o 5.4 5.3 5.2 5.1 5.0 4.9 4.8 4.7

Mean daily substratum temperature ( 4.6 1994 1995 1996 1997 1998 1999 Year

Fig. 5.22. Distribution of regression-derived mean daily ambient and substratum temperature recorded during the calling-activity peak (11 day period) of 1994 - 1999 breeding seasons in sub- alpine habitat. Box plots show median, upper and lower quartiles and interquartile range.

171 1994 1995

140 210 200 130 190 120 180 170 110 160 100 150 140 90 130 80 120 110 70 100 60 90 80 50 70 40 60 50 30 40

Totalrainfall fortnightly (mm) 30 Totalrainfall fortnightly (mm) 20 20 10 10 0 0 1 2 3 4 5 6 7 8 1 2 3 4 5 6 7 8

Date Date

1997 1996 120 140 110 130 100 120 90 110 100 80 90 70 80 60 70 50 60 50 40 40 30 30 20 Total fortnightlyTotal rainfall (mm)

Total fortnightly rainfall (mm) Total fortnightly 20 10 10 0 0 1 2 3 4 5 6 7 8 1 2 3 4 5 6 7 8

Date Date

1999 1998 100 140 90 130 120 80 110 70 100 90 60 80 50 70 60 40 50 30 40 30 20 Totalrainfall fortnightly (mm) Totalrainfall fortnightly (mm) 20 10 10 0 0 1 2 3 4 5 6 7 8 1 2 3 4 5 6 7 8

Date Date

Fig. 5.23. Commencement and completion dates of 1994 - 1999 breeding seasons (horizontal bar with hairs) in sub-alpine habitat in relation to total fortnightly rainfall recorded at Village Flat (1470 m a.s.l.) within the Mt Baw Baw Alpine Resort. Date Code: 1 = 1 – 16 September; 2 = 17 – 30 September; 3 = 1 – 16 October; 4 = 17 – 31 October; 5 = 1 – 16 November; 6 = 17 – 30 November; 7 = 1 – 16 December; 8 = 17 – 31 December. N.B., Rainfall records are absent for the month of September and part of October during the years 1994 and 1995.

172 130 120 110 100 90 calling activity 80

Mean rainfall (mm) 70 60

50 Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec Month

Fig. 5.24. Mean monthly rainfall (± SE) derived from 1932 - 1995 records collected at Erica (40 m a.s.l.), located at the south-eastern end of the Baw Baw Plateau, and the months in which calling activity by Philoria frosti has been recorded.

173

Chapter 6

BREEDING AND POST-BREEDING SEASON PATTERNS OF MOVEMENT, ACTIVITY AND HABITAT USE

6.1 Introduction

Knowledge of the patterns of movement by amphibians is considered a fundamental component of their ecology (Duellman and Trueb 1994), and yet it is not well developed compared to other vertebrates (Whiteman et al. 1996). An understanding of how animals partition their activities and travel among resources is also necessary for species conservation (Pilliod et al. 2002). For amphibians, this might include knowledge of dispersal ability (e.g., Daugherty and Sheldon 1982), the timing of movements (e.g., Paton and Crouch 2002), potential to recolonise disturbed areas (e.g., Kramer et al. 1993), use of aquatic and terrestrial habitat (e.g., Dodd and Cade 1998; Bulger et al. 2003), conditions of environmental stress (e.g., Blaustein and Walls 1995), use of overwintering habitat (e.g., Pilliod et al. 2002), possible gene flow among populations (e.g., Lande 1988) and when and where to release control agents for an introduced species (e.g., Seebacher and Alford 1999).

A recognised difficulty in studying amphibian ecology is that of relocating known individuals. This problem has resulted in many investigations being confined to the breeding season, when amphibians are more concentrated and conspicuous, but not the non-breeding period when they may be involved in other activities in other habitats away from breeding sites (Richards et al. 1994; Madison 1997; Miaud et al. 2000). Studies that have examined amphibian movement following the breeding season have shown that individuals generally disperse into adjacent habitats surrounding the breeding site where they may spend most of their time consuming enough food to grow, preparing for the next breeding season and sheltering from predation, dehydration and freezing (e.g., Denton and Beebee, 1993; Bosman et al. 1996; Madison 1997; Spieler and Linsenmair 1998; Bull and Hayes 2001; Richter et al. 2001). A range of ecological factors are known to influence amphibian movement, including habitat and resources (e.g., Dodd and Cade 1998; Marsh et al. 2000; Pröhl and Berke 2001), demography (Duellman and Trueb 1994), climatic conditions (e.g., Currie and Bellis 1969; Shealy 1975; Kusano and Miyashita 1984; Duellman and Trueb 1994; Palis 1998) and physiological condition (e.g., Licht 1969; Pough et al. 1983).

175 Obtaining information on movement and habitat use by amphibians until recently has been restricted primarily to recapture of animals that have been individually marked (e.g., Currie and Bellis 1969; Kusano and Miyashita 1984; Okuno 1984; Okuno 1986; Staub et al. 1995). However, such procedures often lack the capacity to provide continuous information over time (Duellman and Trueb 1994; Kusano, et al. 1995b; Beebee 1996), particularly when cryptic, hard to capture, or rare species are involved. To obtain continuous information, a number of amphibian studies have used other methods to monitor movement, including radio-active tags (e.g., Kramer 1973; Freda and Gonzalez 1986) and mechanical devices (e.g., Hisai et al. 1987; Sinch 1988). Miniaturisation of electronic components required for radio-tracking has allowed wildlife biologists to remotely monitor many free-ranging animals (White and Garrott 1990). However, for amphibians, such advancements have still confined most studies to large species such as members of the genus Bufo (e.g., Gelder et al. 1986; Denton and Beebee 1993; Kusano, et al. 1995b). More recent advancement in radio-transmitter technology has allowed the development of transmitters with a total mass of less than 1 g for amphibian radio-tracking purposes (Richards et al. 1994), allowing medium-sized amphibians such as P. frosti to be investigated. Radio-tracking remains the only satisfactory approach for obtaining information on animal movement patterns (Koenig et al. 1996).

General descriptions of the habitat of P. frosti have been documented (e.g., Barker et al. 1995; Cogger 2000), however, most records originate from the breeding season from October - December (Atlas of Victorian Wildlife, Dept of Natural Resources and Environment, Victoria) when calling males are more easily located. Malone (1985a) provided the most detailed description of breeding habitat used by P. frosti, with aggregations of calling males being recorded predominantly in unforested, sub-alpine habitats above 1300 m on the Baw Baw Plateau, and a small number detected in habitats with a sparse understorey of Myrtle Beech (Nothofagus cunninghammii) and/or Mountain Tea-tree (Leptospermum grandifolium). Within these habitats, oviposition sites have been recorded predominantly in natural cavities in vegetation, which act as catchments for water travelling down slopes, with others being recorded beneath logs, boulders and rocks (Littlejohn 1963; Malone 1985a, b). To date there are but a few anecdotal records of P. frosti from outside the breeding season. These are from beneath building materials and in disturbed areas within the Mt Baw Baw Alpine Resort prior to 1983, when the species was considered to be relatively common (M. Littlejohn, J. Coventry, P. Robertson unpublished data). It is not known whether P. frosti utilises other habitats in addition to the breeding habitats already described, nor its potential for movement or dispersal within a restricted, high elevation environment.

Due to the endangered conservation status of P. frosti, and its distribution across land tenure managed for different purposes (see Chapter 2), collection of information regarding movement and dispersal, macro and micro-habitat use and general response to climatic conditions was considered important to the development of management measures to ensure the long-term viability of the species. Radio-tracking was considered an appropriate technique to obtain such information, given the rarity of the species, its cryptic nature and complex habitat. A mark-recapture study involving

176 the use of pitfall traps established around a known breeding site was undertaken to supplement the use of radio-tracking in investigating movement and habitat use by P. frosti.

6.2 Materials and Methods

6.2.1 Radio-tracking

Frogs used to examine movement and activity patterns of adult P. frosti were obtained from localities within the Mt Baw Baw Alpine Resort and adjacent Baw Baw National Park. Frog movement was monitored using radio-transmitters weighing 0.9 g with a flexible aluminium antenna (model BD-2 and BD-2T, Holohil Systems Ltd., 3387 Stonecrest Rd, Woodlawn, Ontario, Canada, KOA 3MO) attached to the waist (abdomen) of frogs via a harness of surgical silicon tubing (2 mm diameter) and cotton (Fig. 6.1). To obtain a good fit, different sized harnesses (increments of 1 mm circumference) were placed over the extended rear legs of a frog until a optimum fit was achieved. The use of a harness made of silicon tubing was relatively successful for radio-tracking, although the study was not free of problems. Skin irritations occurred initially on the lateral sides of the abdomen of some individuals, but this was virtually eliminated by modifying the attachment point of the silicon tubing to the radio-transmitter, thereby changing the shape of the harness. The shape of the frogs waist/abdomen was an important aspect to consider when attaching the harness. Male frogs were shaped differently to females, having greater musculature in the hind legs, resulting in fewer transmitters falling off males than females during the study. The lack of easy access to females during both breeding and post-breeding season periods, and the inability to reliably secure radio-transmitters, resulted in the successful telemetry of only a small number of females.

To minimise any effect that transmitter attachment might have on a frog, it was considered appropriate to limit the combined weight of the transmitter and harness to not more than approximately 10 % of a frog’s body weight. Although only arbitrary, this figure was based on females being capable of carrying more than 10 % of body weight in egg mass (unpublished data), and records of other anurans being capable of consuming prey weighing more than 10 % of body weight at one feeding (Duellman and Trueb, 1994). This 10% rule is also recommended by Richards et al. (1994) to minimise impact during radio-tracking. As a result of this size limitation, only adult P. frosti were examined during the study.

Radio-tracking was conducted over three years: November 1995 - April 1996, November 1997 - January 1998, October 1998 - February 1999. These times included the breeding season of P. frosti and a portion of the non-breeding season, or post calling activity period (see Chapter 5). In 1995/96, radio-tracking was conducted in sub-alpine (> 1300 m) and montane (960 - 1300 m) habitat types, whilst in 1997/98 and 1998/99 seasons, radio-tracking was confined to montane

177 elevations, following the discovery that a large portion of the population is confined to this habitat (see Chapter 3).

Frogs were collected mostly during the breeding season when calling males could be located, at which time some females were also located. Some frogs captured in pitfall traps were also used for radio-tracking (see below). Gravid females were not used for radio-tracking because of the potential difficulties in attaching a transmitter and harness and possible adverse affects on reproduction. Because of the cryptic nature of the species, and the reduced population size (see Chapter 3), locating frogs in the absence of calling activity following the breeding season was difficult. An effort was therefore made to locate frogs towards the end of each breeding season for the purpose of radio-tracking individuals during the post-breeding season period. Active searching employed during the post-breeding season period also located a small number of frogs for radio- tracking.

Transmitter frequencies used ranged from 151.049 - 151.561 MHz and frog positions were located by overhead localisation using a Sirtrack H-frame Yagi hand-held antenna (Sirtrack, Havelock, NZ) and a TR-4 Telonics receiver (Telemetry-Electronics Consultants, 932 E. Impala Ave, MESA, Arizona, 85204-6699, USA) (both antenna and receiver were supplied by Faunatech/Austbat, P.O. Box 1655, Bairnsdale, Victoria, AUS).

6.2.1.1 Data Collection

Prior to transmitter attachment, each frog was sexed, measured, weighed and marked by toe- clipping (following Hero 1989). Snout-vent length (SVL) was measured to the nearest millimetre using dial callipers, and weight was measured to the nearest 0.5 g using Pesola spring scales. Each frog was then released at the site of capture. Frogs were radio-tracked diurnally, with a minimum tracking interval between locations (radio-fixes) of one day (approximately 24 h). This radio- tracking regime was chosen because: (1) it was suitable for monitoring frog movement in a sub- alpine, rough, hazardous terrain; (2) it was considered unlikely that frogs would occupy different nocturnal and diurnal sheltering sites on a daily basis; and (3) it was considered suitable for avoiding data autocorrelation problems (Harris et al. 1990) that might be associated with frog movement. On average, radio-tracking intervals were less during the breeding season (mean = 2.1 ± 0.2 days, range = 1 - 13) compared to the post-breeding season period (mean = 3.1 ± 0.2 days, range = 1 - 17). Transmitters were initially left on frogs for a maximum of only four weeks, during which time transmitter and harness attachment techniques were trialed and monitored. Subsequent refinement of the harness attachment method enabled transmitters to be left attached to frogs for close to their total battery life of nine weeks, before which they were removed and replaced.

For each radio-fix, a number of parameters were measured and described: (1) date; (2) time; (3) linear distance moved from last observation; (4) compass bearing from last observation; (5)

178 ecological vegetation class (following Davies et al. 2002) (6) altitudinal environment (sub-alpine and montane); (7) micro-habitat; (8) wet and dry bulb temperature (measured 1 m above sheltering site using a Whirling Sling Psychrometer); and (9) sheltering site temperature - (a) ground surface and (b) substratum (measured with a LCD Insertion Thermometer ± 1.0 o C). In most cases, frogs were located to their exact position by utilising the receiver without the antenna attached. The detection of a signal by the receiver alone was considerably more sensitive than with the antenna attached when in close proximity to the transmitter. When used in association with the gain control, the receiver became sensitive to less than 100 mm distance.

To minimise disturbance, the handling of frogs in most cases was restricted to approximately weekly intervals, at which time the health of individuals was inspected by examining skin condition. Health assessments of frogs that were radio-tracked initially for a period of four weeks showed that body weights were maintained close to the initial capture body weight. It was therefore decided to discontinue the monitoring of body weight due to the additional handling it imposed on the frogs, with general body condition appraisals used to assess health. When frogs were not removed from sheltering sites, substratum temperatures were estimated by measuring temperatures at a distance of 100 mm from the frog, and at a depth of 100 – 150 mm. This depth was considered a reasonable estimate of the sheltering site location based on experience in locating the species (G. Hollis pers. obs.). Surface temperatures were measured approximately 10 mm below the substratum above each sheltering site. A number of meteorological variables, including ambient temperature (oC), relative humidity (%), solar radiation (W/m2) and rainfall (mm), were measured at hourly intervals with a battery-powered data logger (Tain Electronics Pty. Ltd., 10 Rowen Court, Box Hill North, 3129, Victoria) stored in a Stevenson Screen, and located in the Mt Baw Baw Alpine Resort (1470 m a.s.l.).

6.2.2 Pitfall Trapping

A pitfall trap-line was established around the periphery of a known breeding site within Tanjil Plain frost hollow, located within the Baw Baw National Park (see Balkau 1987). The breeding site was along a seepage line beneath several large granite boulders in vegetation comprising sub- alpine wet heathland, and surrounded by sub-alpine woodland. Most of the trap-line was established in sub-alpine woodland, 10 - 15 m from its boundary with sub-alpine wet heathland, and ranging from 15 - 60 m from the breeding site. The remaining portion of the trap-line (approximately 10 m) was within sub-alpine wet heathland. The trap-line consisted of 169 5-litre buckets buried flush with the ground surface and distributed at 2 - 3 m intervals. It was designed and positioned to trap frogs moving to and from the breeding site. To improve capture potential, a continuous 200 mm high drift fence, made of aluminium fly-wire and supported by steel pegs, was constructed between 20 buckets within two portions of the pitfall-trapline, whilst 1.5 m lengths of fly-wire were placed across 90 buckets in areas of more sensitive vegetation. At the remaining 40 buckets, no drift fences were established due to rough terrain, or sensitive vegetation.

179 Breeding and post-breeding season pitfall trapping was conducted at varying intervals during 1994/1995 (November 1994 - February 1995) and 1995/1996 (November 1995 - March 1996) (Fig. 6.17). During exceptionally wet conditions, not all traps were operational due to local drainage effects. Consequently, the total number of traps open each day did not always total 169. Traps were checked at daily intervals during the breeding season. Due to the absence of frog captures during post-breeding season trapping, traps were sometimes checked every second day. Lids were placed on buckets when not in use. Captured P. frosti were individually marked by toe- clipping following Hero (1989) and released at the site of capture.

6.2.3 Breeding and Post-breeding Seasons

Because the period over which radio-tracking was undertaken encompassed the breeding season (spring to early summer) and post-breeding season (summer and autumn), the two seasons were treated as discrete biological units for examining patterns of frog movement and habitat use. Calling activity by P. frosti at a number of breeding sites was monitored in each year to estimate and define the breeding and post-breeding seasons for the general population. The breeding season was considered to encompass the period of calling activity in each year. Daily calling activity was recorded using automatic recording apparatus, and transcribed onto data sheets. The relationship between mean call rates (calls/min) and date, as depicted by distance-weighted least-squares smoothing (tension = 0.5), was used to define commencement and completion of calling activity by the population in each year (see methods and results in Chapter 5).

Because of the possibility that not all individuals participated in movement activity according to season type, as defined above, additional criteria were used to refine the classification of frog movement into breeding and post-breeding season. For males, this included, (1) cessation of calling activity at established breeding sites towards the end of the population breeding season; and (2) a permanent move by an individual into habitat considered unsuitable for breeding activity (i.e., away from seepage habitat). As only non-gravid (spent) females were used in the study, these individuals were considered to have, technically, finished breeding activity, and their movements were subsequently analysed as post-breeding season movement. Season type, as defined for the population, was used to classify movement in the absence of the above information.

6.2.4 Data Analyses

6.2.4.1 Movement

To increase sample size, data recorded during 1995/96, 1997/98 and 1998/99 years were pooled for analysis. Because not all frogs were radio-tracked on a daily basis, frog movement between

180 radio-fixes was first standardised to a 'daily movement' by dividing the distance moved by the number of days between radio-fixes. The distance moved by a frog over these intervals, however, should be interpreted as a minimum linear distance, as it is unlikely that all frogs moved in a straight line between radio-fix localities. To compare rates of movement between seasons (breeding and post-breeding), sex (male and female) and elevation (sub-alpine and montane), the total accumulated distance moved by each frog was divided by the number of days over which radio-tracking was undertaken in each respective season. To exclude any potential bias that may result from frogs that were radio-tracked for a short period during one of the two seasons examined, and thereby being subject to a reduced range of climatic conditions with which to examine movement, frogs radio-tracked for a period less than seven days were excluded from analyses. Due to a lack of independence, movement data from individuals were also not used for more than one season. Due to non-normal data distributions and unequal variances, the nonparametric Mann-Whitney U test was used to compare mean daily movement rates by frogs from each group.

6.2.4.2 Home Range

Analysis of home ranges was restricted by the reliance upon male calling activity as the primary method for capturing frogs. This precluded estimating home range for some frogs over an entire breeding season due to most being captured part way through, or towards the end of the breeding season. Movement data for the greater number of frogs were recorded for a portion of the breeding season followed by varying lengths of time into the post-breeding season. Home ranges were therefore estimated for: (1) frogs for which a sufficient amount of data was collected during the breeding season; and (2) frogs for which a sufficient amount data were recorded during the breeding and post-breeding seasons combined. Because breeding-season movement was likely to be confined to breeding sites for at least a portion of the breeding season, occupation of the initial breeding capture site was used as a base from which a composite of both breeding and post- breeding home range was estimated.

Northing and easting coordinates for each frog were derived from distance and angle information using trigonometry, with the initial breeding capture site as the first set of coordinates. Asymptotic analysis was performed on the data to ensure that sampling provided a sufficient assessment of each individual’s home range (Harris et al. 1990). Frogs were exclude from further analysis if their home range did not stabilise.

As there is no single estimator suitable for all range analyses (Kenward and Hodder 1996), home- range was estimated using two methods: (1) Minimum Convex Polygon, with a kernel fix used for range centre definition, an internationally accepted method used to define home range, particularly for comparisons (see Kenward and Hodder 1996; Burgman and Fox 2003); and (2) Kernel Density Estimator, with an applied smoothing factor (h) derived from Least Squares Cross Validation

181 (LSCV), a robust and flexible nonparametric method that incorporates density contouring, allowing representation of the internal structure of the home range (see Silverman 1986; Worton 1989; Harris et al. 1990; Seaman and Powell 1996). The cross-validated fixed kernel estimator appears to be the better method when compared to the adaptive kernel estimator, producing estimates with very little bias and lower error (Seaman and Powell 1996). In this study, both fixed and adaptive kernel estimates of home range are presented for comparison.

Deriving a satisfactory smoothing factor (h) using LSCV was, however, confounded by the presence of multiple radio-fixes at exactly, or nearly exactly, the same location. A number of ways to deal with this anomaly have been suggested by Seaman and Powell (1997). In this analysis, the procedure of deleting duplicate observations, running LSCV, then using the selected smoothing factor on the complete data set was adopted. However, because of the sedentary nature of most frogs during the breeding season, this resulted in too few radio-fixes for an analysis to be conducted on breeding season locations alone. Kernel estimates were subsequently confined to frogs where observations were collected over both breeding and post-breeding seasons.

Home range estimates are presented as 95% of the utilisation distribution for each frog. Representative examples of the internal structure of frog home ranges are graphically presented as a fixed kernel utilisation distribution depicting 50 - 95% density contours at 5% intervals. Ranges Software (version V, Kenward and Hodder 1996, Institute of Terrestrial Ecology, Furzebrook Research Station, Wareham, Dorset, BH20 5AS, UK) was used to calculate minimum convex polygons whilst KernelHR Software (D. Erran Seaman and Roger A. Powell, 1997, North Carolina State University) was used to estimate fixed and adaptive kernels and provide data for graphical presentation. Graphics were created using ArcInfo version 8.0 and ArcView version 3.2 software (ESRI- Environmental Systems Research Institute, Inc, Redlands, CA, 92373, USA).

6.2.4.3 Habitat and Micro-habitat Use

The characteristics and use of habitat and micro-habitats by P. frosti during the breeding season is reasonably well documented (e.g., Malone 1985a). It was therefore not the purpose of this study to determine if P. frosti utilised habitats and micro-habitats in proportion to that which is available, but rather, if breeding and post-breeding season use was different. Contingency Table Analysis was used to determine if habitat and micro-habitat use by radio-tracked frogs was independent of season type (breeding and post-breeding).

Because different frogs were radio-tracked for different lengths of time, and the time between radio-fixes varied, the frequency at which frogs utilised different habitat and micro-habitats types was first standardised by deriving a proportion of days spent in each type in each season. This proportion was calculated by dividing the number of days a frog spent in each type of habitat by the total number of days it was observed in each season type. Derived proportions were then

182 averaged across all frogs for each type of habitat in each season prior to analysis. For cases when a change in habitat or micro-habitat occurred between two radio-fixes with a time interval greater than one day, the interval time was allocated equally between the two types of habitat or micro- habitat.

The high level of accuracy of radio-fix locations recorded in this study enabled habitat types (Ecological Vegetation Classes) used in the analysis to be modified to include further structural information. This resulted in the addition of habitat ecotones and drainage characteristics for some habitat types. The following habitat types were used: (1) sub-alpine wet heathland (SWH); (2) sub- alpine woodland (SW); (3) montane riparian thicket (MRT), also described as central highlands montane scrub cool temperate rainforest elsewhere in this thesis (see Chapters 2 and 4); (4) MRT/SW ecotone; (5) SWH/MRT ecotone; (6) SWH/SW ecotone; (7) cool temperate mixed forest, drainage line; and (8) cool temperate mixed forest non drainage line.

The micro-habitat of frog sheltering sites obtained at each radio-fix location was categorised into the following combinations of micro-habitat types: (1) roots; (2) roots and soil; (3) rock and soil; (4) log, soil and roots; (5) litter and soil; (6) litter and roots; and (7) surface vegetation. Because of the complexity of micro-habitats available to frogs, described micro-habitats were classified based on the feature dominating the micro-habitat. For each micro-habitat type, the first listed feature was the dominant or primary attribute. Analyses were conducted separately for habitat and micro- habitat at sub-alpine and montane elevations.

6.2.4.4 Activity Patterns

Mean daily temperature, relative humidity, solar radiation and total rainfall were used to assess the relationship between weather patterns and frog movement during the breeding and post-breeding seasons and at sub-alpine and montane elevations. Due to the limited number of female frogs radio-tracked, male and female frogs were pooled for analysis. For radio-fix intervals that were greater than one day, an average of the daily means for each weather variable was derived for the mean day between radio-fixes. Due to the high levels of association between these variables (temperature/solar radiation– Pearson r = 0.63, P < 0.0005; relative humidity/rainfall– Spearman rank rs = 0.54, P < 0.0005; temperature/relative humidity– rs = -0.78, P < 0.0005; temperature/total rainfall– rs = -0.53, P < 0.0005; solar radiation/relative humidity– rs = -0.76, P<0.0005; solar radiation/rainfall– rs = -0.73, P < 0.0005), and subsequent lack of independence, an additional variable called 'mean daily relative humidity/air temperature ratio' (RH/T) was derived to collectively represent the effect of moisture and temperature on frog movement. A high ratio represented very cold, wet conditions and a low ratio, drier, warmer conditions. Separate variables were also retained for analysis because of the subtle differences between them. For example, periods of high temperature are not always associated with low relative humidity levels; periods of high relative humidity are not always associated with periods of high rainfall.

183 Due to the large number of daily movements that were zero in the data, the analysis of movement and weather was conducted in two stages. Firstly, correlation analysis was used to examine the relationship between daily distances moved > 0 m and each variable, with Spearman rank coefficients being derived for non-normally distributed data (look-up tables were used to determine critical values of the Spearman Rank correlation coefficient). Daily movement and RH/T ratio were natural-log transformed to improve linearity. Secondly, univariant logistic regression analysis was used to determine if any of the weather variables adequately predicted ‘movement’ or ‘no movement’ by frogs. Due to the lack of independence between covariates, each was analysed individually as a continuous variable, with movement type (‘movement’ or ‘no movement’) as the dependent variable. The logit within each variable was presumed to be linear. To exclude the occurrence of zero cells, each covariate was collapsed into the following categories for analysis: mean daily temperature (< 0, 0 - 2.9, 3 - 4.9, 5 - 7.9, 8 - 9.9, 10 - 15, > 15 oC); mean daily relative humidity (< 65, 65 - 84.9, 85 - 94.9, ≥ 95 %); mean daily solar radiation (< 150, 150 - 199.9, 200 - 249.9, ≥ 250 W/m2); total daily rainfall ( = 0, 0.1 - 4.9, 5 - 9.9, 10 - 30, > 30 mm); and daily RH/T ratio (< 0, 0.1 - 9.9, 10 - 19.9, 20 - 29.9, ≥ 30). Logits were predicted from category means for each covariate (derived separately for each season and elevation type) and the response variable. Some covariates were natural-log transformed to improve linearity.

Linear regression analysis was further used to examine the relationship between daily distances moved by frogs (> 0 m) and the arrival of frontal weather bearing rainfall (frontal rain), at sub- alpine and montane elevations, and for different seasons. To improve the accuracy of estimated daily movement, the data set containing frog movement was reduced to a subset containing only data that were recorded over consecutive days. Individual daily frog movements were then scored as the number of days since the arrival of frontal rain, with arrival days being scored as day zero. The arrival time of frontal systems was estimated by examining hourly rainfall and temperature records collected from the study area. In most cases, frontal rainfall was relatively easy to distinguish, coinciding with the arrival of rainfall and a decrease in temperature, followed by a period of zero rainfall and increased temperature. Mean daily frog movements derived for each frontal arrival time, for each frog, were used as replicates in the analysis. Both frog movement and frontal arrival time were natural-log transformed to improve normality.

Logistic regression was also used to examine the influence of frontal rainfall on the probability of recording movement by P. frosti. Frontal arrival time was entered into the model as a continuous independent variable and movement type (‘movement’ or ‘no movement’) as the dependent variable. To exclude the occurrence of zero cells, frontal arrival times were collapsed into four categories for analysis (0 - 1, 2 - 3, 4 - 6, > 6 days) and analysed separately for breeding and post- breeding seasons. Logits were predicted from category means and the response variable. Data from sub-alpine and montane elevations were pooled for analysis as separation of these groups resulted in too few frontal arrival categories. Frontal arrival time was natural-log transformed.

184 6.2.4.5 Sheltering Site Selection

Regression analysis was used to examine relationships between substratum, ground surface and ambient temperature recorded at frog sheltering sites during the breeding and post-breeding seasons and at sub-alpine and montane elevations. For significant associations between substratum temperature and ambient temperature, derived regression coefficients were tested to see if they differed significantly from one (Ho:β1 = 1). Regression slopes derived for surface and sheltering site temperature, plotted against ambient temperature, were compared by testing the homogeneity of slopes using ANCOVA, grouped by temperature, season and elevation type. Ambient temperature was entered into the model as the covariate, temperature as the dependent variable and temperature type (surface and substratum), season (breeding and post-breeding) and elevation (sub-alpine and montane) as categorical variables. The significance of the interaction term was used to assess slope homogeneity for the variables examined.

6.3 Results

6.3.1 Radio-tracking

The number of adult P. frosti utilised for radio-tracking purposes over three years was 29, comprising 22 males and seven females (Table 6.1). The small number of females examined in this study precluded most statistical comparisons, with most analyses being descriptive. From the 22 males examined, 21 were observed during the breeding season and 17 post-breeding season. Six females were considered to have been observed during the post-breeding season, given their status as having recently deposited egg masses. One immature female was located at a male calling site during the breeding season. The mean body size of frogs was 44.7 ± 0.5 mm (SVL) and 10.7 ± 0.4 g for males, and 49.6 ± 1.5 mm and 12.5 ± 0.7 g for females. Frogs were radio-tracked for varying lengths of time during the study, averaging 38.2 ± 5.9 days and ranging from 1 - 118. Due to the difficulty in locating and attaching transmitters to female frogs, transmitters were attached to males for longer periods (45.2 ± 7.1, range = 1 - 118 days) than females (16.3 ± 4.7, range = 1 - 29 days) during the study. For females, this statistic is derived from all seven individuals, including the immature individual observed during the breeding season. A total of 397 radio-fixes was obtained from all frogs during the study, including 117 and 240 for males during the breeding and post-breeding season, respectively, and 38 for females during the post-breeding season period. The immature female was radio-tracked for a period of only two days.

185 6.3.2 Breeding and Post-breeding Seasons

Estimated dates for breeding season commencement and completion for each year of the study were: 1995, 26 October - 22 December; 1997, 24 October - 24 November; 1998, 25 September - 4 December (Fig. 6.2). The commencement date of calling activity in 1998 was estimated from the call rate trajectory of the smoothing curve in the absence of data at the beginning of this season. Most of the radio-tracking in this study was undertaken after the peak in calling activity identified in each breeding season (see Chapter 5 for details), during the post-breeding activity period until the end of summer (February), and a small amount during the autumn period prior to winter.

6.3.3 Movement Rate

Daily movement recorded during the breeding and post-breeding season varied. For males, it ranged from 0 - 6 m during the breeding season and 0 - 12 m during the post-breeding season. For females, daily movement during the post-breeding season period ranged from 0 - 7.5 m (Fig. 6.3). The frequency distribution of these movement data was heavily skewed, with 95.5% of daily movements being ≤ 3 m, and 48.6 % of all daily movements being zero.

Mean daily rates of movement for male frogs did not differ significantly between the breeding season (0.26 ± 0.06 m/day, n = 10) and post-breeding season (0.48 ± 0.09 m/day, n = 9) (Mann- Whitney U = 23.0, p = 0.07), although there is an indication that movement may have been less during the breeding season (Fig. 6.4). Examination of the daily movement distributions showed that there was a higher frequency of zero movements recorded during the breeding season (59.0%) than during the post-breeding season (40.8%). As all female movement was categorised as post- breeding season movement, a comparison could be made only with that of males during the post- breeding season period. Female movement did not differ significantly from male movement during the post-breeding season period (Mann-Whitney U = 50.0, p = 0.27), although the mean for females (0.86 ± 0.30 m/day, n = 5) was larger than that for males (Fig. 6.4).

Rates of movement by males did not differ significantly between sub-alpine (0.36 ± 0.10 m/day, n = 3) and montane habitat during the breeding season (0.21 ± 0.07 m/day, n = 7) (Mann-Whitney U = 5.0, p = 0.21; Fig. 6.4). Similarly, during the post-breeding season, rates of movement by males also did not differ significantly between sub-alpine (0.47 ± 0.29 m/day, n = 3) and montane habitat (0.56 ± 0.08 m/day, n = 12) (Mann-Whitney U = 21.0, p = 0.67). In sub-alpine habitat during the post-breeding season, male rates of movement (0.47 ± 0.29 m/day, n = 3) did not differ significantly from that of females (0.86 ± 0.30 m/day, n = 5) (Mann-Whitney U = 11.0, p = 0.30) (Fig. 6.4). No female frogs were radio-tracked in montane habitat.

186 Accumulative and linear distances moved from the point of release varied both within and between breeding and post-breeding seasons. During the breeding season, most males remained within a linear distance of approximately 5 m of initial capture sites (mean = 2.7 ± 0.9 m for frogs with stable home range, n = 6) compared to the post-breeding season when frogs moved linear distances of up to 82 m from initial breeding capture sites (mean = 19.4 ± 6.1 m for frogs with stable home range, n = 12; Fig. 6.5). During the post-breeding season, females remained within 40 m of release sites in sub-alpine habitat (Fig. 6.5), although the sample size was only small (n = 5). The similarity between accumulative and linear distances moved by frogs from site of capture (Figs 6.5 and 6.6) suggests that movement was initially directional, after which there were only small incremental movements (see home range below).

6.3.4 Macro and Micro-habitat Use

Frogs showed preference for different habitat types during the breeding and post-breeding seasons (χ2 = 177.8, df = 7, p < 0.001). Examination of standardised deviates showed that this result was due mainly to the disproportionate use of cool temperate mixed forest, drainage line; cool temperate mixed forest non-drainage line; sub-alpine woodland; sub-alpine wet heathland and sub- alpine wet heathland/sub-alpine woodland ecotone between the two seasons (Fig. 6.7). During the breeding season at sub-alpine elevations, frogs were located only in sub-alpine wet heathland and sub-alpine wet heathland/montane riparian thicket ecotone habitats, whilst at montane elevations, frogs were located predominantly in cool temperate mixed forest, drainage line habitat. During the post-breeding season period at sub-alpine elevations, frogs were located predominantly in sub- alpine wet heathland/montane riparian thicket ecotone; sub-alpine woodland; sub-alpine wet heathland and sub-alpine wet heathland/sub-alpine woodland ecotone, whilst at montane elevations frogs were located mainly in cool temperate mixed forest, non-drainage line habitat.

Micro-habitat use was also dependent on season type, at both sub-alpine (χ2 = 72.6, df = 6, p < 0.001) and montane (χ2 = 101.7, df = 6, p < 0.001) elevations, indicating a preference for different micro-habitat types during the breeding and post-breeding seasons. At sub-alpine elevations, examination of standardised deviates showed that the result was due mainly to the disproportionate use of log, soil and roots, rock and soil and surface vegetation micro-habitats during each season (Fig. 6.7). Rock and soil micro-habitats were more frequently used during the breeding season, whilst logs, soil and roots and surface vegetation were more frequently used during the post- breeding season period. At montane elevations, the result was due mainly to the disproportionate use of surface vegetation, roots, and roots and soil micro-habitats. Surface vegetation and roots micro-habitat were most frequently used during the post-breeding season period, whilst roots and soil micro-habitats were most frequently used during the breeding season.

187 6.3.5 Home Range

Sufficient radio-fixes were obtained to estimate home range for six males during the breeding season and 12 during the breeding and post-breeding season combined. Insufficient data were collected from the small number of females to estimate home range. The results of the asymptotic analysis showed that the number of radio-fixes and days taken for home ranges to stabilise was variable, ranging from 1 - 7 radio-fixes and 1 - 15 days during the breeding season and 5 - 24 radio-fixes and 6 - 49 days for the combined seasons (Table 6.2). During the breeding season, males were confined to smaller home ranges of 0 - 11 m2, mean = 4.8 ± 1.9, (minimum convex polygon) compared to the period encompassing both breeding and post-breeding seasons, when most frogs dispersed into areas adjacent to breeding sites: 3 - 1005 m2, mean = 135.4 ± 80.1, (minimum convex polygon); 4 - 184 m2, mean = 45.7 ± 18.7 (fixed kernel estimate); 9 - 335 m2, mean = 71.6 ± 34.1 (adaptive kernel estimate) (Table 6.2).

Examination of fixed kernel utilisation distributions for frogs radio-tracked over both breeding and post-breeding seasons showed that most participated in a period of dispersal towards the end of the breeding season, after which most remained confined to a relatively centralised home range (see examples in Fig. 6.8a - f). Two frogs (no. 3 and 27), however, remained in close proximity to their initial breeding capture site during the post-breeding season period. Comparison of range estimates derived from each method showed that the minimum convex polygon method produced consistently largest estimates, followed by adaptive and fixed kernel estimates, respectively (Table 6.2).

6.3.6 Weather and Movement

The pattern and extent of association between daily frog movement > 0 m and weather differed for different seasons and elevation types. During the breeding season, distances moved were more associated with weather at montane elevations than at sub-alpine elevations, whilst during the post-breeding season period, distances were more associated with weather at sub-alpine elevations than at montane elevations (Table 6.4). During the breeding season, the pattern of association at sub-alpine elevations (although not significant) was negative with mean daily temperature and solar radiation, and positive with mean daily relative humidity and total daily rainfall (Table 6.4). Contrasting with this pattern was a positive association between mean daily temperature and solar radiation, and a negative association between mean daily relative humidity and total daily rainfall, at montane elevations. During the post-breeding season, the pattern of association between frog movement and weather at sub-alpine elevations followed a similar pattern to that obtained in the breeding season, although all associations were significant. This trend contrasted with patterns of association obtained at montane elevations during the post-breeding period, when all associations were non-significant, and only one variable (mean daily temperature) followed the same trend.

188 Given the patterns of association recorded between movement distances and the temperature and relative humidity variables above, an association between daily RH/T ratio and frog movement for different season and elevation types was also expected. During the breeding season, daily RH/T ratio (natural-log transformed) was positively associated with daily distances moved (natural-log transformed) at sub-alpine elevations and negatively associated at montane elevations, although only the result for montane elevation was significant (Table 6.4, Figs 6.9 and 6.10). During the post-breeding season, daily RH/T ratio was also positively associated with daily distances moved (natural-log transformed) at sub-alpine elevations and negatively associated at montane elevations, except that the result for sub-alpine elevation was only significant in this case. A fitted quadratic curve appears to depict more accurately the relationship between the two variables, compared to the linear relationship (Figs 6.9 and 6.10). At sub-alpine elevations during the post-breeding season, the linear trend line suggests that daily distances moved increased as daily RH/T ratio increased, or as conditions became colder and wetter, whilst the quadratic curve suggests that the relationship is not a simple function, with distances moved decreasing again for the largest RH/T ratios, or coldest, wettest conditions. The largest daily distances moved in this case appear to be undertaken during cool, wet conditions, but not when conditions are very cold and wet. At montane elevations during the breeding season, the quadratic curve for this association suggests a similar relationship to that of the linear trend line for large distances, but that movement reached an asymptote at medium RH/T ratios, or as conditions became moderately cold and wet.

A number of weather variables were suitable predictors of frog movement during the breeding and post-breeding seasons at montane elevations, but not at sub-alpine elevations. During the breeding season, there was an increased probability of movement occurring as mean daily temperature increased and as mean daily relative humidity decreased (Table 6.3, Fig. 6.11). In contrast to this, during the post-breeding season period there was an increased probability of movement occurring as mean daily solar radiation decreased and as mean daily relative humidity (log transformed) and total daily rainfall (log transformed) increased (Fig. 6.12). Mean daily temperature approached significance as a variable for predicting frog movement at montane elevations during the post- breeding season period (p = 0.06).

Daily distance moved by frogs > 0 m was also correlated with the number of days following the arrival of frontal rain during the post-breeding season at sub-alpine and montane elevations (p < 0.0005, r2 = 0.84, n = 13 and p < 0.05, r2 = 0.59, n = 12, respectively) (Fig. 6.13). At both elevations, the magnitude of distances moved by frogs was greater at arrival time compared with subsequent days following arrival. During the breeding season, the relationship between extent of movement and arrival of frontal rain was not significant at sub-alpine and montane elevations (p = 0.22, r2 = 0.45, n = 19 and p = 0.38, r2 = 0.27, n = 14), although the trend was similar to that found during the post-breeding season.

189 The arrival time of frontal rain was also a significant predictor of frog movement during the post- breeding season period (G = 6.431, df = 1, χ2 p < 0.05; Fig. 6.14), but not during the breeding season (G = 0.046, df = 1, χ2 p = 0.83). This trend indicates that during the post-breeding season, there was an increased probability of movement occurring when frontal rain arrived, compared with subsequent days following arrival.

6.3.7 Sheltering Sites

Except for the relationship between substratum and ambient temperatures during the breeding season at sub-alpine elevations, all associations between ambient and sheltering site (surface and substratum) temperatures were significant, with the former relationship approaching significance (p = 0.87) (Table 6.5). Regression coefficients derived for substratum temperature, from its relationship with ambient temperature, were all significantly different from one (β1 ≠ 1), indicating that substratum temperature at frog sheltering sites was different to that of external ambient temperature during breeding and post-breeding seasons and at sub-alpine and montane elevations (Table 6.6). Regression slopes derived for substratum and surface temperature, plotted against ambient temperature, were also significantly different, both within and between season and elevation types (F = 12.17, df = 1, p < 0.005). Figures 6.15 and 6.16 show that frog sheltering sites, as measured by substratum temperature, are buffered against the extremes of high and low external ground surface and ambient temperatures. At montane elevations, both substratum and surface temperatures were slightly higher for all ambient temperatures readings during the post- breeding season compared with the breeding season (i.e., slopes remained similar but y-intercepts were different; Fig. 6.16). At sub-alpine elevations, regression slopes for substratum and surface temperatures during the post-breeding season were steeper than those during the breeding season (Fig. 6.15). Figures 6.15 and 6.16 also show that at sub-alpine elevations, the regression lines for surface and substratum temperatures intersected at ambient temperatures of approximately 6 - 7 oC during the breeding and post-breeding seasons, whilst at montane elevations, intercepts occurred at approximately 3 - 5 oC, respectively. Across the distribution of ambient temperatures, differences in slope and y-intercept of regression lines from the breeding to the post-breeding season indicated a proportional increase in surface and substratum temperature at montane elevation sites, but at sub-alpine elevations, surface and substratum temperatures were both warmer and cooler above and below the temperature associated with the intersection of the regression lines.

6.3.8 Pitfall Trapping

The results of pitfall trapping are summarised in Figure 6.17. Of the 12 P. frosti captured, all were adults, and captures occurred only during the breeding season. During both trapping years, gravid females were captured early in the breeding season (early-mid November), whilst spent females were captured towards the end of the breeding season (mid-late December). One male was also

190 captured early in the breeding season. During both trapping years, three males were heard calling from the known breeding site encompassed by the pitfall trap-line during the breeding season. Trapping effort during the breeding and post-breeding season in 1994/95 was 3460 and 1168 trap- nights, respectively, and 5186 and 6472, respectively, during the post-breeding season in 1995/96.

6.4 Discussion

6.4.1 Study Limitations

It is assumed in this study that neither the weight, size, flexible antenna and methodology of attachment of the transmitter had an adverse effect on frog health and movement. Radio-tracking was conducted on six frogs that weighed less than 10 g, and thus both transmitter and harness together weighed slightly more than the 10% limit adopted initially (up to 11.8% of a frogs body weight). However, movement by these frogs did not appear to be different to that of other frogs. The method of radio-transmitter attachment appears to be the area of most concern when validating the use of external telemetry for amphibians, which have a sensitive moist skin. A variety of external harness systems have been used and tested, each with mixed results (e.g., Rathbun and Murphey 1996; Bull 2000; Goldberg et al. 2002). The influence of an external harness on movement behaviour of an amphibian has also been examined, with impacts to behaviour being detected (Langkilde and Alford 2002). The association of movement with climatic conditions recorded in this study suggest that the use of radio-transmitters, and the method of attachment, had minimal impact on adult P. frosti. It is also likely that the sedentary nature of adult P. frosti contributed to minimising the physical impact of the harness on frogs. However, it is possible that initial movements made by P. frosti shortly after attachment of transmitter and harness may have been as a consequence of handling, or the presence of the transmitter. Although frogs were released at the exact site of capture, most participated in small movements (< 1 m) away from the release site. This behaviour was also recorded by Langkilde and Alford (2002) when tracking Litoria lesueuri. The most limiting aspect of the transmitter attachment method used in this study was considered to be the high percentage (44.8%) that slipped off frogs during radio-tracking. It is likely that the predominantly subterranean nature of P. frosti, coupled with the species reduced waist definition (particularly in females), contributed to this statistic.

Intervals between radio-tracking were larger during the post-breeding season compared with the breeding season. As frogs are unlikely to move in straight lines, daily movements during the post- breeding season are likely to be underestimated when compared with movements from the breeding season. As the minimum radio-tracking interval for frogs in this study was one radio-fix per day, it is also assumed that frogs were not moving from, and back to the same sheltering sites between radio-fixes. Irregular inspections of frog positions suggest that this was not the case. In

191 view of the sedentary nature of P. frosti, the results of this study suggest that a radio-tracking interval of greater than one day may have been satisfactory. Lemckert and Brassil (2000) compared movement data from radio-tracking and spooling methods, and found that radio-tracking may underestimate movement, particularly short-term movements. It is therefore likely that rates of daily movement and accumulated distances moved by P. frosti are underestimated to some extent in this study, but given the sedentary nature of overall movement by the species, this error is unlikely to be great.

Because only adult P. frosti were examined in the study, it is not possible to comment on movement patterns of the species as a whole. Nor is it possible to accurately predict the extent and likelihood of movement during the coldest months of the year (May – September) when radio- tracking was not undertaken, which include the later stages of autumn, winter and the beginning of the breeding season in early spring. However, one could assume that given the high-altitude climate in which the species lives, and that it was found to be relatively sedentary during breeding and post-breeding seasons, movement beyond that which was recorded in this study is unlikely. Few studies have attempted to examine overwintering movement patterns of temperate amphibians when many species are exposed to low ambient temperatures (Lamoureux and Madison 1999). Studies on species of Rana during overwintering periods show that frogs remained active, changing hibernation sites, and often more than once (Holenweg and Reyer 2000; Bull and Hayes 2002). If adult P. frosti are involved in movement during winter, it is likely that it is within the extent of movements recorded during spring, summer and autumn. The ability by P. frosti to locate naturally-occurring insulated refuges that are buffered against temperatures below freezing (see discussion below) suggests that the species may employ a behavioural strategy to avoid freezing conditions as opposed to possessing physiological mechanisms that allow some amphibians to withstand temperatures below freezing (e.g., Storey 1990).

6.4.2 Movement Patterns and Habitat Use

This radio-tracking study, although not comprehensive, has provided an informative account of potential movement patterns and habitat use by adult P. frosti during and following its breeding season. Although only limited data were recorded from pitfall trapping, the results also support the information collected on movement and habitat use through radio-tracking. Due to the cryptic nature and rare status of the species, the collection of data on continuous movement would have been otherwise unobtainable without the use of radio-transmitters.

An examination of specimens of P. frosti from the Museum of Victoria (328 Swanston St, Melbourne, 3000, Victoria) shows that few records of the sub-adult age class have ever been collected, whilst recently metamorphosed frogs and one-year-old individuals have been recorded in the vicinity of breeding sites (Malone 1985b; G. Hollis pers. obs.). The absence of juvenile and sub-adult captures during pitfall trapping suggests that they may remain in close proximity to

192 breeding sites, or move further afield, in this case outside the boundary in which pitfall trapping was conducted. Although most amphibian migrations are associated with reproductive aggregations, dispersal of recently metamorphosed individuals may also involve movement over long distances (Duellman and Trueb 1994). In some species, for example, juveniles have been shown to move distances of up to several hundred metres away from their natal breeding sites (e.g., Jameson 1955, 1956; Zug and Zug 1979), although in some cases, juveniles have also been shown to be more sedentary than adults (e.g., Harris 1975). Juvenile frogs may also stay in close proximity to breeding sites, but also may disperse to more distant places during growth (e.g., Kusano and Miyashita 1984).

Overall, studies on amphibian movement show that most species do not move further from breeding sites than a distance of several hundred metres (see Glandt 1986), with maximum distances of over 1 km being recorded for some genera such as Bufo and Rana (e.g., Lazell et al. 1988; Sinsch 1990; Pilliod et al. 2002). However, distances moved by introduced Bufo marinus in northern Australia may be substantially larger than this (Schwarzkopf and Alford unpublished data in Seebacher and Alford 1999). Unfortunately, there are no studies that document patterns of movement by other species within the genus Philoria with which to compare the movements of P. frosti, nor are there any data from anurans that occupy similar habitats or have a similar life- history strategy. Nevertheless, the extent of movement and home-range estimates derived for individual adult P. frosti suggest that they are relatively sedentary, with all frogs remaining within 82 m linear distance of initial breeding capture sites (mean = 20 m), and combined season home ranges being less than 1005 m2 (mean = 135.4 m2) (MCP) or 184 m2 (mean = 45.7 m2) (fixed kernel). The increased size of home range derived using the MCP method compared with the kernel method is also likely to be biased, particularly as sample sizes increase (Burgman and Fox 2003). The extent of this movement, however, is comparable to other documented cases of amphibian movement (e.g., Nijhuis and Kaplan 1998; Mathews and Pope 1999; Kam and Chen 2000; Lemckert and Brassil 2000; Eggert 2002), and home range (see review by Duellman and Trueb 1994).

Although the range of movement by adult P. frosti was relatively small, there was a consistent pattern of movement and a relatively fixed home range detected for the majority of frogs examined over the duration of breeding and post-breeding seasons. The results suggest that movement between breeding and post-breeding seasons was not random, and that individuals may be able to navigate from breeding sites to feeding and shelter sites, as has been recorded for a number of other amphibians (e.g., Sinsch 1987, 1991). Most males remained in relatively stationary positions for varying portions of time during the breeding season, as indicated by the higher frequency of zero movements, lower rate of movement and small home ranges recorded. This pattern of movement suggests the establishment of a territory by calling adult male P. frosti, as is commonly recorded in many other anurans participating in vocal activity during the breeding season (Duellman and Trueb 1994). However, not all males remained at breeding sites for the entire

193 duration over which calling activity was recorded in each year, with some individuals moving away from breeding habitat before the cessation of calling activity. This finding supports the observation by Malone (1985a) who recorded male P. frosti arriving at, and leaving breeding sites, only to be replaced by other males. By comparison, gravid female P. frosti appear to visit locations at which calling male aggregations have established, and after oviposition, leave the breeding site before the cessation of male calling activity. This visitation period seems to occur over a period of approximately 2 - 4 weeks during the male calling activity peak (as suggested by the 1994 and 1995 pitfall trapping results), and is not dissimilar to the 2 - 3 weeks over which Malone (1985a) recorded oviposition in P. frosti.

Towards the end of the breeding season, most male P. frosti dispersed into habitats adjacent to breeding sites, after which a more stable home range seemed to be established (Fig. 6.8a - f). Dispersal appeared to be directional in that the movement was mostly away from the wetter habitats associated with breeding activity, into relatively drier, adjacent habitats (see discussion below). As a result, home ranges encompassing breeding activity and post-breeding season dispersal tended to be linear in pattern, with an established activity centre for breeding, and feeding and/or shelter following the breeding season. At montane elevations, dispersal was typically perpendicular to the drainage lines associated with breeding habitat, whilst at sub-alpine elevations, it was out of the broader, treeless frost hollows into adjacent treed habitat located up- slope. According to Duellman and Trueb (1994), the orientation of amphibian movements depends on the shape of their local habitat, which appears to be the case for P. frosti. Two radio-tracked males (no. 3 and 27) remained in close proximity to breeding sites during the post-breeding season, suggesting that males may readily move between, or occupy both habitat types following the breeding season. The limited data collected for adult female P. frosti suggest that they also disperse into similar habitat to that of males after oviposition. Seasonal migrations have also been commonly observed in other amphibians (e.g., Kusano et al. 1995b; Madison and Farrand III 1998; Sinsch 1988; Mathews and Pope 1999), particularly those located at temperate latitudes (Stebbins and Cohen 1995).

Movement from ponds, or aquatic habitats, into terrestrial habitats following the breeding season is common in many amphibians (Glandt 1986; Duellman and Trueb 1994). Adult P. frosti also appear to be involved in this pattern of movement. At both sub-alpine and montane elevations, there was a clear preference for wetter habitats in the breeding season and drier habitats following the breeding season, although some frogs used both wet and dry habitats following the breeding season. At sub-alpine elevations (> 1300 m), use of wet habitat (sub-alpine wet heathland and montane riparian thicket) and dry habitat (sub-alpine woodland) is associated with a clear change in vegetation floristics and structure. At lower, montane elevations (960 - 1300 m), the distinction between wet and drier habitat is not as obvious as at sub-alpine elevations, with the major difference being the presence or absence of seepages and soaks. There are, however, major floristic and structural differences between habitats at sub-alpine and montane elevation (see

194 Chapter 4 for details). For male anurans, the home range usually encompasses a suitable calling site, one or more feeding sites and preferred habitat for shelter (Duellman and Trueb 1994). This pattern suggests that movement by P. frosti away from wetter breeding habitat following the breeding season is to occupy suitable habitat for activities such as feeding and shelter. Such movement patterns are also observed in other species, such as Rana clamitans, where movement into terrestrial habitats corresponded with a greater quantity of food compared to aquatic habitats used for breeding (Lamoureux et al. 2002).

Preference by P. frosti for different micro-habitats at sub-alpine and montane elevation during the breeding and post-breeding season periods is analogous to its preference for different habitat types during the same periods and associated change in activity pattern, as described above. During the breeding season, preference for rock and soil micro-habitat at sub-alpine elevations, and roots and soil micro-habitats at montane elevations, corresponds to the greater proportion of frogs that were recorded in wet habitats (sub-alpine wet heathland and montane riparian thicket, and cool temperate mixed forest-drainage line, respectively), and that male calling and/or oviposition sites at each elevation comprise predominantly of these substrata. Similarly, following the breeding season, preference for logs, roots, soil and surface vegetation micro-habitats at sub-alpine elevations, and surface vegetation and roots micro-habitat at montane elevations corresponds to the greater proportion of frogs recorded in drier habitats (sub-alpine woodland and sub-alpine wet heathland/sub-alpine woodland ecotone, and cool temperate mixed forest, non-drainage line, respectively), and that feeding and sheltering are likely to be the primary activities at this time.

Protection from extreme temperatures and dehydration is considered to be the driving mechanism behind selection of sheltering sites in amphibians (e.g., Denton and Beebee 1993; Cohen and Alford 1996; Schwarzkopf and Alford 1996). The use of surface vegetation (e.g., Wittsteinia vaccineaaceae and Blechnum wattsii) as a micro-habitat during the post-breeding season period suggests it may be used for a specific purpose such as feeding or dispersal. Both W. vaccineaaceae and B. wattsii are species that were frequently used, and are structured such that they allow for movement beneath them, whilst providing for a moist, cool micro-habitat. On a number of occasions, individuals were observed moving through or on the surface of vegetation immediately following the arrival of rainfall. Logs, roots and soil are likely to be associated more with sheltering sites where they probably provide for greater protection against desiccation and predation than the surface vegetation micro-habitat. The mainland tiger snake (Notechis scutatus) is relatively common in mid to low-elevation habitats occupied by P. frosti (G. Hollis pers. obs), and may be an important predator of P. frosti. Several individual P. frosti were discovered in the disgorged stomach contents of N. scutatus (Hoplocephalus curtus in Spencer 1901). In this study, an individual P. frosti (no. 20) was also consumed by N. scutatus, whilst a second frog (no. 4) was also excavated from its calling site by an unknown predator (possibly a fox or cat). The use of log, root and soil micro-habitats is likely to provide for greater protection against snakes than surface vegetation micro-habitats. Madison (1997) speculated that the Spotted Salamander (Ambystoma

195 maculatum) reduced its level of movement during spring due to an increased risk of predation by snakes and birds, and that a thicker leaf layer and reduced predator activity allowed for greater movement by the species in the autumn. It is not known whether P. frosti regulates its activity due to predation risk.

6.4.3 Weather Associations

The relationship between weather and the probability and extent of movement by P. frosti indicates that particular weather conditions influence overall movement by the species, but this may vary between season and elevation. The arrival of frontal rainfall resulted in greater movement distances at both elevations and during both seasons, but was a significant predictor of movement only during the post-breeding season period. Rainfall is frequently noted for its importance during movement and dispersal in other amphibians studies (e.g., Currie and Bellis 1969; Kusano and Miyashita 1984; Sinsch 1992; Madison 1997; Palis 1998; Mazerolle 2001), whilst availability of moist habitats is considered to be a particularly important resource for terrestrial amphibians (Ovaska 1991; von Sacken 1998). The lack of significance for frontal arrival time as a predictor of movement during the breeding season could be due to frogs being involved primarily with breeding activity and remaining sedentary. Occupying wet habitats during the spring breeding season, when climatic conditions are wetter, may also minimise potential hydrological constraints that exist during the summer months following the breeding season when temperatures are warmer and humidity lower, and frogs are utilising drier habitat types.

At montane elevations during the breeding season, P. frosti moved greater distances, and there was a greater probability of movement occurring as humidity levels decreased. This, however, is in contrast to the negative association recorded between distances moved by P. frosti and frontal arrival time, when one would expect humidity levels to be high at the time of rainfall arrival. A possible explanation is that prior to the arrival of frontal rainfall, ambient temperatures would be warmer and humidity lower compared to colder temperatures and high humidity levels that would be present after it has moved through. The temperature of sheltering sites (substratum temperature) is also unlikely to change immediately as the frontal rainfall arrives. Substratum temperature was shown to be buffered against external air temperature, such that frogs would remain warmer for longer periods after the passing of a frontal system. This thermal buffering would effectively result in the presence of relatively warm, wet conditions for a period following the arrival of a frontal system. A similar observation was made by Currie and Bellis (1969) who noted a tendency for Rana catesbeiana to participate in movements over land on warm rainy nights compared to cold rainy nights. The non-linear relationship between movement distance and RH/T ratio supports this explanation to some extent in that a minimum threshold of movement distance is reached at a ratio of approximately three when conditions are relatively cool and wet, below which there was no further reduction in movement distance as conditions become wetter and colder (Fig. 6.10).

196 At sub-alpine elevations, although levels of association for temperature, relative humidity, solar radiation, rainfall and RH/T ratio during the breeding season were either poor or not significant, the overall results for both seasons suggests that increased temperature and reduced humidity may limit the extent to which P. frosti can move. Examination of temperature and relative humidity as a combined variable (RH/T ratio) suggests that the observed relationship may not be linear, and that extent of movement may also be limited during very cold, wet conditions (Ln RH/T ratios > 3). No explanation can be given as to why the above variables were not found to be suitable predictors of movement at sub-alpine elevations during both breeding and post-breeding seasons as they were at montane elevations. Nor can an explanation be given for the non-significant correlations recorded during the breeding season. A smaller sample size than that recorded from montane elevations may have obscured any relationship or association, or perhaps the desire by frogs to breed during the breeding season may have nullified or obscured any relationship between weather and movement.

The results recorded from montane elevations are mostly in contrast to that observed from sub- alpine elevations, indicating that reduced temperature and increased relative humidity may reduce distances moved in both seasons, although this was apparent only during the breeding season. No plausible explanation can be given for the non-significant correlations recorded during the post- breeding season. Lack of association might be due to the reduced limitation on movement through the presence of more uniform temperature and humidity levels (see details below). However, if this were the case, then a significant association would not have been recorded from montane elevations during the breeding season when frogs are more sedentary. During the breeding season, temperature and humidity influenced both the probability of movement occurring and distances moved in a similar way (increased probability of movement as temperature increased and humidity decreased), but during the post-breeding season, the results suggest an opposite trend, with an increased probability of movement occurring as humidity and total daily rainfall increased and as solar radiation decreased (Fig. 6.12). Colder, wetter conditions appear to reduce the probability of movement occurring during the breeding season (spring - early summer), although this could also be due to participation in reproductive activities. Warmer, drier conditions appear to reduce the probability of movement occurring during the post-breeding season period (summer - autumn), but when movement does occur, distances are enhanced by increased temperature and reduced humidity.

A noteworthy point regarding the analysis between weather and movement is that all frog movements were correlated against weather data collected from a site at sub-alpine elevation (1470 m a.s.l.), and subsequent differences in climate between sub-alpine and montane habitats could account for the different patterns of association recorded. A closer examination of temperature conditions present at a sub-alpine (1485 m) and montane (1010 m) breeding site during the breeding season showed that although the diel pattern of temperature was almost identical between the two, average temperature was significantly warmer at sub-alpine elevation than at montane elevation, as well as the sub-alpine breeding site being subject to higher and lower

197 thermal maxima and minima (see Chapter 5). A combination of increased average temperature and higher thermal maxima, and associated periods of lower humidity, may possibly limit movement by P. frosti at sub-alpine elevations, whilst colder average temperatures may sometimes be the limiting factor at montane elevations. If movement by P. frosti is limited by warmer temperature and low humidity at sub-alpine elevations, and cooler temperature and high humidity at montane elevations, then overall movement by the species appears to be restricted to within a relatively narrow range of climatic conditions.

Differences between habitat structure and floristics at sub-alpine and montane appear to account for some of the observed movement patterns in conjunction with weather. At sub-alpine elevations, heath and woodland habitats are more exposed to incident solar radiation, and subject to warmer and colder temperature extremes, compared to cool temperate mixed forest and montane riparian thicket at montane elevations, which comprise dense rainforest (Nothofagus cunninghamii, Atherosperma moschatum and Leptospermum grandifolium) and tall eucalypt (Eucalyptus delegatensis, E. nitens, E. glaucescens and E. regnans) canopies (see Chapters 3 and 5 for further details).

The range of temperatures associated with sheltering sites selected by P. frosti suggests that, over and above the compulsion to participate in seasonal activities such as breeding and foraging, the species is selecting sheltering sites that are within a relatively narrow temperature range compared to prevailing ambient temperature. That this temperature range differed between season and elevation type to varying extents can be attributed to changes in climate from spring - summer - autumn, and differences in macro and micro-habitat associated with vegetation at sub-alpine and montane elevations. The data recorded from frog sheltering sites at sub-alpine and montane elevations also support patterns and relationships between frog movement and weather at different elevation discussed above. At montane elevations, for a given ambient temperature, the slight increase in sheltering site temperatures from breeding to post-breeding season can simply be attributed to overall increase in seasonal temperature (i.e., from spring to summer). However, at sub-alpine elevations, the lack of a relatively uniform response to an increase in seasonal temperature from the breeding season to post-breeding season did not occur as clearly as at montane elevations. For ambient temperatures above the point at which the regression lines for surface and substratum temperature intersected (approximately 6 - 7 oC), there was a considerable increase in the temperature of sheltering sites from the breeding season to post-breeding season, below which there was a relative decrease in the temperature of sheltering sites (Fig. 6.15). A possible explanation for this observation could be a combination of macro and micro-habitat use and seasonal changes in temperature. Increases in the temperature of sheltering sites above 6 - 7 oC may have been due to a combination of seasonal temperature increases and the prevalence of greater thermal maxima available at sub-alpine elevations resulting from a more sparse habitat structure, whilst decreases below 6 - 7 oC may have been due to a combination of greater thermal minima that prevail at sub-alpine elevations, a more sparse habitat structure, and a shift from

198 mostly aquatic habitat used for breeding purposes to non-aquatic habitat following the breeding season. The thermal inertia of water, compared to other substrata, may allow frogs to select sites that are more buffered against thermal minima during the breeding season compared to the post- breeding season.

Ambient temperature ranges derived from the intersection of surface and substratum regression lines for sub-alpine (6 - 7 oC )and montane (3 - 5 oC) frog locations appear to be of biological significance. The derived values may represent the temperature range at which frogs are likely to participate in movement, notwithstanding other climatic and environmental factors. Examination of the temperature conditions prevailing on days when frogs at sub-alpine elevations are most likely to move (i.e., days in which frontal rainfall arrived) showed that they are generally within the range of values derived from the regression line intercepts for breeding and post-breeding seasons (mean = 7.3 ± 1.3 oC breeding season; mean = 7.0 ± 1.0 oC post-breeding season) at sub- alpine elevations. ______This chapter has provided insight into the breeding and post-breeding patterns of movement, habitat use and activity of adult P. frosti. Radio-tracking, and the method of transmitter attachment, appear to have had minimal impact on the normal activity patterns of the species, and therefore have provided meaningful data on movement that would have otherwise been unobtainable to due the cryptic nature and rare status of the species. I consider that recorded daily movement and accumulated distances moved by adult P. frosti to be a realistic estimate of the overall movement potential, but may be underestimated to a small extent due to some short-term movements that may not have been detected. Patterns of activity and habitat use recorded during pitfall trapping also support information collected through radio-tracking.

Radio-tracking showed that adult P. frosti are relatively sedentary during breeding and post- breeding seasons, but demonstrate a consistent pattern of movement where frogs disperse from breeding habitats into adjacent terrestrial habitats during and following the breeding season. The limited data collected on female frogs suggests dispersal into terrestrial habitats occurs immediately following oviposition, whilst males occupy breeding sites for a longer period of the breeding season before dispersing into adjacent terrestrial habitats. Movement activities were influenced and limited by climatic conditions in sub-alpine and montane habitats. Sheltering sites chosen by frogs provided micro-climatic conditions that were buffered considerably from external climatic conditions. Due to their different physiography, sub-alpine and montane habitats appear to impose different climatic constraints on movement potential for P. frosti, with overall capacity for movement being confined to a narrow range of macro and micro-climatic conditions. The use of aquatic and terrestrial habitats during the breeding and post-breeding seasons, respectively, coupled with constraints to movement imposed by climate, has important implications when considering the future conservation and management of P. frosti.

199 Table 6.1. Location, sex, size and radio-tracking details for individual frogs radio-tracked between November 1995 and February 1999.

Frog Habitat Type Capture Site Mass (g) SVL Sex Date of Date of No. Elevation (m) (mm) Transmitter Last Attachment Radio-fix 1 Sub-alpine 1485 13.5 46.5 m 14-Nov-95 16-Jan-96 2 Sub-alpine 1375 11.5 43.9 m 15-Nov-95 16-Nov-95 3 Sub-alpine 1505 12.0 49.5 m 17-Nov-95 28-Feb-96 4 Montane 1165 11.5 43.9 m 20-Nov-95 11-Dec-95 5 Sub-alpine 1440 10.5 48.2 m 28-Nov-95 26-Mar-96 6 Sub-alpine 1385 8.5 44.3 m 1-Dec-95 11-Dec-95 7 Sub-alpine 1390 10.0 45.4 m 4-Dec-95 10-Jan-96 8 Montane 1250 12.5 46.2 m 25-Feb-96 26-Mar-96 9 Sub-alpine 1505 14.5 52.4 f 19-Nov-95 17-Dec-95 10 Sub-alpine 1440 12.0 50.2 f 28-Nov-95 22-Dec-95 11 Sub-alpine 1380 11.0 49.9 f 4-Dec-95 11-Dec-95 12 Sub-alpine 1510 14.5 54.9 f 5-Dec-95 6-Dec-95 13 Sub-alpine 1505 13.5 50.6 f 18-Dec-95 10-Jan-96 14 Sub-alpine 1375 11.5 45.7 f 19-Mar-96 19-Apr-96 15 Montane 1200 9.5 44.5 m 24-Nov-97 1-Jan-98 16 Montane 1250 11.0 44.5 m 23-Nov-98 16-Feb-99 17 Montane 1215 10.0 43.5 f 24-Nov-97 26-Nov-97 18 Montane 1210 10.0 43.5 m 22-Nov-97 29-Nov-97 19 Montane 1215 8.5 41.1 m 22-Nov-97 24-Nov-97 20 Montane 1220 11.5 47.0 m 18-Oct-98 7-Jan-99 21 Montane 1220 8.5 40.5 m 24-Nov-97 12-Jan-98 22 Montane 1220 8.5 42.5 m 24-Nov-97 21-Jan-98 23 Montane 1240 11.5 46.3 m 6-Nov-98 16-Dec-98 24 Montane 1250 10.0 44.6 m 25-Oct-98 4-Dec-98 25 Montane 1240 8.7 40.7 m 16-Nov-98 23-Nov-98 26 Montane 1240 14.0 48.8 m 17-Oct-98 29-Dec-98 27 Montane 1090 11.0 40.9 m 2-Nov-98 7-Dec-98 28 Montane 1215 11.7 45.6 m 18-Oct-98 2-Jan-99 29 Montane 1250 10.0 43.8 m 23-Nov-98 16-Dec-98

200 Table 6.2. Movement rate, radio-tracking intensity and results of home range and asymptotic analyses for individual frogs radio-tracked between November 1995 and February 1999.

BS = breeding season; P-bS = post-breeding season; CS = breeding and post-breeding season combined; - denotes individuals omitted from analysis due to insufficient data, * denotes individuals for which fixed and adaptive kernel estimates could not be derived (see methods). Home range estimates represent 95% of the utilisation distribution.

Frog No. No. of Radio-fixes/Days Movement Rate No. of Radio-fixes/Days 95% Home Range (m2) 95% Home Range (m2) Radio-tracked (m/day) Required to Stabilise Minimum Convex Polygon Fixed Kernel/Adaptive Kernel BS P-bS CS BS P-bS BS CS BS CS BS CS 1 6/6 23/5729/63 - 0.42 - 24/37- 65 17/23 2 1/1 0/0 1/1 ------3 21/36 13/65 34/101 0.16 0.01 1/1 22/35 4 5 * * 4 11/21 0/0 11/21 0.00 - 1/1 - 0 - * - 5 3/6 35/112 38/118 - 0.99 - 18/49 - 1005 184/335 6 2/10 0/0 2/10 0.50 ------7 9/31 3/6 12/37 0.41 ------8 0/0 8/30 8/30 - 0.53 ------9 0/0 14/28 14/28 - 0.28 ------10 0/0 10/24 10/24 - 0.08 ------11 0/0 2/7 2/7 - 1.07 ------12 0/0 1/1 1/1 ------13 0/0 8/23 8/23 - 1.67 ------14 0/0 3/29 3/29 - 1.21 ------15 1/1 15/37 16/38 - 0.67 - 14/23 - 91 - 65/80 16 2/3 24/82 26/85 - 0.22 - 13/32 - 28 - 8/9 17 2/2 0/0 2/2 ------18 4/4 1/3 5/7 ------19 2/2 0/0 2/2 ------20 6/15 23/66 29/81 0.48 1.02 2/4 13/29 11 153 * 62/77 21 2/2 15/47 17/49 - 0.58 - 15/29 - 77 - 31/42 22 2/2 17/56 19/58 - 0.47 - 5/6 - 43 - 4/9 23 8/26 5/14 13/40 0.24 0.93 4/10 - 10 - * - 24 13/32 3/8 16/40 0.09 0.67 1/4 14/34 1 16 * * 25 2/7 0/0 2/7 0.44 ------26 13/30 14/43 27/73 0.14 0.86 7/15 18/44 3 109 * 25/49 27 3/4 11/31 14/35 - 0.06 - 5/8 - 3 - * 28 2/6 25/70 27/76 - 0.34 - 5/11 - 30 - 15/20 29 4/9 5/14 9/23 0.10 0.36 ------Mean+SE 4.8 + 1.9 135.4 + 80.1 45.7 + 18.7/71.6 + 34.1

201 Table 6.3. Results of univariant logistic regression analysis on weather variables (collapsed into categories) and movement (Ref' = reference [no movement], Resp = response [movement]) during the breeding and post-breeding seasons, and at sub-alpine and montane elevations.

All variables were modelled with a constant. Est = estimated coefficient, SE =standard error, t ratio = standardised coefficient, G = likelihood ratio statistic, bracketed values = constant. # denotes independent variable natural log transformed, * denotes significant p values.

Weather variable (by category) Season Ref/Resp Est se t ratio G df χ2p value Sub-alpine Elevation Mean daily temperature Breeding 20/22 0.019(-0.194) 0.109(0.639) 0.177(-0.304) 0.031 1 0.860 Post-breeding 50/62 0.040(-0.485) 0.056(0.425) 0.712(-1.140) 0.510 1 0.475 Mean daily solar radiation Breeding 20/22 0.001(-0.349) 0.003(0.739) 0.378(-0.472) 0.144 1 0.705 Post-breeding 50/62 -0.000(-0.174) 0.002(0.500) -0.090(-0.347) 0.008 1 0.928 Mean daily relative humidity Breeding 20/22 -0.041(3.671) 0.033(3.008) -1.261(1.221) 1.694 1 0.193 Post-breeding 50/62 -0.008(0.500) 0.014(1.292) -0.560(0.387) 0.313 1 0.576 Total daily rainfall Breeding 20/22 0.004(-0.144) 0.015(0.353) 0.287(-0.409) 0.083 1 0.774 Post-breeding 50/62 0.004(-0.249) 0.011(0.212) 0.364(-1.175) 0.132 1 0.716 RH/T ratio Breeding 20/22 -0.020(0.462) 0.017(0.565) -1.175(0.818) 1.411 1 0.235 Post-breeding 50/62 -0.013(0.0770 0.010(0.287) -1.337(0.269) 1.841 1 0.175 Montane Elevation Mean daily temperature Breeding 28/47 0.107(-1.393) 0.048(0.472) 2.240(-2.948) 5.383 1 *0.020 Post-breeding 104/62 -0.078(1.178) 0.042(0.396) -1.857(2.975) 3.519 1 0.061 Mean daily solar radiation Breeding 28/47 0.005(-1.755) 0.003(0.861) 1.523(-2.039) 2.426 1 0.119 Post-breeding 104/62 -0.004(1.596) 0.002(0.569) -2.007(2.804) 4.165 1 *0.041 Mean daily relative humidity Breeding 28/47 -0.023(1.161) 0.011(0.845) -2.046(1.373) 4.301 1 *0.038 Post-breeding# 104/62 2.069(-8.578) 0.902(3.967) 2.294(-2.162) 5.492 1 *0 019 Total daily rainfall Breeding 28/47 0.010(-0.556) 0.045(0.295) 0.223(-1.887) 0.050 1 0.824 Post-breeding# 104/62 0.667(-0.244) 0.200(0.271) 3.340(-0.901 12.306 1 *0.000 RH/T ratio Breeding 28/47 -0.003(-0.409) 0.003(0.261) -0.967(-1.571) 1.010 1 0.315 Post-breeding 104/62 -0.000(0.523) 0.003(0.178) -0.080(2.946) 0.006 1 0.936

202 Table 6.4. Results of correlation analysis (Pearson-r and Spearman rank-rs) between weather variables and daily frog movement > 0 m (natural log transformed).

'n.s' denotes non significant p value (α = 0.05).

Weather variables Breeding Season Post-breeding Season Sub-alpine Montane Sub-alpine Montane Mean Daily Temperature r = -0.20, n.s r = 0.38, p < 0.05 r = -0.53, p < 0.001 r = 0.05, n.s Mean Daily Solar Radiation r = -0.25, n.s r = 0.45, p < 0.05 r = -0.66, p < 0.001 r = -0.12, n.s Mean Daily Relative Humidity rs = 0.05, n.s rs = -0.46, p < 0.05 rs = 0.50, p < 0.001 rs = 0.001, n.s Total Daily Rainfall rs = 0.28, n.s rs = -0.31, n.s rs = 0.42, p < 0.01 rs = 0.09, n.s Mean Daily Ln RH/T Ratio r = 0.22, n.s r = -0.42, P < 0.05 r = 0.42, p < 0.01 r = -0.06, n.s

Table 6.5. Results of regression analysis between ambient temperature and temperature of frog sheltering sites (surface and substratum) during the breeding and post-breeding seasons and at sub- alpine and montane elevations.

Elevation Season Variable Coefficient SE r2 n p

Sub-alpine Breeding Substratum 1.21 0.6 0.27 41 =0.087 Breeding Surface 1.12 0.2 0.53 41 <0.0005 Post-breeding Substratum 1.91 0.2 0.58 104 <0.0005 Post-breeding Surface 1.34 0.0 0.83 106 <0.0005

Montane Breeding Substratum 1.73 0.2 0.63 77 <0.0005 Breeding Surface 1.15 0.1 0.78 77 <0.0005 Post-breeding Substratum 1.61 0.1 0.58 162 <0.0005 Post-breeding Surface 1.12 0.0 0.85 163 <0.0005

Table 6.6. Results of hypothesis test (Ho:β1=1) on regression coefficients derived for the relationship between ambient and substratum temperature recorded at frog sheltering sites during the breeding and post-breeding seasons and at sub-alpine and montane elevations.

Testing was only conducted on significant associations.

Elevation Season Variable df f p

Sub-alpine Post-breeding Substratum Temperature 1 52.04 <0.0005

Montane Breeding Substratum Temperature 1 49.91 <0.0005 Post-breeding Substratum Temperature 1 81.12 <0.0005

203 Fig. 6.1. Adult Philoria frosti with radio-transmitter and harness (drawing: Kate Thompson).

Year

Red=1995 35 Brown=1997 30 Black=1998 25

20

15

10

Mean number of calls/minute 5

0 20 30 10 20 30 9 19 29 9 19 29 Sep Sep Oct Oct Oct Nov Nov Nov Dec Dec Dec Date (day/month)

Fig. 6.2. Duration of male calling activity recorded at breeding sites during 1995, 1997 and 1998 breeding seasons, as depicted by least squares smoothing curves.

204 120 1.0 110 Breeding Season 100 Males (n=117) 0.8

90 P

r o

80 p

o

r t

t 70 0.6

i o

n n

u 60

o p

C

50 e r

0.4 B

40 a r 30 20 0.2 10 0 0.0 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 Movement (m)

200 0.8 Post-breeding Season 0.7 Males (n=240) 150 P

0.6 r

o

p o

0.5 r t

t

i o

n n

u

100

o

0.4 p

C

e

r

0.3 B

a r 50 0.2

0.1

0 0.0 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 Movement (m)

40 1.0 Post-breeding Season Females (n=38)

30 0.8 P

r

o

p

o

r t

t

0.6 i o

n n

u

20

o p

C

e

r

0.4 B a 10 r 0.2

0 0.0 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 Movement (m)

Fig. 6.3. Frequency distribution of daily movement recorded for males during the breeding season, and for males and females during the post-breeding season, over the duration of the study.

205 Sex Male 2.0 Female

1.5

1.0

0.5 Rate of movement (m/day) of movement Rate

0.0 Breeding Post-breeding Season

Season 2.0 Breeding Post-breeding

1.5

1.0

0.5 Rate movement of (m/day)

0.0 montane sub-alpine Habitat type

Sex Males 2.0 Females

1.5

1.0

0.5 Rate of movement (m/day)

0.0 sub-alpine Habitat

Fig. 6.4. Distribution of male and female rates of movement recorded during the breeding and post-breeding seasons, and at sub-alpine and montane elevations. Box plots show median, upper and lower quartiles, interquartile range and outside values.

206 Breeding Season

10 Males 9 8 7 6 5 4 3 2 1 0

Linear distance from site of capture (metres) 0 10 20 30 40 Day since transmitter attachment

Post-breeding Season Males 100.0

10.0

1.0

Linear distance site of capture from (metres) 0 50 100 150 Day since transmitter attachment

Post-breeding Season

Females 40

30

20

10

0

Linear distance site of capture from (metres) 0 10 20 30 Day since transmitter attachment

Fig. 6.5. Linear distance moved from site of capture over time by male and female frogs during the breeding and post-breeding seasons. N.B., logged Y axis scale for males during the post-breeding season.

207 Breeding Season Males 15

10

5 ccumulative distancemoved A from site of capture (metres) capture of site from 0 0 10 20 30 40 Day since transmitter attachment

Post-breeding Season Males

1100.0

10.0

ccumulated distance from distance ccumulated 1.0 A site of capture (metres) capture of site

0 50 100 150 Day since transmitter attachment

Post-breeding Season

Females 40

30

20

10

0 0 10 20 30 Day since transmitter attachment Accumulative distance moved from moved distance Accumulative site of capture (metres) Fig. 6.6. Accumulated distance moved from site of capture over time by male and female frogs during the breeding and post-breeding seasons. N.B., logged Y axis scale for males during the post-breeding season.

208 100

) Season Type

%

( 80

s post-breeding

y

a Breeding

d

f

o 60

n

o

i

t

r

o

p 40

o

r

p

n

a

e 20

M

0 l l t h t fd nd r sw sw w r sw m f m t/ s /m h/ ct tm r h w c m sw s Habitat type

Sub-alpine Season Type Elevation Post-breeding 100 Breeding

80

60

40

20 Mean proportion of days (%) days Mean proportion of

0 LS LSR LV RS SV VR VRS Micro-habitat type

Montane Season Type Elevation Post-breeding 100 Breeding

80

60

40

20 Mean proportion of days (%) days of proportion Mean

0 LS LSR LV RS SV R RS Micro-habitat type

Fig. 6.7. Mean proportion of days spent by frogs in different macro and micro-habitat types during the breeding and post-breeding seasons, at sub-alpine and montane elevations. Ctmfdl = cool temperate mixed forest, drainage line; ctmfndl = cool temperate mixed forest, non drainage line; mrt = montane riparian thicket; mrt/sw = montane riparian thicket/sub- alpine woodland ecotone; sw = sub-alpine woodland; swh = sub-alpine wet heathland; swh/mrt =sub-alpine wet heathland/montane riparian thicket ecotone; swh/sw = sub- alpine wet heathland/sub-alpine woodland ecotone; LS = litter and soil; LSR = log, soil and roots; LV = litter, vegetation and roots; RS = rock and soil; SV = surface vegetation; R = roots; RS = roots and soil.

209 Capture Site Capture Site N Ñ Ñ ÑÑ

ÑÑÑÑ

Ñ Ñ Ñ Ñ

Ñ ÑÑ Ñ Ñ

Ñ Radio-fix location

Ñ

Ñ Radio-fixDensity location contour (50 - 95%)

Ñ N Density contour (50 - 95%) Ñ Ñ Sub-alpine wet heathland-montane riparian thicket ecotone Ñ Cool temperate mixed forest - drainage line Sub-alpine woodland

Ñ ÑÑ ÑÑ Cool temperate mixed forest 3036MetersÑ ÑÑ ÑÑÑ

Ñ 80816Metres Ñ Ñ

Fig. 6.8a. Utilisation distribution, habitat use and radio-fix locations for frog number 5 over the breeding and post-breeding seasons. Cross-validated, fixed kernel density contours are illustrated at 5% intervals between 50 and 95%. The capture site represents the calling site at which the frog was initially located.

210 N Capture Site

Ñ Ñ

Ñ

Ñ Ñ Ñ Ñ Ñ Ñ Ñ ÑÑ Ñ Ñ

Ñ ÑÑ Ñ

Ñ

Ñ Ñ

Ñ Ñ Radio-fix location

Density contour (50 - 95%)

Montane riparian thicket

3036Metres Cool temperate mixed forest, non-drainage line

Fig. 6.8b. Utilisation distribution, habitat use and radio-fix locations for frog number 20 over the breeding and post-breeding seasons. Cross-validated, fixed kernel density contours are illustrated at 5% intervals between 50 and 95%. The capture site represents the calling site at which the frog was initially located.

211 N

Ñ Ñ

ÑÑ Ñ

Ñ

Ñ Ñ Ñ

Ñ Ñ

Capture Site

Ñ Radio-fix location

Density contour (50 - 95%)

Sub-alpine wet heathland

Sub-alpine woodland

2024Metres

Fig. 6.8c. Utilisation distribution, habitat use and radio-fix locations for frog number 1 over the breeding and post-breeding seasons. Cross-validated, fixed kernel density contours are illustrated at 5% intervals between 50 and 95%. The capture site represents the calling site at which the frog was initially located.

212 N Ñ Ñ Radio-fix location

Density contour (50 - 95%) Ñ Capture Site

Ñ Montane riparian thicket

Cool temperate mixed forest, non-drainage line

Ñ Ñ Ñ Ñ Ñ Ñ Ñ Ñ Ñ

Ñ Ñ Ñ Ñ Ñ

Ñ

1012Metres Ñ

Fig. 6.8d. Utilisation distribution, habitat use and radio-fix locations for frog number 28 over the breeding and post-breeding seasons. Cross-validated, fixed kernel density contours are illustrated at 5% intervals between 50 and 95%. The capture site represents the calling site at which the frog was initially located.

213 N Ñ

Ñ Ñ Ñ Ñ Ñ Ñ

Ñ

Ñ Radio-fix location

Density contour (50 - 95%)

Cool temperate mixed forest, drainage line Ñ T Ñ Ñ Cool temperate mixed forest, non-drainage line Capture Site

1012Metres Ñ

Fig. 6.8e. Utilisation distribution, habitat use and radio-fix locations for frog number 16 over the breeding and post-breeding seasons. Cross-validated, fixed kernel density contours are illustrated at 5% intervals between 50 and 95%. The capture site represents the calling site at which the frog was initially located.

214 Ñ Radio-fix location

Ñ Density contour (50 - 95%)

Ñ Ñ Cool temperate mixed forest, drainage line Ñ Ñ Ñ Ñ Cool temperate mixed forest, non-drainage line

Ñ

N Ñ Ñ Capture Site

Ñ 2024Metres

Fig. 6.8f. Utilisation distribution, habitat use and radio-fix locations for frog number 15 over the breeding and post-breeding seasons. Cross-validated, fixed kernel density contours are illustrated at 5% intervals between 50 and 95%. The capture site represents the calling site at which the frog was initially located.

215 Sub-alpine Elevation

Breeding Season

1

0

-1

-2

-3

Ln daily frog movement (m) -4

-5 1 2 3 4 Ln mean daily relative humidity/ambient temperature ratio

Sub-alpine Elevation

Post-breeding Season

3 2 1 0 -1 -2 -3

Ln daily frog movement (m) -4 -5 1 2 3 4 5 Ln mean daily relative humidity/ambient temperature ratio

Fig. 6.9. Relationship between daily movement > 0 m (natural-log transformed) and mean daily relative humidity/ambient temperature ratio (natural log transformed) at sub-alpine elevation during the breeding and post-breeding seasons.

216 Montane Elevation

Breeding Season

2

1

0

-1

-2 Ln daily frog movement (m)

-3 0 1 2 3 4 5 Ln mean daily relative humidity/ambient temperature ratio

Montane Elevation

Post-breeding Season 2

1

0

-1

-2

Ln daily frog movement (m) -3

-4 1 2 3 4 5 Ln mean daily relative humidity/ambient temperature ratio

Fig. 6.10. Relationship between daily movement > 0 m (natural-log transformed) by frogs and mean daily relative humidity/ambient temperature ratio (natural log transformed) at montane elevations during the breeding and post-breeding seasons.

217 Breeding Season Montane Habitat 1.0 p=0.02 0.9 0.8 0.7 0.6 0.5 0.4 0.3

Probability of movement 0.2 0.1 0.0 -5 0 5 10 15 20 Mean daily temperature category (oC)

Breeding Season Montane Habitat p=0.04 0.8 0.7 0.6 0.5 0.4 0.3 0.2 Probability of movement 0.1 0.0 40 50 60 70 80 90 100 Mean relative humidity category (%)

Fig. 6.11. Relationship between mean daily temperature and relative humidity and the probability (with upper and lower bounds) of frog movement occurring in montane habitat during the breeding season.

218 Post-breeding Season 0.9 Montane Habitat p=0.04

0.8

0.7

0.6

Probability of movement 0.5

0.4 100 125 150 175 200 225 250 275 300 325 Mean daily solar radiation category (W/m2)

Post-breeding Season Montane Habitat 0.8 p=0.02 0.7 0.6 0.5 0.4 0.3 0.2 Probability of movement Probability of 0.1 0.0 3.8 3.9 4.0 4.1 4.2 4.3 4.4 4.5 4.6 Ln mean daily relative humidity category (%)

Post-breeding Season Montane Habitat 1.0 p<0.0005 0.9 0.8 0.7 0.6 0.5 0.4 0.3

Probability of movement Probability of 0.2 0.1 0.0 -0.5 0.0 0.5 1.0 1.5 2.0 2.5 3.0 Ln total daily rainfall category +1 (mm)

Fig. 6.12. Relationship between mean daily solar radiation, relative humidity (natural-log transformed) and total daily rainfall (natural-log transformed + 1) and the probability (with upper and lower bounds) of frog movement occurring in montane habitat during the post-breeding season.

219 Sub-alpine Elevation Season Type 3 Post-breeding Breeding 2

1

0

-1 Lndistance daily (m)moved

-2 -0.4 0.1 0.6 1.1 1.6 2.1 Ln days since arrival of rain-bearing front +1

Montane Elevation Season Type Post-breeding Breeding 2

1

0

-1

-2

Ln daily distance moved(m) distance daily Ln -3

-4 -0.4 0.2 0.8 1.4 2.0 2.6 Ln days since arrival of rain-bearing front +1

Fig. 6.13. Relationship between daily frog movement > 0 (natural-log transformed) and the arrival of frontal weather bearing rainfall (natural-log transformed + 1) at sub-alpine and montane elevations during the breeding and post-breeding seasons.

Post-breeding Season 0.8 Montane and Sub-alpine Elevation P=0.01 0.7

0.6

0.5

0.4

0.3 Probability of movement 0.2

0.1 -1.0 -0.5 0.0 0.5 1.0 1.5 2.0 Ln days since arrival of rain-bearing front category

Fig. 6.14. Relationship between the arrival of rain-bearing frontal systems (natural-log transformed) and the probability (with upper and lower bounds) of frog movement occurring during the post-breeding season at sub-alpine and montane elevations.

220 Sub-alpine Elevation Temperature Reading Breeding Season Substratum 15 Surface C) o 10

5 Temperature (

0 -5 0 5 10 15 20 Ambient temperature (oC)

Sub-alpine Elevation Temperature Reading Post-breeding Season Substratum 20 Surface

15 C) o

10

Temperature ( 5

0 -5 0 5 10 15 20 25 Ambient temperature (oC)

Fig. 6.15. Relationship between ambient, ground surface and substratum temperature recorded at frog sheltering sites during the breeding and post-breeding seasons at sub-alpine elevation.

221 Montane Elevation Temperature Reading Breeding Season Substratum 25 Surface

20 C) o 15

10 Temperature ( 5

0 0 5 10 15 20 25 30 Ambient temperature (oC)

Temperature Reading Montane Elevation Post-breeding Season Substratum Surface 25

20 C) o 15

10 Temperature ( 5

0 0 5 10 15 20 25 30 Ambient temperature (oC)

Fig. 6.16. Relationship between ambient, ground surface and substratum temperature recorded at frog sheltering sites during the breeding and post-breeding seasons at montane elevations.

222 Gravid Female 1994 Adult Male 1994 Gravid Female 1995 Spent Female 1995 4 Spent Female 1994

3

2

Number of frogs 1

031 20 10 30 19 8 28 20 Oct Nov Dec Dec Jan Feb Feb Mar 1994/1995

1995/1996 Date (day/month)

Fig. 6.17. Number and type of Philoria frosti captured in pitfall traps during spring and summer periods of the 1994/1995 and 1995/1996 seasons. Black horizontal bars show periods in which pitfall trapping was undertaken in each season.

223

Chapter 7

LONGEVITY, MATURATION AND SEX-SPECIFIC GROWTH

7.1 Introduction

Knowledge of population demography, dynamics and life history are fundamental elements of conservation biology and wildlife management disciplines (e.g., Primack 1993; Caughley and Gunn 1996; Meffe and Carroll 1997; Fiedler and Kareiva 1998). For amphibians, whilst larvae have long been used to model ecological processes, knowledge of these factors for post- metamorphic populations is not as well developed (Trenham et al. 2000). The recent global phenomenon of amphibian declines and disappearances has further highlighted the lack of long- term population studies required to assess population stability and identify long-term survival prospects, or extinction risk for species (Pechmann et al. 1991; Blaustein 1994; Blaustein et al. 1994; Pechmann and Wilbur 1994; Driscoll 1999).

Extinction risk increases with rarity (Primack 1993; Beebee 1996; Hunter 1996). Demographic, environmental and genetic stochasticity, and the impact of catastrophes have also been identified as four primary, interacting processes that contribute to an increased probability of extinction (Hunter 1996; Shaffer 1997). To effectively manage and protect animal populations, particularly those that are endangered, rare, or have restricted distributions, an understanding of the potential for natural population fluctuations relative to the time scale and magnitude of reproductive cycles and generation times is required (Primack 1993; Caughley and Gunn 1996). For example, the longevity of individuals may influence the stability of a population in its capacity to buffer against unfavourable conditions, or events, which may result in failed or reduced recruitment (Driscoll 1999). Species with individuals that live longer are likely to be less prone to fluctuations in population size, because a cohort or age class resulting from a single year's reproduction will generally make up a smaller proportion of the population than it would for a species with individuals that are shorter-lived. However, species with longer-lived individuals will also tend to recover more slowly from perturbations due to slow growth rates, delayed maturity and low fecundities (Wheeler et al. 2003). Knowledge of the age structure of a population may also provide insight into the potential for changes in population size to occur. Driscoll (1999) noted that sporadic recruitment or mortality within a population will result in a less regular distribution of age classes, and increased instability, compared to a population with consistent recruitment and mortality where a steady decline in the number of individuals in successive age classes is expected.

225 Mark-recapture studies have been successfully used in documenting population demography, dynamics and life history of amphibians (e.g., Semlitsch 1983; Caldwell 1987; Berven 1995; Jehle et al. 1995; Williamson and Bull 1996; Hels 2002). However, use of mark-recapture procedures can become problematic when dealing with secretive or rare species, or those with lifespans ranging from years to decades (Trenham et al. 2000), and can be very labour intensive or take a long time to yield a result (Kusano et al. 1995a). The recent improvement in use of skeletochronological methods for aging amphibians (e.g., Castanet and Smirina 1990; Castanet et al. 1993; Smirina 1994) has minimised or overcome these problems, and with the successful use of phalanges instead of long bones (e.g., Hemelaar 1988; Bastien and Leclair 1992; Sinsch et al. 2001), has resulted in a non-lethal aging technique that allows the collection of demographic information more rapidly than in mark-recapture studies (Gibbons and McCarthy 1983; Halliday and Verrell 1988). Skeletochronology is based on evidence revealed in the histology of the bone, where annual growth is cylindrical, with broad, less dense zones being associated with periods of rapid growth, and narrower, more dense zones being associated with the temporary cessation of local osteogenesis during periods of decreased growth. Transverse sections obtained from the diaphyses of amphibian bones possess haematoxylinophilic lines of arrested growth (LAG) (Castanet et al. 1993). The pattern of one broad zone and one LAG at the periosteal margin generally constitute one cycle of annual growth for amphibians that occupy regions with predictable climate cycles of warm and cold, or wet and dry (Wake and Castanet 1995). Examination of these LAG can provide information on growth (distance between two successive LAG), age at sexual maturity, longevity and periodic or aperiodic annual activity (e.g., Castanet et al. 1992 [cited in Eggert and Guyétant 1999]; Castanet et al. 1993; Smirina 1994).

There are, however, a number of processes evident in the histology of bone that need to be interpreted or evaluated when using skeletochronological methods. The first of these involves an understanding of processes that destroy or remove growth marks which can hamper age estimates (e.g., Smirna 1972; Leclair and Castanet 1987). These processes include endosteal bone resorption, where the first periosteal deposition (the first LAG) is destroyed, and bone reconstruction (Hemelaar and van Gelder 1980; Castanet and Smirina 1990). Leclair (1990) identifies a number of factors that may contribute to the destruction of the first LAG, including post-metamorphic growth, reproductive phenology and environmental conditions, and the ratio of minimum to maximum size at metamorphosis. The resorption rate of bone, however, can be evaluated by examining individuals of known age (e.g., Caetano et al. 1985; Acker et al. 1986; Forester and Lykens 1991; Eggert and Guyétant 1999; Alcobendas and Castanet 2000).

The second of these processes involves validation of the assumption that formation of LAG reflects annual growth. For example, LAG may not represent annual growth in some species, with double LAG (Caetano et al. 1985; Forester and Lykens 1991), false LAG (Smirina 1994) and double annual LAG (Caetano et al. 1985) being recorded. Validation can be achieved by

226 comparing estimates of age from skeletochronology and other methods, including recapture of individuals of known age (e.g., Cadwell 1987; Berven 1995; Marunouchi et al. 2002.), applying chemical marks to actively growing portions of bone during intervals of growth (Smirina 1972), or by obtaining successive bone samples from the same individual on two or more occasions to examine the number of added LAG (e.g., Smirina 1972; Gibbons and McCarthy 1983; Driscoll 1999; Measey 2001; Khonsue et al. 2001; Ento and Matsui 2002). Other factors that may also require further interpretation when using skeletochronological methods include the distinctiveness and spatial distribution of growth marks, and peripheral LAG when growth has stopped (Castanet and Smirina 1990). The combined use of mark-recapture techniques with skeletochronology and morphometric data appears to provide the most robust method for aging amphibians (Halliday and Verrell 1988; Measey 2001).

To date, little information is available on the demographic attributes of P. frosti. The undertaking of a mark-recapture study using pitfall traps and drift fence over two breeding and post-breeding seasons failed to acquire sufficient captures to demonstrate its usefulness as a technique in obtaining information on population biology of the species (see Chapter 6). Malone (1985a) obtained some information on the site fidelity of calling males within and between breeding seasons by marking individuals that he captured. He found that males continually disappeared from calling sites throughout the breeding season, recapturing only 6.3% of individuals marked in the same breeding season, and no recaptures were obtained in the following breeding season. Given the cryptic nature and current rarity of P. frosti, and its occupancy of a temperate climate, skeletochronology appears to be a highly suitable method for age determination in this species. Its use is further supported by the prospect of using phalangeal bones, thereby precluding the requirement to destroy live individuals.

The current threatened conservation status of P. frosti generates an urgent requirement to acquire demographic information on the species to enable modelling of its long-term viability. In this chapter I use skeletochronological methods to examine: (1) longevity; (2) age to sexual maturity; (3) age-specific growth; and (4) adult age structure in bone samples of P. frosti obtained from the wild population and specimens lodged in the Museum of Victoria.

7.2 Methods

7.2.1 Breeding Biology and Life History

Philoria frosti occupies high elevation, sub-alpine and montane habitats, where it is known to be active during spring, summer and autumn (see Chapters 5 and 6), and presumed to be inactive during the winter months when snow lies on the ground mostly between July and September.

227 Adult females are larger than adult males, measuring on average 51.6 and 44.5 mm snout-vent length (SVL), respectively (Malone 1985a, b). Investigations into the calling behaviour of P. frosti (Chapter 5), and observations from previous studies on breeding activity (Littlejohn 1963; Malone 1985a), show that the breeding season of the species can range from early spring to early summer (September – December). Malone (1985a) observed calling activity to occur between late October and late December, but recorded oviposition over a shorter period of 2 – 3 weeks (11 – 29 November) when a peak in calling was noted. Under natural conditions, he recorded embryonic development lasting 5 – 8 weeks, hatching at stage 22 - 23 of Gosner (1960), and larval development lasting 5 – 10 weeks, at which time individuals metamorphosed at stage 42, and measured an average length of 6.72 mm SVL. Malone’s observations suggest that if most egg masses were deposited in mid November, metamorphosis would have most likely occurred between late January and mid to late March. However, the results from Chapter 5 show that peaks in calling activity can occur as early as the beginning of October, suggesting that metamorphosis could occur as early as mid December in some years, based on Malone’s development statistics. However, it is likely that rates of embryonic development during early October would be significantly slower than at latter periods due to lower temperatures, thereby extending the development period.

7.2.2 Toe-clipping for Mark-recapture and Ageing

During the course of this thesis, tissue samples in the form of toe clips were surgically removed from a total of 92 P. frosti located in the field. Toe clips removed from 88 live-captured individuals were immediately preserved in 4% aqueous formaldehyde solution, whilst four deceased specimens, located during censuses for calling males, were preserved in 75% ethanol prior to toe clips being removed. Live-captured individuals were marked following the toe- clipping scheme of Hero (1989) for recognition if recaptured, prior to their release. Toe clips were sampled from 58 adult males and 28 adult females, with the remaining six frogs being immature. Frogs were captured using different techniques whilst undertaking various components of this study, including pitfall trapping, active searching and acoustic triangulation of calling males. Adult females were also located when searching for calling males. Initially, only terminal phalanx were removed from captured frogs, however, the results of preliminary histology showed that these were inadequate for age determination, and that secondary phalanx were required (see results). Subsequent toe clipping of field specimens therefore involved the removal of the distal two phalangeal elements.

Captured frogs were also measured prior to their release, including SVL to the nearest millimetre (with dial callipers), and weight to the nearest 0.5 g (with Pesola spring scales). Only SVL measurements were collected from the four specimens of P. frosti found deceased in the field. As most frogs were obtained by means of locating calling males during the breeding season, most tissue samples were removed from sexually mature frogs. Secondary sexual characteristics were

228 used to identify the sex of adults; adult males had dark-brown, pigmented vocal sacs and adult females comprised inner fingers that were flanged. Toe clips were also removed from four post- metamorphic frogs to investigate the presence of the first LAG and the potential for endosteal resorption (see section 7.2.5).

To supplement tissue samples removed from frogs captured in the field, samples were also removed from 122 specimens of P. frosti lodged in the Museum of Victoria. This sample included 90 adult males and 23 adult females, identified by secondary sexual characteristics and specimen details from the Atlas of Victorian Wildlife (Department of Natural Resources and Environment, Heidelberg), and nine immature frogs that did not possess secondary sexual characteristics. Four of these nine individuals were considered to be sub-adult females based on their large size (37.7, 39.4, 40.6 and 41.3 mm SVL) and absence of pigmented vocal sacs, whilst the remaining five were considered to be juveniles, but of unknown sex (size range = 19.1 – 32.9 mm SVL). Prior to removing tissue samples, specimens had been preserved in 75% ethanol or 10% buffered formalin solutions, and collected from the period encompassing 1955 – 1983. Due to their preserved state, only SVL was measured for each specimen. No measurements were taken from three specimens that were contorted during the process of preservation.

7.2.3 Histological Procedure

Phalangeal samples were subject to the following histological procedure:

(1) decalcification for 12 h using RDO rapid decalcifying solution (Phoenix Scientific, Melbourne), followed by rinsing under tap-water for approximately 2 h; (2) dehydration by submergence in a series of aqueous ethanol solutions of increasing concentration, followed by a histological solvent (Histolene, Phoenix Scientific, Melbourne); (3) vertical embedment into blocks comprising Paraplast embedding medium (Oxford Labware, St Louis, Missouri); (4) cutting of transverse sections from the mid-shaft diaphyses of the secondary phalanx at a thickness of 5 µm using a rotary microtome (preliminary histology involved sections from the distal phalanx being cut prior to recognition of their inadequacy for age determination); (5) floating of transverse sections onto glass slides using a water bath heated to 55 oC; (6) oven-drying of slides for approximately 12 h at 60 oC, prior to being deparaffinised in Histolene, rehydration in series of aqueous ethanol solutions of decreasing concentrations, and rinsing in de-ionised water; (7) staining of each slide using Ehrlich’s haematoxylin (Humason 1979) for a period of 30 min, followed by rinsing in running tap-water for 5 min to remove excess stain; (8) dehydration of each slide in a series of aqueous ethanol solutions of increasing concentrations, followed by the solvent Histolene; and

229 (9) mounting of each slide with cover-slips using DPX mounting medium (BDH Laboratory Supplies, Poole, England).

7.2.4 Age Determination

A binocular compound microscope was used to examine each stained section from each sample at a magnification of X 400. Each sample was assessed independently by myself and a colleague (Michael Scroggie), and the number of lines of arrested growth (LAG) counted. Diaphysis sections chosen for counting LAG included those in which the size of the medullar cavity was at its minimum and that of the periosteal bone at its maximum. The resulting counts from each observer were then compared for consistency, and a subsequent third examination undertaken to resolve any inconsistent results. Uninterpretable samples, including those that could not be resolved satisfactorily between observers, were excluded from further analyses.

To establish a reference point from which age and growth estimates could be made, metamorphosis was considered to be the birth point, where age = zero. As the known time of metamorphosis can range in P. frosti from late January to late March, or possibly early April, the birth point for different individuals could vary by up to approximately 2.5 months. For ease of interpretation, the birth point of all samples was assigned to March 1, the approximate midpoint of the time period in which metamorphosis has been observed. The period when tissue samples were collected from frogs located in this study was largely confined to the breeding season (mainly the months of October and November), but with several samples from late summer and autumn. The sampling of tissue from museum specimens was taken to be the time of year when each was collected, and presumably preserved. Most of these specimens were also collected during the breeding season. For simplicity, tissue samples obtained during the breeding season were considered to have been removed on November 1, the approximate midpoint of the time period in which breeding has been observed. It is assumed that LAG are deposited in the coldest months of the year (June – September), when growth is suppressed, and that following the deposition of the first LAG at the end of winter, an individual would be approximately six months of age, not one year old. Samples were therefore assumed to have ages of N years minus six months, where N = the number of LAG observed in transverse sections of bones. Because tissue samples acquired later at the end of summer, and in autumn, would comprise the same number of LAG as samples acquired during the breeding season, their age was also determined in the same manner as described above for simplicity, even though they would have been up to several months older.

230 7.2.5 Validation

The method of comparing the number of LAG in successive bone samples from the same individual was considered to be the most appropriate procedure to validate LAG as representing annual growth in P. frosti. As the application of pitfall traps and drift-fences has been shown to be currently unsuitable for obtaining recapture data from the species (Chapter 6), active searching at known breeding sites during the breeding season was considered to provide the greatest opportunity to locate calling males or adult females more than once over different years. Unfortunately, this still resulted in the recapture of only one adult male for analysis.

Deposition of the first LAG, and potential for endosteal resorption, was investigated by examining four captured juvenile frogs known to be offspring from the previous season of breeding (G. Hollis pers. obs.). These frogs were captured on 1 December, 1997, and assuming that they metamorphosed on approximately 1 March, 1996, would have been nine months post- metamorphosis. As these frogs had been subject to three months of growth during autumn, one over-wintering period (approximately June – September), and a further 2 – 3 months of growth up until December, it was expected that transverse sections from these individuals would have one LAG evident if endosteal resorption had not taken place.

7.2.6 Analysis of Growth

The procedure of modelling growth by von Bertalanffy curves (von Bertalanffy 1938) was used to estimate the pattern of growth in P. frosti according age, as determined through skeletochronology. The von Bertalanffy growth curve predicts size (in this study SVL) of an individual at time t, following the equation, SVL= a*(1 – b*exp-kt), where a is the asymptotic body size, b is the hypothetical age at which animals are at size zero, and k is the growth constant, which defines the shape of the curve, and the rate at which a is approached (see Fabens 1965). Point estimates of the parameters a, b and k and their asymptotic 95% confidence intervals were made using the non- linear regression module within SYSATAT 8 (SPSS, Inc, Evanston, Illinois). Best-fitting growth curves were derived separately for males and females at sub-alpine (≥ 1270 m) and montane elevation (< 1270 m) using the iterative, Gauss-Newton, least-squares regression option. Means and 95% confidence intervals of each parameter were used to compare growth trajectories for each population examined. Due to the limited number of non-adult frogs examined in this study, the same non-adults were used in constructing growth curves for each respective population examined. For the purpose of this study, it is therefore assumed that size at metamorphosis and growth of non-adult P. frosti is not different between sex and elevation. An error was judged to have been made in the measurement of SVL for one museum specimen (19.1 mm SVL) aged at eight years. This obvious outlier was excluded from growth analyses.

231 7.3 Results

Tissue sections from a total of 214 toes were examined for aging. This comprised 184 samples from which LAG counts were successfully retrieved, including 121 museum samples, 60 field samples and three deceased samples found in the field. Sections from the distal phalange of 27 field samples were not suitable for age determination (see below). One field sample was lost before sectioning, and one museum sample and one sample from a deceased specimen could not be interpreted.

7.3.1 Interpretation of Histology

Preliminary histology conducted on 27 distal phalanges removed from frogs captured in the field showed that substantial resorption of endosteal bone had occurred in each sample. This resulted in sections that consisted of a large medullar cavity as well as having lost a variable number of inner LAG.

Examination of the diaphyseal region from secondary phalanges showed that there were clearly defined LAG distributed throughout the periosteal bone, with a maximum of 15 observed from all samples. In several samples, individual LAG were made up of two inconsistent lines, but not for all LAG present in a sample. Bone remodelling, through resorption of endosteal bone and reconstruction of periosteal bone surrounding the medullary cavity, was limited, or absent in all samples, and not extensive enough to lead to the removal of LAG. Inner LAG were spatially more separated than outer LAG in most samples. In older individuals the outer LAG were located very close together in the outer region of the periosteal bone (Fig. 7.1a). The close proximity of outer LAG in older individuals, in association with poorer quality sections, resulted in the most recent LAG being difficult to decipher in some sections. It is therefore possible that some of the oldest individuals may have comprised a greater number of LAG, and that the age of these samples may have been slightly under-estimated.

7.3.2 Validation of Skeletochronolgy

Examination of four juvenile frogs known to have undergone one over-wintering period showed that they all had one LAG with limited or absent remodelling of bone (Fig. 7.1a). Examination of the successive bone samples collected from the single recaptured individual demonstrated that LAG are added at the rate of one per year. Four LAG were observed in the tissue sample from the

232 male captured on the 17 October 1999, whilst five LAG were observed in the sample acquired approximately one year later on the 7 November 2000.

7.3.3 Longevity, Maturation and Age Structure

With 15 being the maximum number of LAG counted from all samples, the oldest sample examined was 14.5 years. The maximum age of females (14.5 years) was greater than that of males (13.5 years), whilst adult males were found to reach reproductive maturity one year earlier (3.5 years) than adult females (4.5 years) (Fig. 7.2). Notwithstanding the origin of each population examined (museum or field), the maximum age of males and females from sub-alpine elevations (13.5 and 14.5 years, respectively) was greater than that from the montane elevations (9.5 and 11.5 years, respectively). Adult females from montane elevations reached reproductive maturity one year later (5.5 years) than those at sub-alpine elevations, whilst males reached reproductive maturity at the same age (3.5 years) in both sub-alpine and montane populations. Because of the very small sample size, the ages of four adult, museum specimens from montane elevations are not presented in the histograms in Figure 7.2. These individuals included three males (5.5, 6.5 and 7.5 years of age) and one female (9.5 years of age), all of which did not alter figures for maturation or presented age class distributions.

Examination of the distribution of adult male and female age classes from the sample derived from museum specimens collected at sub-alpine elevations showed that the highest frequency of adult males occurred in the younger age classes (5.5 – 7.5 years) compared with adult females where the highest frequency of individuals occurred in older age classes (9.5 – 10.5 years) (Fig. 7.2a). The distribution pattern of age classes representing adult males from field samples from both sub- alpine and montane elevations contrasted with the museum sample of the sub-alpine population, with the highest frequency of individuals occurring in the youngest age class (3.5 years), particularly in the montane population, which was highly skewed to the left (Fig. 7.2b and c). The highest frequency of adult females from the field sample from sub-alpine elevations occurred in the 4.5 and 7.5 age classes (Fig. 7.2b), whilst the age class representing 8.5 years had the highest frequency of adult females from the field sample from montane elevations (Fig. 7.2c). Compared with the sub-alpine museum sample, there was a distinct lack of individuals in older age class groups in both sub-alpine and montane field samples.

7.3.4 Growth

Best-fitting von Bertalanffy growth curves for male and female samples from sub-alpine and montane elevations are presented in Figure 7.3, whilst values of the parameters a, b and k for each curve, and their 95% confidence intervals, are presented in Table 7.1. The asymptotic body-size (parameter a) of females from sub-alpine elevations (47.9 ± 3.1 mm SVL) was significantly larger

233 than that of males from sub-alpine elevations (41.9 ± 1.9 mm SVL), as identified by their non- overlapping confidence intervals. Estimates of a were not significantly different between other sex or elevation comparisons, although the estimate for montane females (52.8 ± 14.0 mm SVL) was also larger than that for montane males (46.4 ± 3.8 mm SVL), whilst the sizes of males and females from montane elevations were larger than males and females from sub-alpine elevations, respectively. Neither estimates for b or k were significantly different between all sex and elevation comparisons.

7.4 Discussion

Based on the knowledge that P. frosti has a distinctive active and non-active period, and that examination of phalangeal sections taken from a single individual over successive years demonstrated that LAG are deposited annually, one can conclude that counts of LAG are a reliable indicator of age in P. frosti. The limited remodelling of bone that was detected when examining secondary phalanx and that one LAG was present in the four juvenile frogs known to have undergone one inactive (over-wintering) period further supports the use of LAG as an indicator age. The presence of some supplementary LAG, in association with normally observed LAG, in several individuals did not influence age determination. Each supplementary LAG was always positioned very close to, or overlapped with, the LAG it accompanied throughout its circumference. Due to this close proximity, it is likely these supplementary LAG have resulted from short interruptions to the inactive period of P. frosti due to minor climatic fluctuations (e.g., Houck and Francillon-Vieillot 1988). The use of skeletochronology has thus provided a means of obtaining information on the demography and reproductive strategy of P. frosti that would have otherwise been unobtainable.

Tissue samples of P. frosti from both field-captured and museum specimens were also shown to be suitable for skeletochronology examination, and that the different solutions used to preserve museum specimens (see methods) did not appear to alter the detectability of LAG. Tissue samples from deceased frogs that had not undergone considerable decomposition were also found to be suitable. The only minor difficulty confronted in thus study was in deciphering the number of LAG deposited in the outer periosteal bone of a small subset of older individuals (mostly 10 years of age and older), particularly in poorer quality sections. The distinctiveness and spatial distribution of peripheral LAG after growth has stopped is considered by Castanet and Smirina (1990) as a factor that sometimes requires additional interpretation when using skeletochronology. However, it is considered that this difficulty may have resulted in under-estimation of age by only 1 – 2 years.

Unfortunately, there are no data on the longevity of species related to P. frosti with which to compare the results of this study. Among amphibians, salamanders generally live longer than

234 anurans, and larger species tend to live longer than smaller ones (Duellman and Trueb 1994). Tropical amphibians, which generally lack a dormancy period, tend to live shorter lives, whilst species living at high latitudes or altitudes, with relatively short activity periods, may live the longest (Beebee 1996). Inter-populational studies also show that amphibians that occupy cooler climates have greater longevity (Hemelaar 1988; Ryser 1996; Sagor et al. 1998). Due to the small sample sizes, and disparity between numbers of museum and field samples from different elevations, it can not be determined from this study if there are differences in the longevity of P. frosti from sub-alpine and montane populations. The narrow altitudinal range from which P. frosti is known to occur (960 – 1560) may not be substantial enough for differences in longevity to have evolved. The maximum recorded age of P. frosti (14.5 years) falls well within the range of longevity estimates presented for other anurans by Duellman and Trueb (1994), although the authors note discrepancies between longevity estimates from captive and wild samples. If some of the older samples of P. frosti were underestimated by 1 – 2 years (for the reasons given above), then a maximum longevity in the order of 17+ years might be conceivable for P. frosti. The results of this study suggest that P. frosti is relatively long-lived compared to some anurans.

As for longevity, the age at attainment of sexual maturity also tends to be greater for species or populations at high altitudes or latitudes (Duellman and Trueb 1994). The age at which P. frosti was first recorded as being sexually mature (3.5 and 4.5 years for males and females, respectively) falls generally within the range of maturation ages reported by Beebee (1996) for temperate anurans (2 – 5 years). A review by Duellman and Trueb (1994) shows that a large proportion of anurans reproduce at the age of one year, whilst a smaller proportion may require up to four years for males and six years for females. The general pattern of males maturing one year earlier than females among anurans (Duellman and Trueb 1994; Beebee 1996) was also recorded in this study. Although females from montane elevations were recorded as reaching sexual maturity one year later (5.5 years) than from sub-alpine elevations (4.5 years), due to the limited sample size, it is not possible to interpret this result with any certainty. Ambient and substratum temperature measurements made from a sub-alpine and montane breeding site (Chapter 5) indicate that montane habitats may be, on average, cooler that sub-alpine habitats during the breeding season due to increased solar insulation. If this were the case, then males and females from montane elevations may reach sexual maturity at a latter age even though no difference was found in the attainment of sexual maturity by males from either elevation.

The pattern of growth derived for male and female P. frosti from sub-alpine and montane elevations was very similar, with most growth occurring prior to the attainment of sexual maturity, after which it decreased considerably (Fig. 7.3). This pattern of growth was also evident when examining the distribution of LAG in tissue sections, with a general tightening of LAG following the maturation ages identified (e.g., see female P. frosti in figure 7.1b). The tightening of LAG following sexual maturity is described as ‘rapprochement’, and is considered to be related to the onset of sexual maturity where resources utilised for somatic growth are diverted to reproduction

235 (Francillion-Vieillot et al. 1990). However, at sub-alpine elevations, females attained a significantly larger asymptotic size than males. Asymptotic estimates of size from specimens from montane elevations also suggested this size dimorphism between males and females, although they were not statistically different in this case. It is possible that the reduced sample size from montane elevations may explain the larger errors recorded. Malone (1985a) recorded average sizes of adult male and female P. frosti (from predominantly sub-alpine habitat) as 51.6 mm and 44.5 mm SVL, respectively. These values are higher than asymptotic estimates made from specimens collected from sub-alpine elevations in this study, but lower than estimates from specimens collected from montane elevations.

Evaluation of the distribution of adult age classes of museum specimens from sub-alpine (Fig. 7.2a) may provide an indication of the general stability of the male and female population for the period 1955 – 1983. A breakdown of the time period from which specimens were collected shows that there was generally an equal representation of specimens for the periods 1955 – 1961, 1963 – 1969 and 1974 – 1983. Although the museum samples are from a relatively wide time frame (29 years), where there is an increased likelihood of smoothed age class distribution, the relatively gradual reduction in successive age classes either side of the peak for males and females suggest that levels of recruitment and mortality over the time period was relatively consistent (following Driscoll 1999). Alternatively, it is also possible that the observed age class peaks for males and females may indicate favourable conditions that resulted in high recruitment, although such interpretation is likely to be more relevant when examining the age structure of populations over a shorter time period (e.g., Tinsley and Tocque 1995; Driscoll 1999).

Due to their small size, additional caution has to be taken when interpreting the distribution of age classes within the field samples from sub-alpine and montane elevations collected in this study. The addition of 27 tissue samples found to be unsuitable for age determination would have bolstered the sample size from sub-alpine elevations considerably, but could not be used. In comparison with the museum population above, the field samples are from a relatively narrow time period (1991 – 1997) where one might expect either recruitment or mortality from individual years to have a larger impact on the overall age-class distribution of each population. The age- structure distributions of both sub-alpine and montane samples comprising adult males are skewed towards a higher frequency of younger age classes, particularly the montane sample, as well as there being a general lack of individuals in older age classes. This pattern suggests that there has been some recent recruitment of males into each population, but that survivorship of males to an older age may have been compromised. An absence of older adult females from sub-alpine elevations is also evident, suggesting that their survivorship to older age may have also been compromised. The distribution of the very small sample of adult females (5) from montane elevations can not be realistically interpreted, but does have a general modal distribution pattern similar to adult females in the museum sample from sub-alpine elevations.

236 Notwithstanding the sample size and interpretative limitations of the field samples from sub-alpine and montane elevations, if one were to consider them to representative samples of the extant population, their age-class distributions would generally suggest population instability compared with the more stable age structure inferred from the museum sample from sub-alpine elevations. This population instability could be related to the decline of the population of P. frosti some time after 1985 (see Chapter 3). Given that P. frosti is relatively long-lived compared to other anurans, its population should have a greater capacity to buffer against perturbations resulting from stochastic events. Population monitoring in Chapter 3 shows that the decline of the adult male population, from initial estimates made by Malone (1985a) in 1983 and 1984, occurred at the very latest in 1993 (the first post-Malone monitoring survey). This timing suggests that the decline of the adult population from sub-alpine elevations, and possibly montane elevations, has occurred over a period that is less than the longevity of individuals recorded in this study. If the causal mechanism responsible for the decline of P. frosti was acting in a way that substantially reduced recruitment into the population, then one might expect the adult age structure of samples collected between 1991 – 1997 to be dominated more by older age classes. The resulting age class distributions are generally contrary to this pattern, suggesting that the agent responsible for the decline of P. frosti could have impacted on the post-metamorphic life-stage of the population. ______Skeletochronological techniques were successfully used to determine longevity, maturation and growth in P. frosti. Examination of the histology of tissue samples from four juvenile frogs known to have undergone one over-wintering period, and successive samples collected from a single recaptured frog, infer that LAG are added at the rate of one per year in P. frosti, with limited or absent remodelling of bone. Examination of all samples of P. frosti showed that individuals of the species are relatively long-lived (14.5+ years), and attained sexual maturity at an older age when compared generally to other anurans. Adult females captured from montane elevations during this study reached reproductive maturity one year later (5.5 years) than those from sub-alpine elevations (4.5 years), whilst males reached reproductive maturity at the same age from both sub- alpine and montane elevations (3.5 years). Best-fitting von Bertalanffy growth curves for male and female samples from sub-alpine and montane populations show that the asymptotic body-size of females from sub-alpine elevations was significantly larger than that of males from sub-alpine elevations, but there were no other differences in growth parameters for other sex and elevation comparisons. The pattern of growth derived for male and female samples from sub-alpine and montane elevations was very similar, with most growth occurring prior to the attainment of sexual maturity, after which it decreased considerably. A skewed age-class distribution of adult samples from the extant population of P. frosti infered population instability when compared to the binomial age-class distribution of museum samples. The dominance of younger-aged individuals in the extant adult population following the decline of the species suggests that factors responsible for the decline may have impacted on the post-metamorphic life-stage of the population.

237 Table 7.1. Parameter details of von Bertalanffy growth curves derived for male and female samples acquired from sub-alpine (≥ 1270 m) and montane (< 1270 m) elevation.

CI = confidence interval. a = asymptotic body size; b = hypothetical age at which animals are at size zero; k = growth constant; r2 denotes the correlation between the observed values and the predicted values. N.B., the same non-adult frogs were used in the construction of each growth curve. See methods for description of parameters and other growth-curve details.

Elevation Sex estimate of a estimate of b estimate of k nr2 (95% CI) (95% CI) (95% CI) Sub-alpine Male 41.85 (39.91, 43.78) 0.73 (0.57, 0.88) 0.43 (0.27, 0.57) 104 0.54 Sub-alpine Female 47.91 (44.85, 50.97) 0.82 (0.65, 0.99) 0.41 (0.21, 0.61) 45 0.76 Montane Male 46.40 (42.56, 50.25) 0.82 (0.68, 0.96) 0.47 (0.28, 0.67) 41 0.82 Montane Female 52.76 (38.79, 66.73) 0.76 (0.62, 0.87) 0.22 (0.01, 0.43) 18 0.84

238 EB

MC (a) X 323 magnification

EB (b)

X 100 MC magnification

Fig. 7.1. Transverse section through the diaphyseal region of secondary phalanx of Philoria frosti aged six months (a) and 12.5 years (b). Arrows denote lines of arrested growth (LAG), EB = endosteal bone, MC = medullary cavity.

239 (a) adult females 30 adult males

25

) 20

%

(

y

c

n 15

e

u

q

e

r

F 10

5

0 2 3 4 5 6 7 8 9 10 11 12 13 14 15 Age class (years)

adult females 45 (b) adult males

36

)

%

( 27

y

c

n

e

u

q 18

e

r

F

9

0 2 3 4 5 6 7 8 9 10 11 12 13 14 15 Age class (years)

adult females 45 adult males (c) 40 35

) 30

%

(

y

c 25

n

e

u 20

q

e

r

F 15 10 5 0 2 3 4 5 6 7 8 9 10 11 12 13 14 15 Age class (years)

Fig. 7.2. Frequency distribution of age classes for adult frogs from populations comprising (a) male (n = 86) and female (n = 22) museum specimens from sub-alpine elevation (> 1270 m), (b) male (n = 12) and female (n = 10) field samples from sub-alpine elevation, and (c) male (n = 30) and female (n = 5) field samples from montane elevations (≤ 1270 m).

240 50 50

40 40 )

30 30 Sub-alpine males Montane males 20 20

Snout-vent length (mm 10 Snout-vent length (mm) 10

0 0 0 2 4 6 8 10 0 5 10 15 Age (years) Age (years)

60 60

40 40

Sub-alpine females Montane females

20 20 Snout-vent length (mm) Snout-vent length (mm) length Snout-vent

0 0 0 5 10 15 0 5 10 15 Age (years) Age (years)

Fig. 7.3. Relationship between age and size for male and female samples from sub-alpine and montane elevations. Growth is depicted by best-fitting von Bertalanffy curves.

241

Chapter 8

SYNTHESIS AND CONCLUSIONS

8.1 Distribution, Abundance and Pattern of Decline

The comprehensive examination of the distribution of P. frosti in this study has both extended the range of the species, as well as recording a contraction in range from areas that it formerly occupied. This new information increases the area of known distribution from 80 km2 (Malone 1985a) to 134.5 km2, based on the sum of potential areas that could be occupied by the species (Chapter 3, Fig. 3.3). Despite searches in other areas containing potential habitat for P. frosti (Chapter 3), no records of the species are known from outside the vicinity of the Baw Baw Plateau region.

The extension of known range of P. frosti has occurred primarily in lower-elevation, montane habitats on the south-western escarpment of the plateau, where the species was recorded as low as approximately 960 m. This lower limit is a further extension from that initially reported by Osborne et al. (1999), with additional populations being recorded in the headwaters of the Tanjil River West Branch in the north-west of the study area (see Fig. 3.4). A smaller contribution to this range extension has also occurred on the north-eastern escarpment of the plateau, where the species was recorded as low as 1200 m in South Cascade Creek. This range extension of the species, however, is related to search effort only, with previous surveys, or anecdotal searches, being confined primarily to elevations above 1300 m in the Baw Baw National Park and Mt Baw Baw Alpine Resort (e.g., Malone 1985a). The contraction in range of P. frosti recorded in this study has occurred predominantly on the north-eastern side of the Baw Baw Plateau between 1400 and 1500 m elevation, and includes areas from which the species has not been recorded since the survey of Malone (1985a) in 1983 and 1984 (Fig. 3.4).

Historic and extant distributions of P. frosti show that the species is generally confined to the Baw Baw land system comprising granodiorite parent rock. However, within this land system, the species has preference for particular landform components based on soil type, drainage condition, aspect, vegetation and climate (Chapter 4). Populations of P. frosti were not detected on the north- eastern escarpment of the Baw Baw Plateau below 1200 m during this study, even though suitable habitat is present to elevations as low as approximately 1000 m, similar to the south-western escarpment. Without any base-line data to compare to the results of this study, this distribution

243 pattern infers that the species has either declined, or is absent from these apparently suitable habitats. It appears more likely that the absence of P. frosti from this region is related to climatic conditions that prevail in different areas of the plateau and escarpment, with drier, warmer conditions present on the north-eastern side compared with the wetter, cooler conditions on the south-western side (Chapter 3). This explanation of distribution limit is supported by the apparent absence of P. frosti in the upper sections of a number of north-eastern drainage systems when surveyed in 1983 and 1984 by Malone (1985a) at a time when the species was considered to be relatively common. It is also possible that P. frosti may potentially occupy habitat on the north- eastern escarpment during periods of more favourable climate, but has not occupied the region recently. The extension in distribution range of P. frosti recorded in this study highlights the need for amphibian surveys to be based on potential habitat rather than perceived habitat or historical locations.

The results of this study clearly show that the relative abundance of P. frosti in sub-alpine habitats above 1300 m has declined by approximately 98% from counts reported by Malone (1985a) in 1983 and 1984. When comparing this study’s estimate of total male population size above 1300 m (2,352) with Malone’s estimate of 10,000 – 15,000, a decline in the order of 84% has occurred. However, this level of decline is probably an underestimate, due to the fact that Malone’s original count of 4,248 males did not account for a substantial amount of potential habitat between 1300 and 1400 m elevation, where the highest density of calling males was recorded in this study (Chapter 3). With relatively little base-line data available to assess montane populations below 1300 m, it can not be determined conclusively if these populations have also declined. The limited data available suggest that these populations have declined (Chapter 3).

Closer examination of 1993 – 2002 survey data at different elevations suggests that the population of P. frosti is continuing to decline in sub-alpine habitats between 1400 and 1500 m, but has increased very slightly in habitats between 1300 and 1400 m. In montane habitats below 1300 m, the results of surveys conducted between 1997 and 2002 suggest a slight decline in population numbers, although the data are from only a short period. Within the Mt Baw Baw Alpine Resort, P. frosti appears to have disappeared from former breeding localities that have been subjected to disturbance associated with the development of ski runs and building infrastructure. However, populations remain on resort land in habitats subjected to little or no disturbance, and in similar proportions to other localities on the Baw Baw Plateau at similar elevation. In all years of survey, including those of Malone (1985a) in 1983 and 1984, the number of calling males recorded has increased in number progressively from the east to the west side of the plateau. It is in the vicinity of the eastern side of the plateau that a contraction in distribution range of P. frosti was recorded in this study. Estimates of the total adult male population of P. frosti from different regions of elevation across the plateau also support this overall pattern of abundance (Chapter 3).

244 Philoria frosti is one of 33 amphibians reported to have declined in Australia (Campbell 1999, and references therein), although for many of these species there is clearly a lack of long-term distribution and abundance information to interpret population declines with any degree of confidence. It is considered that population data from this study was of sufficient size and quality to examine population trends and determine the current conservation status of P. frosti (Chapter 3). Data collected from systematic surveys of P. frosti were examined over a period of 20 years (1983 – 2002), longer than the estimated life-span of the species (14.5+ years), and approximately 3 – 4 times greater than the period taken for the species to reach sexual maturity (see Chapter 7). Results of retrospective power analyses also showed that there was sufficient power to detect changes in population size of less than 10% in sub-alpine (> 1400 m) and sub-alpine-montane (1300 – 1400 m) habitats where surveys date back to 1983. Sufficient power to detect changes in population size of less than 10% was also attained for monitoring of post-1996 survey transects at montane elevations.

This study has also attempted to account for some of the limitations that exist when using counts of calling males as a census method, thereby reducing potential error in estimates of abundance for P. frosti (Chapter 3). The results highlight the need for other amphibian monitoring programs contemplating the use of acoustic surveys to consider temporal, spatial and environmental factors in survey design, as well as statistical power of the monitoring program.

8.2 Factors that Predispose P. frosti to being Sensitive to Environmental Change

Information recorded on the habitat preferences and moisture and temperature tolerances of P. frosti during sheltering, movement and breeding activities infers that the species is likely to be sensitive to natural and anthropogenic influences. Notwithstanding the recent decline and range contraction of the species, P. frosti and other members of the genus Philoria in northern New South Wales and southern Queensland, are considered to be naturally rare due to their restricted distributions and specific habitat requirements (Malone 1985a; Hines et al. 1999). In addition to the relictual status of members of the genus (see 1.3.1), habitats upon which P. frosti is dependent, such as cool temperate rainforest and sub-alpine wet heathland, are also considered to be relictual (AHC 1994; Peel 1999). Montane vegetation communities, such as those used by P. frosti, are also noted for their sensitivity to disturbance, and prolonged periods of recovery (Curtin 1995).

The rarity of P. frosti exacerbates its risk of extinction through stochastic variability (environmental, demographic and genetic) and catastrophes (e.g., Hunter 1996; Shaffer 1997). This study has demonstrated that P. frosti is confined to a very narrow range of ecological conditions, and together with its decline and contraction in range, is particularly vulnerable to natural or anthropogenic disturbances. Primack (1993) identifies categories of species that are

245 more vulnerable to extinction than others, and therefore require careful monitoring and management in conservation efforts. Of 14 categories identified, P. frosti can be assigned in part, or entirely to at least seven based on knowledge gained in this thesis and previous studies.

• Species with a very narrow geographical range The confinement of P. frosti to a small area on the Baw Baw Plateau and escarpment between 960 and 1564 m elevation (134.5 km2 potential habitat, 93.4 km2 encompassing extant and historic records - Chapter 3) makes the species one of the most restricted amphibians in Australia.

• Species in which population size is small Relative to populations of other amphibians, whose populations extend over significantly greater areas, P. frosti can be considered to have a small population size (approximately 7000 adult males), particularly given its recent population decline and contraction in range (Chapter 3).

• Species with low rates of population increase The potential for P. frosti to increase population size relative to other amphibians is reduced. The species has low fecundity, producing only 50 – 185 eggs per clutch (Littlejohn 1963; Malone 1985a, b), and recruitment to the terrestrial stage has been estimated to be 8.1% (Malone 1985b). The longevity of P. frosti (~14.5+ years), and prolonged time taken to reach sexual maturity (3.5 years for males and 4.5 – 5.5 years for females) (Chapter 7) relative to other amphibians, allude to a breeding strategy that has evolved to recruit in a gradual rather than explosive manner.

• Species that are not effective dispersers Investigation into the movement patterns of adult P. frosti in this study showed that they are relatively sedentary, remaining within 82 m distance of initial capture localities at breeding sites (mean = 20 m), and average home range size being 135.4 m2 (MCP) or 45.7 m2 (fixed kernel) for male frogs (Chapter 6). The dispersal capability of juvenile and sub-adult P. frosti remains unknown.

• Species with specialised niche requirements Breeding and non-breeding activity by P. frosti is restricted to macro and micro-habitats that offer specific biophysical and climatic conditions (Chapters 4, 5 and 6). For example, sheltering sites chosen by frogs were shown to be within a narrow temperature range compared to prevailing ambient temperature conditions. The extent and probability of movement by frogs was also shown to be influenced by weather conditions, also being constrained to a relatively narrow range of climatic conditions (Chapter 6). Reliance by P. frosti on conditions of low temperature and high relative humidity for optimal movement suggests that habitat niches utilised by the species may play an important role in extending opportunities for movement and dispersal during periods of less optimal weather. The unusual life history of P. frosti also reflects the specialised niche requirements of the species, where unpigmented embryos are deposited in natural cavities away

246 from sunlight, and non-feeding larvae develop to the terrestrial stage primarily at the site of oviposition.

• Species that are characteristically found in stable environments Habitat utilised by P. frosti, particularly for breeding purposes, can be considered to be environmentally stable. Cool temperate rainforest and cool temperate mixed forest are noteworthy for their low level of variability in rainfall, and in combination with topographic protection, ensure a fire frequency and intensity that is lower than most other localities in the landscape (Peel 1999). The Baw Baw Plateau is also noted for its high rainfall in the region (Chapter 2), and for the presence of saturated atmospheric conditions (Abbs and Jensen 1993). The relictual nature of cool temperature rainforest and sub-alpine wet heathland communities (noted above) also implies long- term ecological stability. The use of these vegetation types for breeding purposes, and the narrow climatic window within which breeding activity takes place, emphasises the importance of environmental stability for successful breeding and subsequent embryonic and larval development in P. frosti.

• Species that form permanent or temporary aggregations Species that form permanent or temporary aggregations are more at risk from natural stochastic events or human disturbances (Primack 1993). The metapopulation dynamics of P. frosti are currently unknown. However, monitoring surveys and radio-tracking of adult frogs in this study suggest that aggregations of breeding adults are formed during the breeding season, and to a lesser extent, in the non-breeding season. Adult male and female frogs were recorded as being relatively sedentary during radio-tracking, remaining in proximity of breeding sites following the breeding season (Chapter 6). Although not examined in this study, the recorded distribution of calling males along survey transects was not uniform, with nodes, or aggregations of males typically present. I consider this aggregative pattern of males during the breeding season to be due to social interactions between conspecifics and the natural distribution of suitable breeding habitat in the landscape. The whereabouts of juvenile and sub-adult P. frosti during the breeding and non- breeding seasons are mostly unknown, although recently metamorphosed frogs, and a small number of juvenile and sub-adult frogs, are know to occupy breeding sites (G. Hollis pers. obs.; M. Littlejohn, J. Coventry pers. comm.).

8.2.1 Headwater, Relictual Amphibians

The ecological requirements of P. frosti recorded from this, and previous studies, show that the species is similar to a number of endemic amphibians in north-west California and Oregon that have been the focus of recent studies due to their perceived sensitivity to land management practices. The species Rhyacotriton variegatus, Rhyacotriton olympicus and Plethodon elongatus (Caudata), and Ascaphus truei (Anura) have been described as relictual, headwater amphibians, due to their general preference for cold, clear headwater drainage systems comprising loose, coarse

247 substrates, in stabilised, late seral stage forests (e.g., Corn and Bury 1989; Welsh 1990; Welsh and Lind 1995, 1996, 2002; Diller and Wallace 1999; Wilkins and Peterson 2000; Adams and Bury 2002; Mills and Bury 2002). By causing habitat fragmentation, and altering micro-habitat and micro-climatic conditions, inappropriate timber harvesting regimes are thought be a considerable threat to these sensitive amphibians (Bury and Corn 1988; Welsh 1990; Welsh and Lind 1996; Adams and Bury 2002). Like the North American amphibians above, P. frosti may also be sensitive to disturbances such as that which might occur during forestry activities (see 8.4 and 8.5).

8.3 The Decline of P. frosti and Potential Causative Agents

Unlike some amphibian populations, where the causal agents responsible for their decline have been linked to anthropogenic activities (e.g., Corn and Bury 1989; Davidson et al. 2001; Gillespie and Hollis 1996; Gillespie 2001a, 2002), the overall decline of P. frosti has yet to be directly related to anthropogenic disturbances. Except for a number of historical and more recent incursions into the habitat of P. frosti (Mt Baw Baw Alpine Resort, walking and ski tracks, and some historical (1930 – 1950) and recent clearfell timber harvesting in State Forest), a large portion of the habitat of the species remains in a relatively undisturbed condition. Nevertheless, the disappearance of P. frosti from formerly known breeding sites within a small area comprising ski runs and building infrastructure in the Mt Baw Baw Alpine Resort can be attributed to clearance, modification and fragmentation of breeding and non-breeding habitat. However, this disturbance does not explain the overall decline of P. frosti above 1300 m. Global phenomena have therefore been implicated as potential agents responsible for the decline of P. frosti in the absence of obvious anthropogenic disturbances over the broader distribution of the species.

Below, I further discuss potential agents of decline for P. frosti in relation to factors reviewed by Osborne et al. (1999) for amphibians in the Australian Alps (see 1.2.2.1). The population data on P. frosti presented in this review are from this study, excluding an additional five years of monitoring from 1998 - 2002. Subsequent inferences made about the decline of P. frosti were made by myself in collaboration with the other authors, and with consideration to the nature of declines of other species in the Australian Alps.

8.3.1 Increased UV-B Radiation

A greater intensity of UV-B radiation recorded in sub-alpine habitat compared with lower elevation, montane habitats utilised by P. frosti suggests that increased UV-B may in some way be linked to the decline of the species (see 1.2.2.1.1). However, given that P. frosti has seemingly evolved a breeding strategy that mostly avoids exposure to UV-B (possessing unpigmented

248 embryos and larvae), and that the post-metamorphic life-stage of the species appears to be mostly subterranean (Chapter 6), it is difficult to envisage how increased UV-B radiation might be directly related to the decline of the species. Examination of the breeding macro and micro-habitat of P. frosti (Chapter 4) also demonstrates preference for habitats that are almost entirely protected from sunlight. A number of observed cases where egg masses of P. frosti were located in semi- exposed situations, particularly at sites constructed of vegetation in sub-alpine wet heathland (B. Malone pers. comm., G. Hollis pers. obs.), still does not explain the extent of decline of the species.

If increased UV-B radiation is involved in the decline of P. frosti, it is more likely that its interaction with other stressors, or agents, is a more plausible explanation. For example, Kiesecker and Blaustein (1995) demonstrated increased embryonic mortality as a result of a synergism between UV-B radiation and a pathogen. Synergisms between UV-B and low pH values (Long et al. 1995; Pahkala et al. 2002), and UV-B, low pH and high nitrate levels (Hatch and Blaustein 2000), have also been shown to significantly reduce survival of embryos and larvae, respectively. This study showed that relatively low soil pH values (mean = 5.7, range = 4.9 – 7.2) were recorded at breeding localities currently used by P. frosti, and that the species demonstrated preference for sites with more neutral values of pH at lower elevation (Chapter 4). Because intensities of UV-B radiation reaching the ground are greater within sub-alpine habitats, the resulting effect of synergism between low soil pH and UV-B requires investigation with respect to the species decline.

8.3.2 Pathogens

Of the pathogens reported to have been associated with amphibian declines, amphibian chytridiomycosis provides the most plausible link so far (Daszak et al. 2003). Among amphibians reported by Speare and Berger (2000) to have been infected by Batrachochytrium on the east coast of Australia, three species (Litoria spenceri, Pseudophryne corroboree and Pseudophryne pengilleyi) occur in the Australian Alps, and all have suffered population declines (see Osborne et al. 1999; Gillespie 2001b). Skeletochronological examination of tissue samples from captured specimens of P. frosti between 1994 and 2000 (Chapter 7) show that the age-class distribution of adult frogs was generally skewed towards the presence of younger individuals (Figs 7.2b and c). If these tissue samples are representative of the extant population of adult P. frosti, this age-class distribution suggests that the agent(s) responsible for the decline of the species may have impacted negatively on individuals pertaining to the post-metamorphic life-stage of the population. To date, no investigations have been undertaken to determine the presence or absence of Batrachochytrium in P. frosti. Given that Batrachochytrium is known to be virulent to the post-metamorphic life- stage of amphibians, and not eggs and larvae (Berger et al. 1999), the fungus should be considered a possible agent in the decline of P. frosti. The presence of the fungus in high-elevation

249 amphibians, and its preference for cooler temperatures, provides further support for Batrachochytrium as a threat to P. frosti (Berger et al. 1999; Daszak et al. 2003).

8.3.3 Climate Change

The impact of climate change is particularly relevant to P. frosti, given that it is geographically restricted, confined to sub-alpine and montane habitats, and has specialised life-history characteristics. Bennett et al. (1991) predicted that the bioclimate of P. frosti would disappear with a rise in temperature of 1 - 3 oC. Additional support for the suggestion that P. frosti is sensitive to climate change is provided by results from this study, including: (1) the species apparent climate- driven pattern of distribution and population density (Chapter 3); (2) the species current decline and contraction in range from sub-alpine habitat (Chapter 3); (3) the species narrow moisture and temperature tolerances recorded during activities relating to sheltering, movement and breeding at sub-alpine and montane elevations (Chapters 5 and 6); and (4) the species current use of a smaller subset of breeding habitats, characterised by topographically protected, cool, moist environments, compared with those used historically (Chapter 4).

Although studies have shown that there have been warming trends in climate in the general vicinity of P. frosti (see 1.2.2.1.1), the pattern of decline of the species from higher, sub-alpine habitats, and persistence of a larger population in lower, montane habitats, is in contrast to an upward contraction in range that might be expected under a warming scenario. Temperature conditions (ambient and substratum) measured at a sub-alpine (1485 m) and montane (1090 m) breeding site in this study, however, showed that the montane site was on average cooler than the sub-alpine site over the breeding season (Chapter 5). This unexpected difference in temperature was considered to be due to the significantly greater solar insulation provided by canopy and understorey vegetation at the montane site (comprising cool temperate rainforest), and that the altitudinal range over which P. frosti occurs is relatively narrow (606 m, range = 960 – 1566 m). If this pattern in temperature is generally consistent between sub-alpine and montane habitats during the activity period of P. frosti, and average temperatures have increased in the vicinity of the Baw Baw Plateau to the detriment of the species, the contraction of the species to lower, cooler, montane habitats could be considered an expected response to warming temperatures. Under this scenario, vegetation types such as cool temperate rainforest and cool temperate mixed forest, which predominate on the south-western escarpment of the Baw Baw Plateau, could be refugial habitats for P. frosti. The observation of breeding activity starting approximately four weeks earlier than previously recorded for P. frosti in this study might also suggest a response by the species to warming temperatures, as has been reported in other amphibians recently (Chapter 5).

Examination of rainfall data from Erica (near the Baw Baw Plateau) during the calling activity period for P. frosti showed that the high levels of rainfall received between 1932 and 1962 have not occurred since, and that the number of days with less than 5 mm rainfall have increased

250 significantly from 1959 to 1998 (Smith et al. 1999). The narrow climatic preferences of P. frosti indicate that the species may be influenced by even the most subtle climatic changes. Climate change has also been implicated in the decline of Pseudophryne corroboree in the Australian Alps (Osborne 1990; Osborne et al. 1999). Osborne et al. (1999) summarise the decline of P. corroboree with respect to the distribution of rainfall and elevation, indicating a progressive decline of the species from the drier, lower-elevation areas to the wetter, higher-elevation areas of its distribution. The decline of Litoria verreauxii alpina and Pseudophryne pengilleyi has also been greater at higher elevation elsewhere in the Australian Alps (Osborne et al. 1999). The situation is somewhat similar for P. frosti, where populations have also disappeared from areas at the drier, eastern edge if its range, but have persisted in wetter areas at the western extent of its range. However, with respect to elevation, P. frosti appears to have declined more at higher elevation than lower elevation (Chapter 3) compared with P. corroboree.

Changes in climate due to factors operating at a regional, or catchment level, may also explain the decline of P. frosti. Helps (2001) reports that the long-term downward trend in total annual rainfall, and smoothing of fluctuations in annual rainfall, at Erica and Noojee (also noted generally by Smith et al. 1999 above) may be as a result of the construction of the Thomson Reservoir in 1982. Research in China shows that large volumes of water in mountainous areas, like the Thomson Reservoir, act as a temperature moderator, altering rainfall patterns due to changes in temperature range (Helps 2001). Filling of the Thomson reservoir in 1989, and subsequent downward trend in rainfall and smoothing of peaks of rainfall, correlate with the timing in decline and contraction in range of P. frosti after the surveys of Malone (1985a) in 1983 and 1984. Observations by Malone (1985b) that desiccation was the primary cause of reduced embryonic and larval survivorship in P. frosti supports the hypothesis that reduced rainfall resulting from the construction of the Thomson Reservoir may be involved with the decline and contraction in range of the species. It is also interesting to note the decline of P. corroboree, L. v. alpina and P. pengilleyi in the Snowy Mountains, where there are also numerous water empoundments that were established in the 1950s and 1960s. The influence of global and regional climate change requires further investigation with respect to the decline and ecology of P. frosti.

8.3.4 Natural Population Fluctuations and Weather Patterns

Long-term population monitoring of amphibians, and an understanding of their demography, is required to separate apparent declines from natural population fluctuations (Pechmann and Wilbur 1994). With respect to the longevity and maturation of P. frosti (Chapter 7), I consider that sufficient population data have been collected and examined in this study to ascertain population trends (see 8.1). With an additional five years of survey data (1998 – 2002) examined in this study compared to data reported on previously by Osborne et al. (1999), I continue to support the conclusion of Osborne et al. (1999) that the decline of P. frosti is not due to natural population fluctuations in response to normal patterns of weather. There appears to be no general relationship

251 between annual relative abundance indices and weather to suggest that the dramatic decline of the species could be interpreted as natural fluctuations in population numbers. Further support is given to this consensus in the biogeographical sense, where natural population fluctuations in response to weather conditions have also been excluded as a possible explanation for the decline of other amphibians in the Australian Alps (Osborne et al. 1999).

8.3.6 Atmospheric Pollution

Atmospheric pollution has not as yet been investigated as a potential agent in the decline of P. frosti. Atmospheric studies suggest that pollutants originating from industrial activities in the Latrobe Valley are not contributing to acidification (1.2.2.2). If the pollutants were a threat to P. frosti, prevailing west to south-westerly winds in the vicinity of the Latrobe Valley suggest that any pollutants would be generally directed away from the Baw Baw Plateau. However, the Baw Baw Plateau is situated within another pollution corridor, which originates in the Melbourne area and extends to the north-east of Victoria, where urban and industrial air pollution have been implicated as a suppressant of precipitation by influencing cloud properties (see Rosenfeld 2000).

Forest edges resulting from land management practices may also function as traps and concentrators for wind-borne nutrients and pollutants (Weathers et al. 2001). The high intensity of timber harvesting around the periphery of the Baw Baw Plateau over the past 15 years has resulted in multiple forest edges, although most harvesting to date has occurred below the distributional limit of P. frosti. The influence of potential atmospheric pollutants on P. frosti in association with its ecological requirements, requires further investigation.

8.3.7 Multiple and Interacting Factors

Investigations into the decline of amphibian populations in the past have focused primarily on the direct effects of single factors (Blaustein and Kiesecker 2002). However, more recent studies investigating multiple factors, and their interactions (e.g., Carey 1993; Pounds and Crump 1994; Kiesecker and Blaustein 1995; Long et al. 1995; Moore and Klerks 1998; Alexander and Eischeid 2001), increasingly support the consensus that amphibian losses are the result of interactions between a number of highly context-dependent causal factors (see review by Blaustein and Kiesecker 2002). Collins and Storfer (2003) also note that amphibian declines are likely to be due to complex interactions of several anthropogenic factors, separating decline agents into two groups: alien species, over-exploitation and land use change (class I hypotheses); and global change, contaminants and diseases (class II hypotheses). The decline of P. frosti may also be as a result of multiple factors acting together, or synergisms between factors, some of which have been noted above. With respect to knowledge gained on the distribution, abundance and ecological

252 requirements of P. frosti from this and previous studies, the following multiple, and interacting factors, require further investigation with regards to the ecology of the species and its decline:

1. Multiple factors resulting in reduced precipitation, including global and regional climate change (see 8.3.3) and atmospheric pollution (see 8.3.6); 2. Synergism between UV-B radiation and other factors including low pH (e.g., Long et al. 1995; Pahkala et al. 2002), low pH and high nitrate levels (e.g., Hatch and Blaustein 2000), pathogens (e.g., Kiesecker and Blaustein 1995) and pollutants (e.g., Long et al. 1995; Monson et al. 1999; Blaustein et al. 2003); and 3. Synergism between climate change and pathogens (e.g., Lips 1998; Pounds et al. 1999; Harvell et al. 2002; Carey and Alexander 2003) and high temperature and low pH (Moore and Klerks 1998).

8.4 Implications for Conservation

This study has confirmed a decline and contraction in range of P. frosti, obtained fundamental information on autoecology and demography of the species to assess the context of the decline, and reviewed conceivable agents of decline. Having addressed four of the five disciplines of Endangered Species Recovery Analysis (see Caughley and Dunn 1996), a platform has now been established to commence diagnosis of the decline of P. frosti, involving the development and testing of hypotheses for identified decline agents, as well as to investigate for potential impacts of land management on the species (see 8.5).

Components of this thesis have primarily addressed principles from the declining-population paradigm in conservation biology (see 1.1), focusing on the detection and confirmation of the decline of P. frosti. Other components have examined the autoecology of the species, alluding to the potential vulnerability of the species to environmental stochasticity (small-population paradigm). Although P. frosti has suffered a significant decline and contraction in range from sub- alpine elevations, and a possible decline from newly-discovered populations at montane elevations, I consider that this decline has been detected before the critical stage of ‘low population numbers’. Conservation agencies and land managers are therefore in a position of dealing with downward trends in numbers, rather than critically small numbers that can potentially change conservation management into crisis management (Caughley and Dunn 1996). However, given the extent of decline from sub-alpine elevations, there is still a degree of urgency in identifying the decline agent(s) and implementing appropriate management. Recommendations for future research on P. frosti (see 8.5.3 below) continue to draw on principles from both conservation biology paradigms, which have much to contribute to each other in the discipline of conservation biology (Caughley 1994; Caughley and Dunn 1996).

253 Knowledge of the distribution, abundance and ecological requirements of P. frosti gained from this study also provides the impetus to consider the adequacy of existing conservation measures for the species. Based on the estimate of potential habitat area for P. frosti (134.5 km2, Chapter 3), approximately 56% of this area occurs within the Baw Baw National Park (75.4 km2), which is managed primarily for conservation purposes. A further 42% of the distribution of P. frosti occurs in State Forest (56.1 km2), primarily on the south-western escarpment of the Baw Baw Plateau down to 960 m elevation, with only a small contribution from the north-eastern escarpment above 1200 m. Approximately 3% of the distribution of P. frosti occurs within the Mt Baw Baw Alpine Resort (4.2 km2) where approximately 15% of the total area has been developed for recreational activities and supporting infrastructure. This study’s estimate of adult males above 1300 m elevation (Chapter 3) indicates that approximately 35% of the total population of P. frosti occurs on land managed within the Baw Baw National Park. Approximately 65% of the total population of P. frosti occurs in State Forest, with almost all of this estimate occurring on the south-western escarpment of the plateau between 960 and 1299 m (Chapter 3). Less than 10% of the total population estimated to occur in State Forest and National Park occurs on land managed within the alpine resort.

Prior to this study's discovery of substantial populations of P. frosti in State Forest, the distribution of the species was generally considered to be confined to reserve in the Baw Baw National Park (Chapter 2). However, with approximately two-thirds of the extant population of P. frosti now believed to occur in State Forest, out of a reserve, and the remaining one-third in the National Park continuing to decline, conservation agencies and land managers are now faced with the challenge of diagnosing and reversing the decline of the species, as well as managing forestry activities in State Forest and recreational activities in the ski resort and National Park. Unlike species where the agents of decline have been identified as tangible entities, such as habitat loss, introduced predators or overexploitation, the agents responsible for the decline of P. frosti appear not to be tangible (e.g., climate change or an introduced pathogen), and will require additional thought by conservation agencies when considering a conservation strategy for P. frosti. A clear preference by P. frosti for wetter, cooler, habitats on the south-western escarpment of the Baw Baw Plateau emphasises the refugial nature and importance of this region in the future management and conservation of the species. If non-tangible agents are diagnosed as being responsible for the decline of P. frosti, and there is no short-term solution for treating the decline (e.g., under a climate change scenario), the relevance of the south-western escarpment as a refuge for the future conservation of the species will increase in importance.

Due to a lack of information about the potential impacts of timber harvesting on the survival prospects of P. frosti, there is currently no recognised strategy to manage for both timber values and the ecological requirements of the frog in State Forest. The uncertainty associated with this dilemma is a common problem confronted by managers in the field of conservation management (Caughley 1994; Caughley and Gunn 1996; Meffe and Carroll 1997). Managers can be placed in a

254 position of having to take quick action to reduce ecological risk arising from a hazard before all information is available, or when the quality of information is suspect (Meffe and Carroll 1997). Risk and decision analysis is a technique that allows all the criteria for a decision to be disclosed and objectively evaluated (e.g., Caughley and Gunn 1996; Maguire 1997; Meffe and Carroll 1997). Risk analysis involves estimating the likelihood that an identified hazard will have a negative effect, and estimating the ecological consequences of that negative effect (Meffe and Carroll 1997). The information gathered from the risk analysis is then integrated into a decision analysis process (e.g., Maguire 1997). Managers faced with developing a management strategy for P. frosti in State Forest may benefit from undertaking such an analysis.

If the result of risk and decision analyses determines that forestry activities are to be undertaken within and adjacent to the habitat of P. frosti, incorporation of adaptive management principles into forestry operations is likely to be beneficial in achieving a balanced result between resource use whilst ensuring the long-term survival prospects of P. frosti (see 8.5). Adaptive management is the systematic acquisition and application of reliable information to improve natural resource management over time (e.g., Walters 1986; Salafsky et al. 2001; Wilhere 2002). The approach aims to reduce uncertainty underlying ecological relations to environmental variation by monitoring biotic responses to management actions, and by comparing responses to predictions generated by alternative hypotheses (Irwin and Freeman 2001).

8.5 Recommendations for Management and Research

This study has contributed to the completion of research recommendations identified in the Action Statement (CNR 1993) and National Recovery Plan (Hollis 1997) for P. frosti. The findings of this research have been used to commence development of a management strategy for P. frosti in State Forest, and for the revision of research actions in a revised Recovery Plan (Hollis 2002).

8.5.1 Management in State Forest

The sensitivity of P. frosti to habitat disturbance suggests that forestry activities may impact directly or indirectly on the long-term survivorship prospects of the species. This impact may occur through: (1) direct destruction of frogs and habitat; (2) changes to climatic and hydrological conditions from activities in and adjacent to frog habitat; (3) sedimentation of breeding habitat following activities in and adjacent to frog habitat; and (4) fragmentation of populations, and/or destruction or modification of dispersal corridors. Intensive timber harvesting in forest management blocks on the north-eastern and south-western escarpments of the Baw Baw Plateau over the past 20 years, including a number of areas within the potential habitat of P. frosti

255 (Department of Sustainability and Environment logging history data), may have also impacted on the population of P. frosti. A study of forest rotation and stream-flow benefits in the water catchment for the Thomson Reservoir showed that stream-flow water yields are reduced significantly in regenerating forest (through increased rates of transpiration) up to age of 25 years following timber harvesting, and thereafter decreases with age (Clarke 1994). This observation suggests that the hydrological and climatic requirements of P. frosti could be influenced by substantial areas of re-growth forest following timber harvesting. Areas of re-growth forest, of an even age and uniform type, following timber harvesting may also be detrimental to the habitat requirements of P. frosti. Logs and other coarse woody debris, which are utilised by P. frosti for breeding and shelter (Chapter 4 and 6), may become limited in re-growth forest as they pass beyond favourable decay classes (e.g., Davis 2002).

The initial discovery of P. frosti in State Forest areas on the south-western and north-eastern escarpments of the Baw Baw Plateau in 1996 led to the development of an interim guideline to manage forestry activities in State Forest in the species habitat (NRE 1998). Due to the infancy of the Recovery Plan for P. frosti (Hollis 1997) at the time, together with a lack of information on distribution and habitat of the species in State Forest, the management guideline was developed to be both precautionary, and of an interim nature. This precautionary approach allowed time to collect further information on breeding habitat and key ecological requirements of the species whilst excluding most forestry activities from within its habitats (see NRE 1998).

In 2000, personnel from the then Department of Natural Resources and Environment (DNRE) commenced to revise the above interim management guideline using the findings of research from this study. Newly acquired information on the distribution of breeding habitat of P. frosti in State Forest, based on a digital elevation model (Liu and White 1999) and refinement of the 1:25000 hydrological layer (internal Department of Sustainability and Environment mapping), was also used to assist this process. A number of management strategy options, designed to both protect areas containing habitat for P. frosti, and allow forestry activities within its habitat, were investigated. However, a lack of information about the response of the species to habitat disturbance and fragmentation in addition to the species' endangered status and predicted sensitivity to disturbance (this study), hampered progress towards adopting a strategy that was satisfactory to timber industry and conservation interests. In 2003, the basis for a revised conservation and management strategy for P. frosti was:

1. the establishment of a management zone encompassing the potential habitat envelope of the species, based on the distribution of known frog localities and modelling of breeding habitat (Liu and White 1999). 2. the exclusion of areas and volumes of timber within the management zone from the forestry industry for the next license period;

256 3. the undertaking of experimental timber harvesting (adaptive management) within the management zone to determine the impact of timber harvesting on populations of P. frosti and its ecological requirements; and 4. modification of forestry practices to accommodate the findings of experimental timber harvesting.

8.5.2 Management in the Baw Baw National Park and Mt Baw Baw Alpine Resort

A number of management and conservation initiatives contained in the Recovery Plan for P. frosti (Hollis 1997) remain as ongoing actions in the Baw Baw National Park and Mt Baw Baw Alpine Resort. These actions pertain to the control, eradication and monitoring of introduced plants and animals, protection of sensitive vegetation communities, rehabilitation of previously disturbed areas, and the general maintenance of skiing/walking tracks and infrastructure used for recreational activities.

8.6 Priority Research Directions

Based on the findings of this study, it is recommended that the research actions listed below be considered a priority for incorporation into a revised Recovery Plan for the species. I consider that these research actions are currently the most appropriate to further address the decline and contraction in range of the species, and to assess the potential impact of proposed land use activities within the species habitat.

Population Monitoring

A statistically-defendable, efficient long-term monitoring program is required to distinguish natural population fluctuations from real declines in amphibians (Chapter 3). Population monitoring is also required to investigate potentially adverse agents in the decline of P. frosti, as well as to measure the effectiveness of implemented management actions or programs involving adaptive management (Clark et al. 2002). Evaluation of the monitoring program conducted in this study, and by others (M. Scroggie, unpublished data), has provided guidance as to the effectiveness of the survey methodology. The survey monitoring program for P. frosti has to date yielded valuable information on the distribution, habitat use and abundance of the species over time. Its continuation is considered vital for future studies investigating the species decline and sensitivity to land management practices (see below).

257 Chytrid Fungus (Batrachochytrium dendrobatidis)

The introduced Chytrid Fungus (Batrachochytrium dendrobatidis) is now well established in a number Australian amphibian populations, and has been implicated as a proximate cause of the recent decline of some amphibians (Berger et al. 1998). Given the detection of this pathogen in other amphibians within the Australian Alps, it is possible that this pathogen could explain the decline and contraction in range of P. frosti following the survey of Malone (1985a) in 1983 and 1984. Archived specimens of P. frosti in the Museum of Victoria and tissue samples taken from live specimens during this study can be used to determine the presence of B. dendrobatidis in P. frosti, and the timing of introduction.

Impact of Forestry Activities

The narrow ecological requirements of P. frosti suggests that forestry activities may have a potentially adverse effect on the habitat and biophysical preferences of the species (see 8.4). Research is required to determine the potential impact of forestry activities on P. frosti and its habitat to ensure that they will not impinge upon the long-term survival of the species (see 8.5.1 for further details).

Conservation Genetics

Knowledge of the movement patterns and potential for dispersal by adult P. frosti was examined in this study (Chapter 6). However, movement patterns of adults may be a poor basis for predicting the behaviour of dispersing juveniles (e.g., Rothermel and Semlitsch 2002). The potential for movement by non-adult age classes (juveniles and sub-adults) of P. frosti, and overall dispersal potential, remains unknown. Knowledge of overall dispersal potential is required to gain an understanding of metapopulation dynamics of the species. An understanding of metapopulations is considered necessary to effectively assess population declines (Alford and Richards 1999), as well as providing useful information for interpreting research investigating the potential impact of land use activities on the species. Due to the rarity, cryptic nature and endangered conservation status of P. frosti, conservation genetic analysis (e.g., see Steinberg and Jordan 1998) appears to be the most feasible method available to determine the overall dispersal potential of the species. Such an analysis would provide: (1) an ecologically relevant estimate of dispersal rates, or effective neighbourhood size; (2) an estimation of the metapopulation structure; (3) insight into potential sex-bias in dispersal rates; and (4) insight into effective population size.

Population Modelling

Modelling has become an important tool for guiding management and conservation planning (Lacy 1993; Shebitz 2002), and is widely used for setting conservation policy (Ellner et al. 2002).

258 Analysis of the viability of populations, commonly referred to as population viability analysis (PVA), typically involves an assessment of a population’s risk of extinction, or its projected growth under current conditions, or conditions resulting from proposed management (Reed et al. 2002). There is now sufficient information on the ecology, population structure and life history of P. frosti to undertake population modelling to investigate its prospect of long-term survivorship under different management scenarios (e.g., Meir and Kareiva 1998). Population modelling is required to examine the probability of survival of the species under the current management regime, where a dramatic decline in population from higher elevation, sub-alpine habitats has occurred, as well as under changed management regimes, particularly examining the potential impact of forestry practices on the species.

Weather and Climate Change

Climate change may be a potential threat to the long-term survivorship of P. frosti, and may be responsible for the decline of the species (see 8.3.3). The results of this study clearly emphasis the importance of the wetter, south-western escarpment of the Baw Baw Plateau as a current refuge for the species. One difficulty in assessing the influence of climate change on populations is separating short-term fluctuations, due to natural weather conditions, from longer-term trends. New, dynamic conservation strategies for biodiversity are required to accommodate changes in global climate (Hannah et al. 2002). For P. frosti, further research is required to understand the influence of climate change on the species, and of the influence of weather on population fluctuations over time, before strategies can be developed to accommodate climate change.

259

References

Abbs D.J. and Jensen J.B. (1993). Numerical modeling of orographically forced postfrontal rain. Monthly Weather Review, 121: 189-206.

Acker P.M., Kruse K.C. and Krehbiel E.B. (1986). Aging Bufo americanus by skeletochronology. Journal of Herpetology, 20: 570-574.

Adams M.J. and Bury R.B. (2002). The endemic headwater stream amphibians of the American Northwest: associations with environmental gradients in a large forested preserve. Global Ecology and Biogeography, 11: 169-178.

AHC (1994). National Estate Values in the Central Highlands of Victoria: Draft Project Report. Department of Conservation and Natural Resources and Australian Heritage Commission, Australian Capital Territory.

Ahern L.D. (1982). Threatened wildlife in Victoria and issues related to its conservation. Fisheries Wildlife Paper, 27: 34

Alcobendas M. and Castanet J. (2000). Bone growth plasticity among populations of Salamandra salamandra: interactions between internal and external factors. Herpetologica, 56: 14-26.

Aldrick J.M., Hook R.A., van de Graaff R.H.M., Nicholson B.M., O'Beirne D.A. and Schoknecht N.R. (1992). A Study of the Land in the Catchment of the Gippsland Lakes. Volume 2, Department of Conservation and Natural Resources, Land and Catchment Protection Branch, Victoria, Australia.

Alexander M.A. and Eischeid J.K. (2001). Climate variability in regions of amphibian declines. Conservation Biology, 15: 930-942.

Alford R.A. (1999). Tadpole ecology: resource use, competition and predation. In: McDiarmid R.W. and Altig R., (Eds.), Biology of Anuran Larvae. University of Chicago Press, Chicago, pp. 240-278.

Alford R.A., Dixon P.M. and Pechmann J.H.K. (2001). Global amphibian declines. Nature, 414: 449-500.

Alford R.A. and Richards S.J. (1997). Lack of evidence for epidemic disease as an agent in the catastrophic decline of Australian rain forest frogs. Conservation Biology, 11: 1026-1029.

Alford R.A. and Richards S.J. (1999). Global amphibian declines: a problem in applied ecology. Annual Review of Ecology and Systematics, 30: 133-165.

261 Ankley G.T., Diamond S.A., Tietge J.E., Holcombe G.W., Jensen K.M., Defoe D.L. and Peterson R. (2002). Assessment of the risk of solar ultraviolet radiation to amphibians. I. dose- dependent induction of hindlimb malformations in the northern leopard frog (Rana pipiens). Environmental Science and Technology, 36: 2853-2858.

Anstis M. (1981). Breeding biology and range extension for the New South Wales frog Kyarranus sphagnicolus (Anura: Leptodactylidae). Australian Journal of Zoology, 1: 1-9.

Anstis M. (2002). Tadpoles of South-eastern Australia: a Guide with Keys. Reed New Holland, Australia.

Anstis M., Alford R.A. and Gillespie G.R. (1998). Breeding biology of Litoria booroolongensis (Moore, 1961) and Litoria Lesueuri (Dumeril and Bibron, 1841) (Anura: Hylidae) and comments on population declines of L. booroolongensis. Transactions of the Royal Society of South Australia, 122: 33-43.

Anthony L.L. and Blumstein D.T. (2000). Integrating behaviour into wildlife conservation: the

multiple ways that behaviour can reduce Ne. Biological Conservation, 95: 303-315.

Anzalone C.R., Kats L.B. and Gordon M.S. (1998). Effects of solar UV-B radiation on embryonic development in Hyla cadaverina, Hyla regilla, and Taricha torosa. Conservation Biology, 12: 646-653.

ANZECC (1991). Australian and New Zealand Environment and Conservation Council. List of Endangered Vertebrate Fauna, Australian National Parks and Wildlife Service, Canberra.

ANZECC (1995). Australian and New Zealand Environment and Conservation Council. List of Endangered Vertebrate Fauna, Australian Nature Conservation Agency, Canberra.

Babik W. and Rafiński J. (2001). Amphibian breeding site characteristics in the western Carpathians, Poland. Herpetological Journal, 11: 41-51.

Balkau F. (1982). Map of the Baw Baw Plateau 1:25 000. Snowgum Press. Carlton, Victoria.

Balkau F. (1987). Baw Baw Plateau 1:25 000. Revised edition. Snowgum Press. Carlton, Victoria.

Baringa M. (1990). Where have all the froggies gone? Science, 247: 1033-1034.

Barker J., Grigg G.C. and Tyler M.J. (1995). A Field Guide to Australian Frogs. Surrey Beatty and Sons, Sydney.

Bastien H. and Leclair Jr., R. (1992). Aging wood frogs (Rana sylvatica) by skeletochronology. Journal of Herpetology, 26: 222-225.

Bastos R.P. and Haddad C.F.B. (2002). Acoustic and aggressive interactions in Scinax rizibilis

262 (Anura : Hylidae) during the reproductive activity in southeastern Brazil. Amphibia- Reptilia, 23: 97-104.

Beard K.H., McCullough S. and Eschtruth A.K. (2003). Quantitative assessment of habitat preferences for the Puerto Rican terrestrial frog, Eleutherodactylus coqui. Journal of Herpetology, 37: 10-17.

Beebee T.J.C. (1995). Amphibian breeding and climate. Nature, 374: 219-220.

Beebee T.J.C. (1996). Ecology and Conservation of Amphibians. Chapman and Hall, London.

Beebee T.J.C., Flower R.J., Stevenson A.C., Patrick S.T., Appleby P.G., Fletcher C., Marsh C., Natkanski J., Rippey B. and Battarbee R.W. (1990). Decline of the natterjack toad Bufo calamita in Britain: palaeoecological, documentary and experimental evidence for breeding site acidification. Biological Conservation, 53: 1-20.

Belden L.K. and Blaustein A.R. (2002a). Exposure of red-legged frog embryos to ambient UV-B radiation in the field negatively affects larval growth and development. Oecologica, 130: 551-554.

Belden L.K. and Blaustein A.R. (2002b). Population differences in sensitivity to UV-B radiation for larval long-toed salamanders. Ecology, 83: 1586-1590.

Bennett S., Brereton R., Mansergh I., Berwick S., Sandiford K. and Wellington C. (1991). The Potential Effect of the Enhanced Greenhouse Climate Change on Selected Victorian Fauna. Arthur Rylah Institute Technical Report Series No. 123, Department of Conservation and Environment, Victoria.

Berger L., Speare R., Daszak P., Earl Green D., Cunningham A.A., Goggin C.L., Slocombe R., Ragan M.A., Hyatt A.D., McDonald K.R., Hines H.B., Lips K.R., Marantelli G. and Parkes H. (1998). Chytridiomycosis causes amphibian mortality associated with population declines in the rainforests of Australia and Central America. Proceedings of the National Academy of Sciences, 95: 9031-9036.

Berger L., Speare R. and Hyatt A. (1999). Chytrid fungi and amphibian declines: overview, implications and future directions. In: Campbell A., (Ed.), Declines and Disappearances of Australian Frogs. Biodiversity Group, Environment Australia, Canberra, pp. 23-33.

Bertoluci J. (1998). Annual patterns of breeding activity in atlantic rainforest anurans. Journal of Herpetology, 32: 607-611.

Bertoluci J. and Rodrigues M.T. (2002). Seasonal patterns of breeding activity of atlantic rainforest anurans at Boracéia, southeastern Brazil. Amphibia-Reptilia, 23: 161-167.

263 Berven K.A. (1995). Population regulation in the wood frog, Rana sylvatica, from three diverse geographic localities. Australian Journal of Ecology, 20: 385-392.

Bevier C.R. (1997). Utilization of energy substrates during calling activity in tropical frogs. Behavioral Ecology and Sociobiology, 41: 343-352.

Bider J.R. and Morrison A.K. (1981). Changes in toad (Bufo americanus) responses to abiotic factors at the northern limit of their distribution. American Midland Naturalist, 106: 293- 304.

Blaustein A.R. (1994). Chicken little or nero's fiddle? A perspective on declining amphibian populations. Herpetologica, 50: 85-97.

Blaustein A.R., Belden L.K., Olson D.H., Green D.M., Root T.L. and Kiesecker J.M. (2001). Amphibian breeding and climate change. Conservation Biology, 15: 1804-1809.

Blaustein A.R., Chivers D.P., Kats L.B. and Kiesecker J.M. (2000). Effects of ultraviolet radiation on locomotion and orientation in roughskin newts (Taricha granulosa). Ethology, 106: 227-234.

Blaustein A.R., Hays J.B., Hoffman P.D., Chivers D.P., Kiesecker J.M., Leonard W.P., Marco A., Olson D.H., Reaser J.K. and Anthony R.B. (1999). DNA repair and resistance to UV-B radiation in western spotted frogs. Ecological Applications, 9: 100-1105.

Blaustein A. R., Hoffman P.D., Hokit D.G., Kiesecker J.M., Walls S.D. and Hays J.B. (1994b). UV repair and resistance to solar UV-B in amphibian eggs: a link to population declines? Proceedings of the National Academy of Science, USA, 91: 1791-1795.

Blaustein A. R., Hoffman P.D., Kiesecker J.M. and Hays J.B. (1996). DNA repair activity and resistance to solar UV-B radiation in eggs of the red-legged frog. Conservation Biology, 10: 1398-1402.

Blaustein A.R., Hokit D.G. and O'Hara R.K. (1994c). Pathogenic fungus contributes to amphibian losses in the Pacific northwest. Biological Conservation, 67: 251-254.

Blaustein A.R. and Kiesecker J.M. (2002). Complexity in conservation: lessons from the global decline of amphibian populations. Ecology Letters, 5: 597-608.

Blaustein A.R., Kiesecker J.M., Chivers D.P. and Anthony R.G. (1997). Ambient UV-B radiation causes deformities in amphibian embryos. Proceeding of the National Academy of Sciences, USA, 94: 13735-13737.

Blaustein A.R., Romansic J.M. and Kiesecker J.M. (2003). Ultraviolet radiation, toxic chemicals and amphibian population declines. Diversity and Distributions, 9: 123-140.

264 Blaustein A.R. and Wake D.B. (1990). Declining amphibian populations: a global phenomenon? Trends in Ecology and Evolution, 5: 203-204.

Blaustein A.R. and Wake D.B. (1995). The puzzle of declining amphibian populations. Scientific American, April 1995: 56-61.

Blaustein A.R., Wake D.B. and Sousa W.P. (1994a). Amphibian declines: Judging stability, persistence, and susceptibility of populations to local and global extinctions. Conservation Biology, 8: 60-71.

Blaustein A.R. and Walls S.C. (1995). Aggregation and kin recognition. In: Heatwole H. and Sullivan B.K., (Eds.), Amphibian Biology. Vol. 2. Social Behaviour. Chipping Norton, New South Wales, Australia, pp. 568-602.

Blumthaler M. and Ambach W. (1990). Indication of increasing solar ultraviolet-B radiation flux in alpine regions. Science, 248: 206208

Bosch J. and Márquez R. (2001). Call timing in male-male acoustic interactions and female choice in the midwife toad Alytes obstetricans. Copeia, 2001: 169-177.

Bosch J., Martínez-Salano I. and García-París M. (2001). Evidence of a chytrid fungus infection involved in the decline of the midwife toad (Alytes obstetricians) in protected areas of Central Spain (Anura: Discoglossoidae). Biological Conservation , 97: 331-337.

Bosman W., van Gelder J.J. and Strijbosch H. (1996). Hibernation sites of the toads Bufo bufo and Bufo calamita in a river floodplain. Herpetological Journal, 6: 83-86.

Bourne G.R., Collins A.C., Holder A.M. and McCarthy C.L. (2001). Vocal communication and reproductive behavior of the frog Colostethus beebei in Guyana. Journal of Herpetology, 35: 272-281.

Bradford D.F. (1991). Mass mortality and extinction in a high-elevation population of Rana mucosa. Journal of Herpetology, 25: 174-177.

Bradford D.F., Gordon M.S., Johnson D.F., Andrews R.D. and Jennings W.B. (1994). Acid deposition as an unlikely cause for amphibian population declines in the Sierra Nevada, California. Biological Conservation, 69: 155-161.

Bradford D.F., Tabatabai F. and Graber D.M. (1993). Isolation of remaining populations of the native frog, Rana mucosa, by introduced fishes in Sequoia and Kings Canyon National Parks, California. Conservation Biology, 7: 882-888.

Brattstrom B.H. (1979). Amphibian temperature regulation studies in the field and in the laboratory. American Zoologist, 19: 345-356.

265 Braun-Blanquet J. (1928). Pflanzensoziologie. Springer, Berlin.

Bray J.R. and Curtis J.T. (1957). An ordination of the upland forest communities of southern Wisconsin. Ecological Monographs, 27: 325-349.

Brenowitz E.A. and Rose G.J. (1999). Female choice and plasticity of male calling behaviour in the Pacific treefrog. Animal Behaviour, 57: 1337-1342.

Brererton R., Bennett S. and Mansergh I. (1995). Enhanced greenhouse climate change and its potential effect on selected fauna of south-eastern Australia: a trend analysis. Biological Conservation, 72: 339-354.

Bridges A.S. and Dorcas M.E. (2002). Temporal variation in anuran calling behavior: implications for surveys and monitoring programs. Copeia, 2002: 587-592.

Brooke P.N., Alford R.A. and Schwarzkopf L. (2000). Environmental and social factors influence chorusing behaviour in a tropical frog: examining various temporal and spatial scales. Behavioural Ecology and Sociobiology, 49: 79-87.

Brooks T.M., Mittermeier R.A., Mittermeier C.G., da Fonseca G.A.B., Rylands A.B., Konstant W.R., Flick P., Pilgram J., Oldfield S., Magin G. and Hilton-Taylor C. (2002). Habitat loss and extinction in the hotspots of biodiversity. Conservation Biology, 16: 909-923.

Broomhall S.D., Osborne W.S. and Cunningham R.B. (2000). Comparative effects of ambient ultraviolet-B radiation on two sympatric species of Australian frogs. Conservation Biology, 14: 420-427.

Bulger J.B., Scott Jr., N.J. and Seymour R.B. (2003). Terrestrial activity and conservation of adult California red-legged frogs Rana aurora draytonii in coastal forests and grasslands. Biological Conservation, 110: 85-95.

Bull E.L. (2000). Comparison of two radio transmitter attachments on Columbia spotted frogs (Rana luteiventris). Herpetological Review, 31: 26-28.

Bull E.L. and Hayes M.P. (2001). Post-breeding season movements of Columbia spotted frogs (Rana luteiventris) in north-eastern Oregon. Western North American Naturalist, 61: 119- 123.

Bull E.L. and Hayes M.P. (2002). Overwintering of Columbia spotted frogs in northeastern Oregon. Northwest Science, 76: 141-147.

Bull E.L. and Marx D.B. (2002). Influence of fish and habitat on amphibian communities in high elevation lakes in northeastern Oregon. Northwest Science, 76: 240-248.

266 Burgman M.A. and Fox J.C. (2003). Bias in species range estimates from minimum convex polygons: implications for conservation and options for improved planning. Animal Conservation, 6: 19-28.

Burgman M.A. and Lindenmayer D.B. (1998). Conservation Biology for the Australian Environment. Surrey Beatty and Sons, Chipping Norton, New South Wales.

Burton T.M. and Likens G.E. (1975a). Energy flow and nutrient cycling in salamander populations in the Hubbard Brook Experimental Forest, New Hampshire. Ecology, 56: 1068-1080.

Burtons T.M. and Likens G.E. (1975b). Salamander populations and biomass in the Hubbard Brook Experimental Forest, New Hampshire. Copeia, 1975: 541-546.

Bury R.B. and Corn P.S. (1988). Responses of aquatic and streamside amphibians to timber harvest: a review. In: Raedeke K., (Ed.), Streamside Management: Riparian Wildlife and Forestry Interactions. Institute of Forest Resources, Cont. No. 59, University of Washington, Seattle, pp. 165-181.

Busby J.R. (1988). Potential impacts of climate change on Australia's flora and fauna. In: Pearman G.I., (Ed.), Greenhouse Planning for Climate Change. CSIRO, Melbourne.

Bush S.L. (1997). Vocal behavior of males and females in the majorcan midwife toad. Journal of Herpetology, 31: 251-257.

Caetano M.H., Castanet J. and Francillon H. (1985). Détermination de l´âge de Triturus marmorataus (Latreille 1800) du Parc National de Peneda Geres (Portugal) par squelettochronologie. Amphibia-Reptilia, 6: 117-132.

Caldwell J.P. (1987). Demography and life history of two species of chorus frogs (Anura: Hylidae) in South Carolina. Copeia , 1987: 114-127.

Campbell A. (ed.) (1999). Declines and Disappearances of Australian Frogs. Biodiversity Group, Environment Australia, Canberra.

Carey C. (1978). Factors affecting body temperature in toads. Oecologica, 35: 179-219.

Carey C. (1993). Hypothesis concerning the causes of the disappearance of boreal toads from the mountains of Colorado. Conservation Biology, 7: 355-362.

Carey C. and Alexander M.A. (2003). Climate change and amphibian declines: is there a link? Diversity and Distributions, 9: 111-121.

Carey C., Cohen N. and Rollins-Smith L. (1999). Amphibian declines: an immunological perspective. Developmental and Comparative Immunology, 23: 459-472.

267 Castanet J., Francillon-Viellot H., Meunier F.J. and de Ricqles A. (1993). Bone and individual aging. In: Hall B.K., (Ed.), Bone, Volume 7: Bone Growth. CRC Press, Boca Raton, pp. 245-283.

Castanet J., Meunier F.J. and Francillon-Vieillot H. (1992). Squelettochronologie a partir des et des dents chez les vertébrés. In: Baglinière J.L., Castanet J., Conand F. and Meunier F.J., (Eds.), Tissus durs et âge individuel des vertébrés. Colloque national, Bondy, France, 4-6 mars 1991. Colloques et Séminaires ORSTOM-INRA, Paris, pp. 257-280.

Castanet J. and Smirina E. (1990). Introduction to the skeletochronological method in amphibians and reptiles. Annales des Sciences Naturelles, Zoologie, 11: 191-196.

Castellano S., Rosso A., Doglio S. and Giacoma C. (1999). Body size and calling variation in the green toad (Bufo viridis). Journal of Zoology, London, 248: 83-90.

Caughley G. (1994). Directions in conservation biology. Journal of Animal Ecology, 63: 215-244.

Caughley G. and Gunn A. (1996). Conservation Biology in Theory and Practice. Blackwell Science, Inc., Cambridge, Massachusetts, U.S.A.

Clark J.A., Hoekstra J.M., Boersma P.D. and Kareiva P. (2002). Improving U.S. Endangered Species Act recovery plans: key findings and recommendations of the SCB recovery plan project. Conservation Biology, 16: 1510-1519.

Clarke H.R. (1994). Forest rotation and streamflow benefits. Australian Forestry, 57: 37-44.

Clarke K.R. (1993). Non-parametric multivariant analysis of changes in community structure. Australian Journal of Ecology, 18: 117-143.

CNR (1993). Baw Baw Frog (Philoria frosti). Action Statement No. 55, Department of Conservation and Natural Resources, Victoria.

CNR (1995). Threatened Fauna in Victoria - 1995. Department of Conservation and Natural Resources, Victoria.

Cogger H.G. (2000). Reptiles and Amphibians of Australia. 6th edn, Reed, Sydney.

Cohen J. (1988). Statistical power analysis for the behavioral sciences. 2nd edn, NJ: Lawrence Earlbaum Associates, Hillsdale.

Cohen M.P. and Alford R.A. (1996). Factors affecting diurnal shelter site use by the cane toad, Bufo marinus. Herpetologica, 52: 172-181.

Collins J.P. and Storfer A. (2003). Global amphibian declines: sorting the hypotheses. Diversity and Distributions, 9: 89-98.

268 Commonwealth of Australia (1992). A National Strategy for the Conservation of Australia's Biological Diversity. BDAC, Australian Government Printing Service, Canberra.

Connell J.H. and Sousa W.P. (1983). On the evidence needed to judge ecological stability or persistence. American Naturalist, 121: 789-824.

Cooke A.S. and Frazer J.F.D. (1976). Characteristics of newt breeding sites. Journal of Zoology, London, 178: 223-236.

Corn P.S. (2003). Amphibian breeding and climate change: importance of snow in mountains. Conservation Biology, 17: 622-625.

Corn P.S. and Bury R.B. (1989). Logging in west Oregon: responses of headwater habitats and stream amphibians. Forest Ecology and Management, 29: 39-57.

Corn P.S. and Fogleman J.C. (1984). Extinction of montane populations of the northern leopard frog (Rana pipiens) in Colorado. Journal of Herpetology, 18: 147-152.

Courtice G.P. and Grigg G.C. (1975). A taxonomic revision of the Litoria aurea complex (Anura: Hylidae) in south-eastern Australia. Australian Zoologist, 18: 149-163.

Covacevich J.A. and McDonald K.R. (1993). Distribution and conservation of frogs and reptiles of Queensland rainforests. Memoirs of the Queensland Museum, 34: 189-199.

Crump M.L., Hensley F.R. and Clark K.L. (1992). Apparent decline of the golden toad: Underground or extinct? Copeia, 1992: 413-420.

Cunningham J.M., Rawson R.P. and Roth N.J. (1971). Baw Baw Alpine Reserve survey of vegetation and notable natural features. Victoria File 72/2239, Forests Commission, Melbourne.

Currie W. and Bellis D. (1969). Home range and movements of the bullfrog, Rana catesbeiana Shaw, in an Ontario pond. Copeia, 1969: 688-692.

Curtin C.G. (1995). Can montane landscapes recover from human disturbance? long-term evidence from disturbed sub-alpine communities. Biological Conservation, 74: 49-55.

Czech B., Krausman P.R. and Borkhataria R. (1998). Social construction, political power, and the allocation of benefits to endangered species. Conservation Biology, 12: 1103-1112.

Czechura G.V. and Ingram G.L. (1990). Taudactylus diurnus and the case of the disappearing frogs. Memoirs of the Queensland Museum, 29: 361-365.

Daszak P., Cunningham A.A. and Hyatt A.D. (2003). Infectious disease and amphibian population declines. Diversity and Distributions, 9: 141-150.

269 Daugherty C.H. and Sheldon A.L. (1982). Age-specific movement patterns of the frog Ascaphus truei. Herpetologica, 38: 468-474.

Davidson C., Shaffer H.B. and Jennings M.R. (2001). Declines of the California red-legged frog: climate, uv-b, habitat, and pesticides hypotheses. Ecological Applications, 11: 464-479.

Davidson C., Shaffer H.B. and Jennings M.R. (2002). Spatial tests of the pesticide drift, habitat destruction, UV-B, and climate-change hypotheses for California Amphibian declines. Conservation Biology, 16: 1588-1601.

Davies J.B. (2000). Preliminary Vegetation Description of the Baw Baw Plateau and the Adjacent Montane Escarpment. Unpublished report, Department of Natural Resources and Environment, Gippsland Region, Victoria.

Davies J.B., Carter R.L., Drummond M.B., Hollis G.J., Pascoe C.G., Wallis R.P. and Lester K.P. (1994). Flora and Fauna of the Eastern and Western Tyers Forest Blocks and Adjacent South-eastern Slopes of the Baw Baw National Park, Central Gippsland, Victoria. Unpublished report, Department of Conservation and Natural Resources, Gippsland Area.

Davies J.B., Oates A.M. and Trumbull-Ward A.V. (2002). Ecological Vegetation Class Mapping at 1:25 000 in Gippsland. Final Report. Department of Natural Resources and Environment, Gippsland Region, Victoria.

Davis T.M. (2002). Microhabitat use and movements of the wandering salamander, Aneides vagrans, on Vancouver Island, British Columbia, Canada. Journal of Herpetology, 36: 699-703.

DCE (1991). Baw Baw National Park, Resources and Uses. Department of Conservation and Environment, Victoria. de Bavay J.M. (1993). The developmental stages of the Kyarranus sphagnicolus Moore (Anura: Myobatrachidae). Australian Journal of Zoology, 41: 151-201.

Defonseca P.H.H. and Jocque R. (1982). The palmate newt Triturus helveticus in Flanders (Belgium). Distribution and habitat references. Biological Conservation, 23: 297-307.

Denton J.S. and Beebee T.J.C. (1993). Summer and winter refugia of natterjacks (Bufo calamita) and common toads (Bufo bufo) in Britain. Herpetological Journal, 3: 90-94.

Diller L.V. and Wallace R.L. (1999). Distribution and habitat of Ascaphus truei in streams on managed, young growth forests in north coastal California. Journal of Herpetology , 33: 71-79.

DNRE (2000). A systematic list of vertebrate fauna considered extinct, at risk of extinction or in

270 major decline in Victoria. Department of Natural Resources and Environment, East Melbourne, Victoria.

Docherty S., Bishop P.J. and Passmore N.I. (1995). Calling behavior and male condition in the frog Hyperolius marmoratus. Journal of Herpetology, 29: 616-618.

Dodd Jr., C.K. and Cade B.S. (1998). Movement patterns and the conservation of amphibians breeding in small, temporary wetlands. Conservation Biology, 12: 331-339.

Douglas J.G. and Ferguson J.A. (1976). Geology of Victoria. Geological Survey of Victoria, Sydney.

Driscoll D.A. (1996). Understanding the metapopulation structure of frogs in the Geocrinia rosea complex through population genetics and population biology: implications for conservation and evolution. Ph.D Thesis. University of Western Australia.

Driscoll D.A. (1998). Counts of calling males as estimates of population size in the endangered frogs Geocrinia alba and G. vitellina. Journal of Herpetology, 32: 475-481.

Driscoll D.A. (1999). Skeletochronological assessment of age structure and population stability for two threatened frog species. Australian Journal of Ecology, 24: 182-189.

Drost C.A. and Fellers G.M. (1996). Collapse of a regional frog fauna in the Yosemite area of the California Sierra Nevada, USA. Conservation Biology, 10: 414-425.

Duellman W.E. (1995). Temporal fluctuations in abundances of anuran amphibians in a seasonal Amazonian rainforest. Journal of Herpetology, 29: 13-21.

Duellman W.E. and Trueb L. (1994). Biology of Amphibians. The John Hopkins Press Ltd, London.

Eggert C. (2002). Use of fluorescent pigments and implantable transmitters to track a fossorial toad (Pelobates fuscus). Herpetological Journal, 12: 69-74.

Eggert C. and Guyétant R. (1999). Age structure of a spadefoot toad Pelobates fuscus (Pelobatidae) population. Copeia, 1999: 1127-1130.

Ellner S.P., Fieberg J., Ludwig D. and Wilcox C. (2002). Precision of population viability analysis. Conservation Biology, 16: 258-261.

Ento K. and Matsui M. (2002). Estimation of age structure by skeletochronology of a population of Hynobius nebulosus in a breeding season (Amphibia, Urodela). Zoological Science , 19: 241-247.

271 ESAV (1982). Baw Baw Plateau Course Notes. Baw Baw Course 82/1, Environment Studies Association of Victoria, Melbourne. Unpublished.

Fabens A.J. (1965). Properties and fitting of the von Bertalanffy growth curve. Growth, 29: 265- 289.

Fahrig L. (2001). How much habitat is enough. Biological Conservation, 100: 65-74.

Fahrig L., Pedlar J.H., Pope S.E., Taylor P.D. and Wegner J.F. (1995). Effect of road traffic on amphibian density. Biological Conservation, 73: 177-182.

Fellers G.M. and Drost C.A. (1993). Disappearance of the cascades frog Rana cascadae, at the southern end of its range. Biological Conservation, 65: 177-181.

Fiedler P.L. and Kareiva P.M. (Eds.) (1998). Conservation Biology: For the Coming Decade. 2nd edn, Chapman and Hall, New York.

Fisher R.N. and Shaffer H.B. (1996). The decline of amphibians in California's Great Central Valley. Conservation Biology, 10: 1387-1397.

Fite K.V., Blaustein A., Bengston L. and Hewitt H.E. (1998). Evidence of retinal damage in Rana cascadae: a declining amphibian species. Copeia, 1998: 906-914.

Flood J. (1980). The Moth-Hunters-Aboriginal Prehistory of the Australian Alps. Australian Institute of Aboriginal Studies, Canberra.

Forester D.C. and Lykens D.V. (1991). Age structure in a population of red-spotted newts from the Allegheny Plateau of Maryland. Journal of Herpetology, 25: 373-376.

Francillon-Vieillot H., Arntzen H., Geraudie J.W. and Geraudie J. (1990). Age, growth and longevity of sympatric Triturus cristatus, T. marmoratus and their hybrids (Amphibia, Urodela): a skeletochronological comparison. Journal of Herpetology, 24: 13-22.

Freda J. and Gonzalez R.J. (1986). Daily movements of the treefrog, Hyla andersoni. Journal of Herpetology, 20: 469-471.

Gayou D.C. (1984). Effects of temperature on the mating call of Hyla versicolor. Copeia, 1984: 733-738.

Gehlbach F.R. and Kennedy S.E. (1978). Population ecology of a highly productive aquatic salamander (Siren intermedia). Southwest Naturalist, 23: 423-430.

Gerhardt H.C. (1988). Acoustic properties used in call recognition by frogs and toads. In: B. Fritzsch, M.J. Ryan, W. Wilczynski, T.E. Hetherington and W. Walkowiak, (Eds.), The Evolution of the Amphibian Auditory System. John Wiley and Sons, New York, pp. 455-

272 483.

Gerrodette T. (1987). A power analysis for detecting trends. Ecology, 68: 1364-1372.

Gibbons M.M. and McCarthy T.K. (1983). Age determination of frogs and toads (Amphibia, Anura) from north-western Europe. Zoologica Scripta, 12: 145-151.

Gibbs J.P. (1995). MONITOR users manual: software for estimating the power of population monitoring programs to detect trends in plant and animal abundance. Version 6.2, Yale University, New Haven.

Gibbs J.P. and Breisch A.R. (2001). Climate warming and calling phenology of frogs near Ithaca, New York, 1990-1999. Conservation Biology, 15: 1175-1178.

Gilbert J.M. (1959). Forest succession in the Florentine Valley, Tasmania. Papers and Proceedings of the Royal Society of Tasmania, 93: 129-149.

Gillespie G. and Hero J-M. (1999). Potential impacts of introduced fish and fish translocations on Australian amphibians. In: Campbell A., (Ed.), Declines and Disappearances in Australian Frogs. Biodiversity Group, Environment Australia, Canberra, pp. 131-144.

Gillespie G. and Hines H. (1999). Status of temperate riverine frogs in south-eastern Australia. In: Campbell A., (Ed.), Declines and Disappearances in Australian Frogs. Biodiversity Group, Environment Australia, Canberra, pp. 109-130.

Gillespie G.R. (2000). Survey, Monitoring and Management of the Booroolong Frog and Spotted Tree Frog in New South Wales. Final Report for the 1999/2000 Program. Report Commissioned by the Threatened Species Unit, NSW National Parks and Wildlife Service Southern Zone, Queanbeyan.

Gillespie G.R. (2001a). The role of introduced trout in the decline of the spotted tree frog (Litoria spenceri) in south-eastern Australia. Biological Conservation, 100: 187-198.

Gillespie G.R. (2001b). Ecology of the spotted tree frog Litoria spenceri: an investigation of causes of population declines. Ph.D Thesis. Department of Zoology, University of Melbourne.

Gillespie G.R. (2002). Impacts of sediment loads, tadpole density, and food type on the growth and development of tadpoles of the spotted tree frog Litoria Spenceri: an in-stream experiment. Biological Conservation, 106: 141-150.

Gillespie G.R. and Hollis G.J. (1996). Distribution and habitat of the spotted tree frog, Litoria spenceri Dubois (Anura: Hylidae), and an assessment of potential causes of population declines. Wildlife Research, 23: 49-75.

273 Gillespie G.R., Osborne W.S. and McElhinney N.A. (1995). The Conservation Status of Frogs in the Australian Alps: A Review. Unpublished Report to the Australian Alps Liaison Committee, Canberra.

Glandt D. (1986). Die saisonalen Wanderungen der mitteleuropäischen Amphibien. Bonner Zoologische Beiträge, 37: 211-228.

Goldberg C.S., Goode M.J., Schwalbe C.R. and Jarchow J.L. (2002). External and implanted methods of radiotransmitter attachment to a terrestrial anuran (Eleutherodactylus augusti). Herpetological Review, 33: 191-194.

Gosner K.L. (1960). A simplified table for staging anuran embryos and larvae with notes on identification. Herpetologica, 16: 183-190.

Gower J.C. and Legendre P. (1986). Metric and euclidean properties of dissimilarity coefficients. Journal of Classification , 3: 1-48.

Grant K.P. and Licht L.E. (1995). Effects of ultraviolet radiation on life-history stages of anurans from Ontario, Canada. Canadian Journal of Zoology, 73: 2292-2301.

Griffiths R.A., de Wijer P. and Brady L. (1993). The effect of pH on embryonic and larval development in smooth and palmate newts, Triturus vulgaris and T. helveticus. Journal of Zoology, London, 230: 401-409.

Grover M.C. (1998). Influence of cover and moisture on abundance of the terrestrial salamanders Plethodon cinereus and Plethodon glutinosus. Journal of Herpetology, 32: 489-497.

Gullan P.K., Earl G.E., Forbes S.J., Barley R.H. and Walsh N.G. (1984). Sites of Botanical Significance in Central Gippsland. Environmental Studies Series Report No. 400, Department of Conservation, Forests and Lands, Melbourne.

Gullan P.K., Forebes S.J., Earl G.E., Barley R.H. and Walsh N.G. (1985). Vegetation of South and Central Gippsland. Muelleria, 6: 97-146.

Gullan P.K., Walsh N.G. and Forbes S.J. (1981). Vegetation of the Gippsland Lakes Catchment. Muelleria, 4: 333-383.

Halliday T. and Tejedo M. (1995). Intrasexual selection and alternative mating behaviour. In: Heatwole H. and Sullivan B.K., (Eds.), Amphibian Biology, Volume 2. Social Behaviour. Surrey Beatty and Sons, Chipping Norton, pp. 419-468.

Halliday T.R. and Verrell P.A. (1988). Body size and age in amphibians and reptiles. Journal of Herpetology, 22: 253-265.

274 Hamer A.J, Lane S.J and Mahony M.J (2002). Management of freshwater wetlands for the endangered green and golden bell frog (Litoria aurea): roles of habitat determinants and space. Biological Conservation, 106: 413-424.

Hannah L., Midgley G.F., Lovejoy T., Bond W.J., Bush M., .Lovett J.C., S.D. and Woodward F.I. (2002). Conservation of biodiversity in a changing climate. Conservation Biology, 16: 264-268.

Harris R.T. (1975). Seasonal activity and microhabitat utilisation in Hyla cadaverina (Anura: Hylidae). Herpetologica, 31: 236-239.

Harris S., Cresswell W.J., Forde P.G., Trewhella W.J., Woollard T. and Wray S. (1990). Home- range analysis using radio-tracking data - a review of problems and techniques particularly as applied to the study of mammals. Mammal Review, 20: 97-123.

Harvell C.D., Mitchell C.E., Ward J.R., Altizer S., Dobson A.P., Ostfeld R.S. and Samuel M.D. (2002). Climate warming and disease risks for terrestrial and marine biota. Science, 296: 2158-2162.

Hatano F.H, Rocha C.F.D and Van Sluys M. (2002). Environmental factors affecting calling activity of a tropical diurnal frog (Hylodes phyllodes : Leptodactylidae). Journal of Herpetology, 36: 314-318.

Hatch A.C. and Blaustein A.R. (2000). Combined effects of UV-B, nitrate, and low pH reduce the survival and activity level of larval cascades frogs (Rana cascadae). Archives of Environmental Contamination and Toxicology, 39: 494-499.

Hays J.B., Blaustein A.R., Kiesecker J.M., Hoffman P.D., Pandelova I., Coyle D. and Richardson T. (1996). Developmental responses of amphibians to solar and artificial UVB sources: a comparative study. Photochemical Photobiology, 64: 449-455.

Hecnar S.J. and M'Closkey R.T. (1997). The effects of predatory fish on amphibian species richness and distribution. Biological Conservation, 79: 123-131.

Helps A. (2001). Climate Change Impacts at a Catchment Level. Traralgon 2001 - Planning for the Inevitable, Victorian Flood Management Conference. Victoria.

Hels T. (2002). Population dynamics in a Danish metapopulation of spadefoot toads Pelobates fuscus. Ecography, 25: 303-313.

Hels T. and Buchwald E. (2001). The effect of road kills on amphibian populations. Biological Conservation, 99: 331-340.

Hemelaar A.S.M. (1988). Age, growth and other population characteristics of Bufo bufo from

275 different latitudes and altitudes. Journal of Herpetology, 22: 369-388.

Hemelaar A.S.M. and van Gelder J.J. (1980). Annual growth rings in phalanges of Bufo bufo (Anura, Amphibia) from the Netherlands and their use for age determination. Netherlands Journal of Zoology, 30: 129-135.

Hermans K., Pinxten R. and Eens M. (2002). Territorial and vocal behaviour in a captive dart- poison frog, Epipedobates tricolor Boulenger, 1899 (Anura : Dendrobatidae). Belgian Journal of Zoology, 132: 105-109.

Hero J-M. (1989). A simple code for toe clipping anurans. Herpetological Review, 20: 66-67.

Hero J-M. and Gillespie G.R. (1997). Epidemic disease and amphibian declines in Australia. Conservation Biology, 11: 1023-1025.

Heyer W.R., Donnelly M.A., McDiarmid R.W., Hayek L-A. C. and Foster M.S. (eds.) (1994). Measuring and Monitoring Biological Diversity: Standard Methods for Amphibians. Smithsonian Institute Press, Washington.

Heyer W.R., Rand A.S., da Cruz C.A.G. and Peixoto O.L. (1988). Deciminations, extinctions, and colonisations of frog populations in southeast Brazil and their evolutionary implications. Biotropica, 20: 230-235.

Hills E.S. (1959). Physiography of Victoria. 4th edn, Whitcombe and Tombs, Victoria.

Hines H., Mahony M. and McDonald K. (1999). An assessment of frog declines in wet subtropical Australia. In: Campbell A., (Ed.), Declines and Disappearances of Australian Frogs. Biodiversity Group, Environment Australia, Canberra, pp. 44-63.

Hisai N., Chiba S., Yano M. and Sugawara T. (1987). Ecological studies of Bufo japonicus formosus Boulenger. (IX) Spring migration to breeding sites and homing. Reports from the Institute of Natural Studies, 18: 1-13. (in Japanese with English summary).

Holenweg A.K. and Reyer H.U. (2000). Hibernation behavior of Rana lessonae and R. esculenta in their natural habitat. Oecologia, 123: 41-47.

Hollis G.J. (1995). Reassessment of the distribution, abundance and habitat of the Baw Baw Frog (Philoria frosti Spencer): Preliminary findings. Victorian Naturalist, 112: 190-201.

Hollis G.J. (1997). Recovery Plan for the Baw Baw Frog ( Philoria frosti). Unpublished report to the Endangered Species Program, Environment Australia, Canberra.

Hollis G.J. (2002). Baw Baw Frog (Philoria frosti) Recovery Plan: Major Project Review. Unpublished report to the Biodiversity Group, Environment Australia, Canberra.

276 Hollis G.J., Carter R.L., Davies J.B., Davies M.J., Drummond M.B., Wallis R.P. and Weeks J.M. (1995). Survey of Vegetation and Vertebrate Fauna in the Cascade Forest Block and Adjacent Areas of Baw Baw National Park, Central Gippsland, Victoria. Unpublished report, Department of Conservation and Natural Resources, Gippsland Area.

Holper P. (1995). Acid air? Tracking acidification. ECOS, 86 (Summer 1995/96): 30-34. Melbourne, CSIRO.

Houck L.D. (1982). With rates and age maturity for the Plethodontid salamander Bolitoglossa subpalmata. Copeia, 1982: 474-478.

Houlahan J.E., Findley C.S., Schmidy B.R., Meyer A.H. and Kuzmin S.L. (2000). Quantitative evidence for global amphibian declines. Nature, 404: 752-755.

Humason G.L. (1979). Animal Tissue Techniques. W.H. Freeman and Company, San Francisco.

Humphries R.B. (1979). Dynamics of a breeding frog community. Ph.D Thesis. Australian National University.

Hunter D. and Gillespie G.R. (1999). The distribution, abundance and conservation status of riverine frogs in Kosciusko National Park. Australian Zoologist, 31: 198-209.

Hunter D.A., Osborne W.S. and Smith M.J. (1998). Distribution and abundance of the alpine tree frog (Litoria verreauxii alpina) in the Australian Alps National Parks. Report on the first seasons survey (1996-97). Unpublished Report to the New South Wales National Parks and Wildlife Service, Applied Ecology Research Group, University of Canberra, Canberra.

Hunter Jr., M.L. (1996). Fundamentals of Conservation Biology. Blackwell Science. Inc., Cambridge, Massachusetts, U.S.A.

Ingram G.J. (1990). The mystery of the disappearing frogs. Wildlife Australia, 27: 6-7.

Ingram G.J. and Corben C.J. (1975). A new species of Kyarranus (Anura: Leptodactylidae) from Queensland. Memoirs of the Queensland Museum, 17: 335-339.

Ingram G.J. and McDonald K.R. (1993). An update on the decline of Queensland's frogs. In: Lunney D. and Ayers D., (Eds.), Herpetology in Australia: a diverse discipline. Royal Zoological Society of New South Wales, Mosman, pp. 297-303.

Irwin E.R. and Freeman M.C. (2002). Proposal for adaptive management to conserve biotic integrity in a regulated segment of the Tallapoosa River, Alabama, U.S.A. Conservation Biology, 16: 1212-1222.

277 Jablonski N.G. (1998). Ultraviolet light-induced neural tube defects in amphibian larvae and their implications for the evolution of melanized pigmentation and declines in amphibian populations. Journal of Herpetology, 32: 455-457.

Jaeger R.G. (1981). Activity of Neotropical frogs in relation to ambient light. Biotropica, 13: 50- 65.

Jameson D.L. (1955). The population dynamics of the cliff frog, Syrrhophus mamocki. American Midland Naturalist, 54: 342-381.

Jameson D.L. (1956). Growth, dispersal and survival of the Pacific tree frog. Copeia, 1956: 25-29.

Jehle R., Hodl W. and Thonke A. (1995). Structure and dynamics of central European amphibian populations: a comparison between Triturus dobrogicus (Amphibia, Urodela) and Pelobates fuscus (Amphibia, Anura). Australian Journal of Ecology, 20: 362-366.

Jongman R.H.G., ter Braak C.J.F. and van Tongeren O.F.R. (1987). Data Analysis in Community and Landscape Ecology. Pudoc, Wageningen, The Netherlands.

Kam Y.C. and Chen T.C. (2000). Abundance and movement of a riparian frog (rana swinhoana) in a subtropical forest of guandau stream, Taiwan. Zoological Studies, 39: 67-76.

Kantvilas G. and Minchin P.R. (1989). An analysis of epiphytic lichen communities in Tasmanian cool temperate rainforest. Vegetatio, 84: 99-112.

Kats L.B. and Ferrer R.P. (2003). Alien predators and amphibian declines: review of two decades of science and the transition to conservation. Diversity and Distributions, 9: 99-110.

Kenkel N.C. and Orlóci L. (1986). Applying metric and non-metric multidimensional scaling to ecological studies: some new results. Ecology, 67: 919-928.

Kenward R.E. and Hodder K.H. (1996). RANGES V: An analysis system for biological location data. Institute of Terrestrial Ecology, Dorset, United Kingdom.

Kerr J.B. and McElroy C.T. (1993). Evidence for large upward trends of ultraviolet-B radiation linked to ozone depletion. Science, 262: 1032-1034.

Khonsue W., Matsui M., Hirai T. and Misawa Y. (2001). Age determination of wrinkled frog, Rana rugosa with special reference to high variation in postmetamorphic body size (Amphibia: Ranidae). Zoological Science, 18: 605-612.

Kiesecker J.M. and Blaustein A.R. (1995). Synergism between UV-B radiation and pathogen magnifies amphibian embryo mortality in nature. Proceedings of the National Academy of Sciences, U.S.A., 92: 11049-11052.

278 Knowles R. (1994). The systematics of the endangered montane rainforest frogs of the genus Philoria (Anura: Myobatrachidae), and implications for conservation management. B.Sc. Honours Thesis. Department of Biological Sciences, University of Newcastle.

Knowles R., Mahony M., Armstrong J. and Donnellan S. (in press). Systematics of sphagnum frogs of the genus Philoria (Anura: Myobatrachidae) in eastern Australia. Records of the Australian Museum.

Knutson M.G., Sauer J.R., Olsen D.A., Mossman M.J., Hemesath L.M. and Lannoo M.J. (1999). Effects of landscape composition and wetland fragmentation on frog and toad abundance and species richness in Lowa and Wisconsin, U.S.A. Conservation Biology, 13: 1437- 1446.

Koenig W.D., van Vuran D. and Hooge P.N. (1996). Detectability, philopatry, and the distribution of dispersal distances in vertebrates. Trends in Ecology and Evolution, 11: 514-517.

Kramer D.C. (1973). Movements of western chorus frogs Pseudacris triseriata triseriata tagged with Co-60. Journal of Herpetology, 7: 231-235.

Kramer P., Reichenbach N., Hayslett M. and Sattler P. (1993). Population dynamics and conservation of the peaks otter salamander, Plethodon hubrichti. Journal of Herpetology, 27: 431-435.

Krebs C.J. (1994). Ecology: The Experimental Analysis of Distribution and Abundance. 4th edn, Harper Collins, New York.

Kruskal J.B. (1964). Multidimensional scaling by optimizing goodness of fit to a nonmetric hypothesis. Psychometrika, 29: 1-27.

Kruskal J.B., Young F.W. and Seery J.B. (1973). How to use KYST, a flexible program to do multidimensional scaling and unfolding. Unpublished Report, Bell Telephone Laboratories, U.S.A.

Kusano T., Fukuyama K. and Miyashita N. (1995a). Age determination of the stream frog, Rana sakuraii, by skeletochronology. Journal of Herpetology, 29: 625-628.

Kusano T., Maruyama K. and Kaneko S. (1995b). Post-breeding dispersal of the Japanese toad, Bufo japonicus formosus. Journal of Herpetology, 29: 633-638.

Kusano T. and Miyashita K. (1984). Dispersal of the salamander, Hynobius nebulosus tokyoensis. Journal of Herpetology, 18: 349-353.

La Marca E. and Reinthaler H.P. (1991). Population changes in Atelopus species of the Cordeillera de Merida, Venezuela. Herpetological Review, 22: 125-128.

279 Laan R. and Verboom B. (1990). Effects of pool size and isolation of amphibian communities. Biological Conservation, 54: 251-262.

Lacy R.C. (1993). Vortex: a computer simulation model for population viability analysis. Wildlife Research, 20: 45-65.

Lamoureux V.S. and Madison D.M. (1999). Overwintering habitats of radio-implanted green frogs, Rana clamitans. Journal of Herpetology, 33: 430-435.

Lamoureux V.S., Maerz J.C. and Madison D.M. (2002). Premigratory autumn foraging forays in the green frog, Rana clamitans. Journal of Herpetology, 36: 245-254.

Lande R. (1988). Genetics and demography in biological conservation. Science, 241: 1455-1460.

Langhelle A., Lindell M.J. and Nyström P. (1999). Effects of ultraviolet radiation on amphibian embryonic and larval development. Journal of Herpetology, 33: 449-456.

Langkilde T. and Alford R.A. (2002). The tail wags the frog: harmonic radar transponders affect movement behavior in Litoria lesueuri. Journal of Herpetology, 36: 711-715.

Laurance W.F. (1996). Catastrophic declines of Australian rainforest frogs: is unusual weather responsible? Biological Conservation, 77: 203-212.

Laurance W.F., McDonald K.R. and Speare R. (1996). Epidemic disease and the catastrophic decline of Australian rain forest frogs. Conservation Biology, 10: 406-413.

Lazell J.D., Goodyear N.C. and Liao W-P. (1988). Rana areolata (Florida gopher frog). Movement. Herpetological Review, 19: 33

LCC (1991). Melbourne Area District 2 Review, Descriptive Report. Land Conservation Council, Victoria.

Lecis R. and Norris K. (2003). Habitat correlates of distribution and local population decline of the endemic Sardinian newt Euproctus platycaphalus. Biological Conservation, 115: 303- 317.

Leclair Jr., R. (1990). Relationships between relative mass of the skeleton, endosteal resorption, habitat and precision of age determination in ranid amphibians. Annales des Sciences Naturelles, Zoologie, 11: 205-208.

Leclair R. and Castanet J. (1987). A skeletochronological assessment of age and growth in the frog Rana pipiens Shreber (Amphibia, Anura) from southwestern Quebec. Copeia, 1987: 361- 369.

Lemckert F. and Brassil T. (2000). Movement and habitat use of the endangered giant barred river

280 frog (Mixophyes iteratus) and the implications for its conservation in timber production forests. Biological Conservation, 96: 177-184.

Letnic M. and Fox B.J. (1997). The impact of industrial fluoride fallout on faunal succession following sand-mining of dry sclerophyll forest at Tomago, NSW, II. Myobatrachid frog recolonisation. Biological Conservation, 82: 137-146.

Levene H. (1960). Robust tests for equality of variance. In: I. Olkin, (Ed.), Contributions to Probability and Statistics. Standard University Press, Palo Alto, California, pp. 278-292.

Licht L.E. (1969). Comparative breeding biology of the red-legged frog (Rana aurora aurora) and the western spotted frog (Rana pretiosa pretiosa) in southwestern British Columbia. Canadian Journal of Zoology, 47: 505-509.

Lindenmayer D.B. and Franklin J.F. (2002). Conserving Forest Biodiversity: a comprehensive multiscaled approach. Island Press, Washington, DC.

Lips K.R. (1998). Decline of a tropical montane amphibian fauna. Conservation Biology, 12: 106- 117.

Lips K.R. (1999). Mass mortality and population declines of anurans at an upland site in western Panama. Conservation Biology, 13: 117-125.

Lips K.R., Green D.E. and Papendick R. (2003). Chytridiomycosis in wild frogs from southern Costa Rica. Journal of Herpetology, 37: 215-218.

Littlejohn M.J. (1963). The breeding biology of the Baw Baw Frog. Proceedings of the Linnean Society of New South Wales, 88: 273-276.

Littlejohn M.J. (1977). Long-range acoustic communication in anurans: an integrated and evolutionary approach. In: D.H. Taylor and S.I. Guttman, (Eds.), The Reproductive Biology of Amphibians. Plenum Press, New York, pp. 263-294.

Littlejohn M.J. (1987). Calls of Victorian Frogs. Compiling and mixing by D. Smith. Department of Zoology, University of Melbourne, Parkville, Victoria.

Liu G-J. and White J. (1999). DEM-based Habitat Modelling for the Baw Baw Frog (Philoria frosti). Natural Resource Systems, Department of Natural Resources and Environment, Melbourne.

Long L.E., Saylor L.S. and Soulé M.E. (1995). A pH/UV-B synergism in amphibians. Conservation Biology, 9: 1301-1303.

Lüddecke H. (2002). Male and female responses to call playback in the Andean frog Colostethus

281 subpunctatus. Amphibia-Reptilia, 23: 141-150.

Luddecke H. and Sanchez O.R. (2002). Are tropical highland frog calls cold-adapted? The case of the Andean frog Hyla labialis. Biotropica, 34: 281-288.

Lumsden L.F., Alexander J.S.A., Hill F.A.R., Krasna S.P. and Silveira C.E. (1991). The Vertebrate Fauna of the Land Conservation Council Melbourne-2 Study Area. Arthur Rylah Institute for Environmental Research Technical Report No. 115, Department of Conservation and Environment, Victoria.

Macnally R.C. (1985). Habitat and microhabitat distributions in relation to ecological overlap in two species of Ranidella (Anura). Australian Journal of Zoology, 33: 329-338.

Madison D.M. (1997). The emigration of radio-implanted spotted salamanders, Ambystoma maculatum. Journal of Herpetology, 31: 542-551.

Madison D.M. and Farrand III L. (1998). Habitat use during breeding and emigration in radio- implanted tiger salamanders, Ambystoma tigrinum. Copeia, 1998: 402-410.

Maguire L.A. (1997). Decision analysis in conservation biology. In: Meffe G.K. and Carroll C.R., (Eds.), Principles of Conservation Biology. 2nd edn. Sinauer Associates, Sunderland, Massachusetts, pp. 377-383.

Mahony M. (1999). Review of the declines and disappearances within the bell frog species group (Litoria aurea species group) in Australia. In: Campbell A., (Ed.), Declines and Disappearances of Australian Frogs. Biodiversity Group, Environment Australia, Canberra, pp. 81-93.

Mahony M. and Knowles R. (1994). A taxonomic review of selected frogs of north-east New South Wales forests. North-east Forests Biodiversity Study Report No. 3g., Unpublished report to the New South Wales National Parks and Wildlife Service.

Mahony M.J. (1993). The status of frogs in the Watagan Mountains area, the Central Coast of New South Wales. In: Lunney D. and Ayers D., (Eds.), Herpetology in Australia: a diverse discipline. Royal Zoological Society of New South Wales, Mosman, pp. 257-264.

Mahony M.J. (1996). The decline of the green and golden bell frog Litoria aurea viewed in the context of declines and disappearances of other Australian frogs. Australian Zoologist, 30: 237-247.

Malone B.S. (1985a). Status, distribution and ecology of the Baw Baw Frog (Philoria frosti). Arthur Rylah Institute for Environmental Research Technical Report Series No. 36, Department of Conservation, Forests and Lands, Victoria.

282 Malone B.S. (1985b). Mortality during the early life history stages of the Baw Baw Frog (Philoria frosti). In: G. Grigg, S. Shine and H. Ehmann, (Eds.), Biology of Australasian Frogs and Reptiles. Surrey Beatty and Sons, Sydney, pp. 1-5.

Mansergh I. and Norris K.C. (1982). Sites of Zoological Significance in Central Gippsland. Environmental Studies Series No. 359,Volume 1, Ministry for Conservation, Victoria.

Marnell F. (1998). Discriminant analysis of the terrestrial and aquatic habitat determinants of the smooth newt (Triturus vulgaris) and the (Rana temporaria) in Ireland. Journal of Zoology, London, 244: 1-6.

Marsh D.M. (2000). Variable responses to rainfall by breeding tungara frogs. Copeia, 2000: 1104- 1108.

Marsh D.M. (2001). Fluctuations in amphibian populations: a meta-analysis. Biological Conservation, 101: 327-335.

Marsh D.M., Rand A.S. and Ryan M.J. (2000). Effects of inter-pond distance on the breeding ecology of tungara frogs. Oecologia, 122: 505-513.

Martin A.A. (1967). Australian anuran life histories: some evolutionary and ecological aspects. In: Weatherly A.H., (Ed.), Australian Inland Waters and Their Fauna. Australian National University Press, Canberra, pp. 175-191.

Marunouchi J., Kusano T. and Ueda H. (2002). Fluctuation in abundance and age structure of a breeding population of a Japanese brown frog, Rana japonica Günther (Amphibia, Anura). Zoological Science, 19: 343-350.

Matthews K.R. and Pope K.L. (1999). A telemetric study of the movement patterns and habitat use of Rana muscosa, the mountain yellow-legged frog, in a high-elevation basin in Kings Canyon National Park, California. Journal of Herpetology, 33: 615-624.

Mazerolle M.J. (2001). Amphibian activity, movement patterns, and body size in fragmented peat bogs. Journal of Herpetology , 35: 13-20.

McCarthy M.A. (1998). Identifying declining and threatened species with museum data. Biological Conservation, 83: 9-17.

McCoy E.D. (1994). "Amphibian decline": a scientific dilemma in more ways than one. Herpetologica, 50: 98-103.

McDonald K.R. (1990). Rheobatrachus liem and Taudactylus Straughan and Lee (Anura: Leptodactylidae) in Eungella National Park, Queensland: distribution and decline. Transactions of the Royal Society of South Australia, 114: 187-194.

283 McDonald R.C., Isbell R.F., Speight J.G., Walker J. and Hopkins M.S. (1990). Australian Soil and Land Survey Field Hand Book. 2nd edn, Inkata Press, Melbourne.

Measey G.J. (2001). Growth and ageing of feral Xenopus laevis (Daudin) in south Wales, U.K. Journal of Zoology, London, 254: 547-555.

Meffe G.K. and Carroll C.R. (1997). Principles of Conservation Biology. 2nd edn, Sinauer Associates, Sunderland, Massachusetts.

Meir E. and Kareiva P.M. (1998). Contributions of spatially explicit landscape models to conservation biology. In: Fiedler P.L. and Kareiva P.M., (Eds.), Conservation Biology: For the Coming Decade, 2nd edn. Chapman and Hall, New York, pp. 497-507.

Miaud C., Sanuy D. and Avrillier J-N. (2000). Terrestrial movements of the natterjack toad Bufo calamita (Amphibia, Anura) in a semi-arid, agricultural landscape. Amphibia-Reptilia, 21: 357-369.

Michael S.F. (1996). Courtship calls of three species of Eleutherodactylus from Puerto Rico (Anura: Leptodactylidae). Herpetologica, 52: 116-120.

Middleton E.M., Herman J.R., Celarier E.A., Wilkinson J.W., Carey C. and Rusin R.J. (2001). Evaluating ultraviolet radiation exposure with satellite data at sites of amphibian declines in central and south America. Conservation Biology, 15: 914-929.

Mills L.S. and Bury R.B. (2002). Terrestrial and stream amphibians across clearcut-forest interfaces in the Siskiyou Mountains, Oregon. Northwest Science, 76: 129-140.

Minchin P.R. (1987). An evaluation of the relative robustness of techniques for ecological ordination. Vegetatio, 69: 89-107.

Minchin P.R. (1991). DECODA database for ecological community data, documentation notes. Australian National University, Canberra.

MMBW (1975). Report on the Environmental Study into the and Associated Works. Volume 1 and 2, Melbourne and Metropolitan Board of Works, Melbourne.

Monello R.J. and Wright R.G. (1999). Amphibian habitat preferences among artificial ponds in the Palouse region of northern Idaho. Journal of Herpetology, 33: 298-303.

Monson P.D., Call D.J., Dean D.A., Cox A., Liber K. and Ankley G.T. (1999). Photoinduced toxicity of fluoranthene to northern leopard frogs (Rana pipens). Environmental Toxicology and Chemistry, 18: 308-312.

Moore J.A. (1961). The frogs of eastern New South Wales. Bulletin of the American Museum of

284 Natural History, 121: 149-386.

Moore M.K. and Klerks P.L. (1998). Interactive effect of high temperature and low pH on sodium flux in tadpoles. Journal of Herpetology, 32: 588-592.

Moorrees A. and Molnar C. (1992). Sites of Botanical Significance in the Land Conservation Council Melbourne District-2 Study Area. Unpublished report, Department of Conservation and Environment, Victoria.

Morgan L.A. and Buttemer W.A. (1996). Predation by the non-native fish Gambusia holbrooki on small Litoria aurea and L. dentata tadpoles. Australian Zoologist, 30: 143-149.

Morrison C. and Hero J-M. (2003). Geographic variation in life-history characteristics of amphibians: a review. Journal of Animal Ecology, 72: 270-279.

Murphy C.G. (1999). Nightly timing of chorusing by male barking treefrogs (Hyla gratiosa): the influence of female arrival and energy. Copeia, 1999: 333-347.

Murphy M.L. and Hall J.D. (1981). Varied effects of clear-cut logging on predators and their habitat in small streams of the Cascade Mountains, Oregon. Canadian Journal of Fisheries and Aquatic Science, 38: 137-145.

Muths E., Corn P.S., Pessier A.P. and Green D.E. (2003). Evidence for disease-related amphibian decline in Colorado. Biological Conservation, 110: 357-365.

Navas C.A. (1996a). The effect of temperature on the vocal activity of tropical anurans: a comparison of high and low-elevation species. Journal of Herpetology , 30: 488-497.

Navas C.A. (1996b). Thermal dependency of field locomotor and vocal performance of high- elevation anurans in the tropical Andes. Journal of Herpetology , 30: 478-487.

Navas C.A. and Bevier C.R. (2001). Thermal dependency of calling performance in the eurythermic frog Colostethus subpunctatus. Herpetologica, 57: 384-395.

Nijhuis M.J. and Kaplan R.H. (1998). Movement patterns and life history characteristics in a population of the cascade torrent salamander (Rhyacotriton cascadae) in the Columbia River Gorge, Oregon. Journal of Herpetology, 32: 301-304.

NRE (1998). Forest Management Plan for the Central Highlands. Department of Natural Resources and Environment, Melbourne, Victoria.

NRE (1999). Victoria's Biodiversity Education Resource Book 1 CSF Levels 3 and 4. Department of Natural Resources and Environment, Melbourne, Victoria.

NRE (2001). Victorian Flora Species Index, Including Vascular and Non-vascular Taxa.

285 Department of Natural Resources and Environment, Heidelberg, Victoria.

Nyman S. (1986). Mass mortality in larval Rana sylvatica attributable to the bacterium, Aeromonas hydrophilum. Journal of Herpetology, 20: 196-201.

Nystrom P., Birkedal L., Dahlberg C. and Bronmark C. (2002). The declining spadefoot toad Pelobates fuscus: calling site choice and conservation. Ecography, 25: 488-498.

Odendaal F.J. and Bull C.M. (1983). Water movements, tadpole competition and limits to the distribution of the frogs Ranidella riparia and R. signifera. Oecologia (Berlin), 57: 361- 367.

Odendaal F.J., Bull C.M. and Nias R.C. (1982). Habitat selection in tadpoles of Ranidella signifera and R. riparia (Anura: Leptodactylidae). Oecologia, 52: 411-414.

Okuno R. (1984). Studies on the natural history of the Japanese toad, Bufo japonicus japonicus. I. Habitat groups and the interactions among them. Japanese Journal of Ecology, 34: 113- 121. (in Japanese with English synopsis).

Okuno R. (1986). Studies on the natural history of the Japanese toad, Bufo japonicus japonicus. X. Mating and spawning behaviors. Japanese Journal of Ecology, 36: 11-18. (in Japanese with English synopsis).

Orians G.H. and Soulé M.E. (2001). Whither conservation biology research? Conservation Biology, 15: 1187-1188.

Osborne W., Hunter D. and Hollis G. (1999). Population declines and range contraction in Australian alpine frogs. In: Campbell A., (Ed.), Declines and Disappearances of Australian Frogs. Biodiversity Group, Environment Australia, Canberra, pp. 145-157.

Osborne W.S. (1989). Distribution, relative abundance and conservation status of Corroboree frogs, Pseudophryne corroboree Moore (Anura: Myobatrachidae). Australian Wildlife Research, 16: 537-547.

Osborne W.S. (1990a). Declining frog populations and extinctions in the Canberra region. Bogong, 11: 4-7.

Osborne W.S. (1990b). The conservation biology of Pseudophryne Moore (Anura: Hylidae): a study of insular populations. Ph.D Thesis. Australian National University.

Osborne W.S. (1992). Declines and extinctions in populations of frogs in the Australian Capital Territory: a discussion paper. Internal Report, 92/8, ACT Parks and Conservation Service, Canberra.

286 Osborne W.S., Littlejohn M.J. and Thomson S.A. (1996). Former distribution and apparent disappearance of the Litoria aurea complex from the Southern Tablelands of New South Wales and the Australian Capital Territory. Australian Zoologist, 30: 190-198.

Osborne W.S. and Norman J.A. (1991). Conservation genetics of corroboree frogs, Pseudophryne corroboree: population subdivision and genetic divergence. Journal of Zoology, 39: 285- 297.

Ovaska K. (1991). Reproductive phenology, population structure, and habitat use of the frog Eleutherodactylus johnstonei, West Indies. Journal of Herpetology, 25: 424-430.

Ovaska K.E. and Caldbeck J. (1997a). Courtship behavior and vocalizations of the frogs Eleutherodactylus antillensis and E. cochranae on the British Virgin Islands. Journal of Herpetology, 31: 149-155.

Ovaska K.E. and Caldbeck J. (1997b). Vocal behaviour of the frog Eleutherodactylus antillensis on the British Virgin Islands. Animal Behaviour, 54: 181-188.

Ovaska K.E. and Caldbeck J. (1999). Courtship call of the frog Eleutherodactylus schwartzi from the British Virgin Islands. Journal of Herpetology, 33: 501-504.

Oza G.M. (1990). Ecological effects of the frog's legs trade. Environmentalist, 10: 39-41.

Pahkala M., Räsänen K., Laurila A., Johanson U., Björn L.O. and Merilä J. (2002). Lethal and sublethal effects of UV-B/pH synergism on common frog embryos. Conservation Biology, 16: 1063-1073.

Palis J.G. (1998). Breeding biology of the gopher frog, Rana capito, in Western Florida. Journal of Herpetology, 32: 217-223.

Parris K.M. (2001). Distribution, habitat requirements and conservation of the cascade treefrog (Litoria pearsoniana, Anura: Hylidae). Biological Conservation, 99: 285-292.

Parris K.M. (2002). The distribution and habitat requirements of the great (Mixophyes fasciolatus). Wildlife Research, 29: 469-474.

Paton P.W.C. and Crouch III W.B. (2002). Using the phenology of pond-breeding amphibians to develop conservation strategies. Conservation Biology, 16: 194-204.

Pearson O.P. and Bradford D.F. (1976). Thermoregulation of lizards and toads at high altitudes in Peru. Copeia, 1976: 155-170.

Pechmann J.H.K., Scott D.E., Semlitsch R.D., Caldwell J.P., Vitt L.J. and Gibbons J.W. (1991). Declining amphibian populations: the problem of separating human impacts from natural

287 fluctuations. Science, 253: 892-895.

Pechmann J.H.K. and Wake D.B. (1997). Declines and disappearances of amphibian populations. In: Meffe G.K. and Carroll C.R., (Eds.), Principles of Conservation Biology. 2nd edn. Sinauer Associates, Sunderland, Massachusetts, pp. 135-137.

Pechmann J.H.K. and Wilbur H.M. (1994). Putting declining amphibian populations in perspective: natural fluctuations and human impacts. Herpetologica, 50: 65-84.

Peel B. (1999). Rainforests and Cool Temperate Mixed Forests of Victoria. Flora and Fauna Program, Department of Natural Resources and Environment, Victoria.

Pengilley R.K. (1971). Calling and associated behaviour of some species of Pseudophryne (Anura: Leptodactylidae). Journal of Zoology, London, 163: 73-92.

Penna M. (1997). Selectivity of evoked vocal responses in the time domain by frogs of the genus Batrachyla. Journal of Herpetology, 31: 202-217.

Perlman D.L. and Adelson G. (1997). Biodiversity: Exploring Values and Priorities in Conservation. Blackwell Science Inc., cambridge, Massachusetts.

Peters D. (1983). TASFORHAB. In: Myers K., Margules C.R. and Musto I., (Eds.), Proceedings of the Symposium: Survey Methods for Nature Conservation. CSIRO Division of Water and Land Resources, Adelaide.

Peterson G.S., Johnson L.B., Axler R.P. and Diamond S.A. (2002). Assessment of the risk of solar ultraviolet radiation to amphibians. II. in situ characterization of exposure in amphibian habitats. Environmental Science and Technology, 36: 2859-2865.

Phillips K. (1990). Where have all the frogs and toads gone? Bioscience, 40: 422-424.

Pilliod D.S., Peterson C.R. and Ritson P.I. (2002). Seasonal migration of Columbia spotted frogs (Rana luteiventris) among complementary resources in a high mountain basin. Canadian Journal of Zoology, 80: 1849-1862.

Pough F.H., Magnusson W.E., Ryan M.J. and Wells K.D. (1992). Behavioral energetics. In: M.E. Federa and W.W. Burggren, (Eds.), Environmental Physiology of the Amphibians. University of Chicago Press, Chicago, pp. 395-436.

Pough F.H., Taigen T.L., Stewart M.M. and Brussard P.F. (1983). Behavioural modification of evaporative water loss by a Puerto Rican frog. Ecology, 64: 244-252.

Pounds J.A. and Crump M.L. (1994). Amphibian declines and climate disturbance: the case of the golden toad and the harlequin frog. Conservation Biology, 8: 72-85.

288 Pounds J.A., Fogden M.P.L., Savage J.M. and Gorman G.C. (1997). Tests of null models for amphibian declines on a tropical mountain. Conservation Biology, 11: 1307-1322.

Pounds J.A., Fogden P.L. and Campbell J.H. (1999). Biological response to climate change on a tropical mountain. Nature, 398: 611-615.

Primack R.B. (1993). Essential of Conservation Biology. Sinauer Associates, Inc., Sunderland, Massachusetts.

Pröhl H. and Berke O. (2001). Spatial distributions of male and female strawberry poison frogs and their relation to female reproductive resources. Oecologia, 129: 534-542.

Radwan N.M. and Schneider H. (1988). Social behaviour, call repertory and variations in the calls of the pool frog, Rana lessonae (Anura, Ranidae). Amphibia-Reptilia, 9: 329-351.

Rathbun G.B. and Murphry T.G. (1996). Evaluation of a radio-belt for ranid frogs. Herpetological Review, 27: 187-189.

Reed J.M. and Blaustein A.R. (1995). Assessment of "nondeclining" amphibian populations using power analysis. Conservation Biology, 9: 1299-1300.

Reed J.M., Mills L.S., Dunning J.R., J.B., Menges E.S., McKevey K.S., Frye R., Beissinger S.R., Anstett M-C. and Miller P. (2002). Emerging issues in population viability analysis. Conservation Biology, 16: 7-19.

Richards S.J., McDonald K.R. and Alford R.A. (1993). Declines in populations of Australia's endemic tropical rainforest frogs. Pacific Conservation Biology, 1: 66-77.

Richards S.J., Sinsch U. and Alford R.A. (1994). Radio Tracking. In: W. Heyer, M.A. Donnelly, R.W. McDiarmid, L-A. C. Hayek and M.S. Foster, (Eds.), Measuring and Monitoring Biological Diversity. Standard Methods for Amphibians. Smithsonian Institution Press, Washington, pp. 155-158.

Richter S.C., Young J.E., Seigel R.A. and Johnson G.N. (2001). Postbreeding movements of the dark gopher frog, Rana sevosa Goin and Netting: implications for conservation and management. Journal of Herpetology, 35: 316-321.

Roberts I. (1996). Baw Baw Plateau Vegetation Map. Earth Resource Analysis. Forrest, Victoria.

Roberts J.D., Horwitz P., Wardell-Johnson G., Maxson G. and Mahony M.J. (1997). Taxonomy, relationships and conservation of a new genus and species of myobatrachid frog from the high rainfall region of south-western Australia. Copeia, 1997: 373-381.

Robertson D.R. (1976). Diurnal and lunar periodicity of intestinal calcium in the frog, Rana

289 pipiens. Biochemical Physiology, 54A: 225-231.

Rosenfeld D. (2000). Suppression of rain and snow by urban and industrial air pollution. Science, 287: 1793-1796.

Rosengren N.J., McRae-Williams M.S. and Kraemers S.M. (1981). Sites of Geological Significance in Central Gippsland, Victoria. Environmental Studies Report No. 341, Ministry for Conservation, Melbourne.

Rothermel B.B. and Semlitsch R.D. (2002). An experimental investigation of landscape resistance of forest versus old-field habitats to emigrating juvenile amphibians. Conservation Biology, 16:

Roush W. (1995). When rigor meets reality. Science, 269: 313-315.

Russell K.R., Hanlin H.G., Wigley T.B. and Guynn Jr., D.C. (2002). Responses of isolated wetland herpetofauna to upland forest management. Journal of Wildlife Management, 66: 603-617.

Ryan M.J. (1988). Constraints and patterns in the evolution of anuran acoustic communication. In: B. Fritzsch, M.J. Ryan, W. Wilczynski, T.E. Hetherington and W. Walkowiak, (Eds.), The Evolution of the Amphibian Auditory System. John Wiley and Sons, New York, pp. 637-677.

Ryan M.J. (1991). Sexual selection and communication in frogs. Trends in Ecology and Evolution, 6: 351-354.

Ryser J. (1996). Comparative life histories of a low- and a high-elevation population of the common frog Rana temporaria. Amphibia-Reptilia, 17: 183-195.

Sagor E.S., Ouellet M., Barten E. and Green D.M. (1988). Skeletochronology and geographic variation in age structure in the wood frog, Rana sylvatica. Journal of Herpetology, 32: 469-474.

Salafsky N., Margoluis R., Redford K.H. and Robinson J.G. (2002). Improving the practice of conservation: a conceptual framework and research agenda for conservation science. Conservation Biology, 16: 1469-1479.

Salvador A. and Carrascal L.M. (1990). Reproductive phenology and temporal patterns of mate access in Mediterranean anurans. Journal of Herpetology, 24: 438-441.

Santiago R.R., Duellman W.E., Coloma L.A. and Bustamante M.R. (2003). Population decline of the jambato toad Atelopus ignescens (Anura: Bufonidae) in the Andes of Ecuador. Journal of Herpetology, 37: 116-126.

290 Sartorius S.S. and Rosen P.C. (2000). Breeding phenology of the lowland leopard frog (Rana yavapaiensis): implications for conservation and ecology. The Southwestern Naturalist, 45: 267-273.

Schindler D.W., Curtis P.L., Parker B.R. and Stainton M.P. (1996). Consequences of climate warming and lake acidification for UV-B penetration in North American boreal lakes. Nature, 379: 705-708.

Schwarzkopf L. and Alford R.A. (1996). Desiccation and shelter-site use in a tropical amphibian: comparing toads with physical models. Functional Ecology, 10: 193-200.

Scotts D.J. (1991). Old-growth forests: their ecological characteristics and value to forest- dependent vertebrate fauna of south-east Australia. In: Lunney D., (Ed.), Conservation of Australia's Forest Fauna. The Royal Zoological Society of New South Wales, New South Wales, pp. 147-159.

Scribner K.T., Arntzen J.W., Cruddace N., Oldham R.S. and Burke T. (2001). Environmental correlates of toad abundance and population genetic diversity. Biological Conservation, 98: 201-210.

Seale D.B. (1980). Influence of amphibian larvae on primary production, nutrient flux, and competition in a pond ecosystem. Ecology, 61: 1531-1550.

Seaman D.E. and Powell R.A. (1996). An evaluation of the accuracy of kernel density estimators for home range analysis. Ecology, 77: 2075-2085.

Seaman D.E. and Powell R.A. (1997). Kernel Home Range Estimation Program. Version 4.26, North Carolina State University, North Carolina.

Seebacher F. and Alford R.A. (1999). Movement and microhabitat use of a terrestrial amphibian (Bufo marinus) on a tropical island: seasonal variation and environmental correlates. Journal of Herpetology, 33: 208-214.

Semlitsch R.D. (1983). Structure and dynamics of two breeding populations of the eastern tiger salamander, Ambystoma tigrinum. Copeia, 1983: 608-616.

Seymour R.S., Mahony M.J. and Knowles R. (1995). Respiration of embryos and larvae of the terrestrially breeding frog Kyarranus loveridgei. Herpetologica, 51: 369-376.

Shaffer M.L. (1997). Population viability analyses: determining nature's share. In: Meffe G.K. and Carroll C.R., (Eds.), Principles of Conservation Biology. 2nd edn. Sinauer Associates, Sunderland, Massachusettes, pp. 215-218.

Shealy R.M. (1975). Factors influencing activity in the salamanders Desmognathus ochrophaeus

291 and D. monticola (Plethodontidae). Herpetologica, 31: 94-102.

Shebitz D. (2002). Assessing demographic change for conservation purposes. Conservation Biology, 16: 566-568.

Sibley G.T. (1975). A Study of Land in the Tyers River Catchment. Soil Conservation Authority, Victoria.

Silverman B.W. (1986). Density Estimation for Statistics and Data Analysis. Chapman and Hall, London, United Kingdom.

Sinsch U. (1987). Orientation behaviour of toads (Bufo bufo) displaced from the breeding site. Journal of Comparative Physiology A, 161: 715-727.

Sinsch U. (1988). Seasonal changes in the migratory behaviour of the toad Bufo bufo: direction and magnitude of movements. Oecologia, 76: 390-398.

Sinsch U. (1989). Behavioral thermoregulation of the Andean toad (Bufo spinulosus) at high altitude. Oecologia, 80: 32-38.

Sinsch U. (1990). Migration and orientation in anuran amphibians. Ethology, Ecology and Evolution, 2: 65-79.

Sinsch U. (1991). The orientation behaviour of amphibians. Herpetological Journal, 1: 541-544.

Sinsch U., Di Tada I.E. and Martino A.L. (2001). Longevity, demography and sex-specific growth of the pampa de achala toad, Bufo achalensis CEI, 1972. Studies on Neotropical Fauna and Environments, 36: 95-104.

Skelly D.S., Yurewicz K.L., Werner E.E. and Relyea R.D. (2003). Estimating decline and distributional change in amphibians. Conservation Biology, 17: 744-751.

Smirina E.M. (1972). Annual layers in bones of Rana temporaria. Zoologischeskii Zhurnal, 51: 1529-1534. (In Russian).

Smirina E.M. (1994). Age determination and longevity in amphibians. Gerontology, 40: 113-146.

Smith M.A., Berrill M. and Kapron C.M. (2002). Photolyase activity of the embryo and the ultraviolet absorbence of embryo jelly for several Ontario amphibian species. Canadian Journal of Zoology, 80: 1109-1116.

Smith M.J., Osborne W.S. and Hollis G.J. (1999). Long-term Variability in Precipitation and Temperature on the Baw Baw Plateau: Is there a Link with the Decline of the Endemic Baw Baw Frog (Philoria frosti)? Report Commissioned by the Victorian Department of Natural Resources and Environment, University of Canberra and Department of Natural

292 Resources and Environment.

Soulé M.E. (1985). What is conservation biology? Bioscience, 35: 727-734.

Soulé M.E. and Wilcox B.A. (1980). Conservation Biology: An Evolutionary and Ecological Approach. Sinauer Associates, Sunderland, Massachusetts.

Soulé M.E. (ed.) (1986). Conservation Biology: Science of Scarcity and Diversity. Sinauer Associates, Sunderland, Massachusetts.

Speare R. and Berger L. (2000). Chytridiomycosis in amphibians in Australia. http://www.jcu.edu.au/school/phtm/PHTM/frogs/chyspec.htm. 9 October 2000.

Spencer B. (1901). Two new species of frogs from Victoria. Proceedings of the Royal Society of Victoria, 13: 175-178.

Spieler M. and Linsenmair K.E. (1998). Migration patterns and diurnal use of shelter in a ranid frog of a west African savannah: a telemetric study. Amphibia-Reptilia, 19: 43-64.

Stallard R.F. (2001). Possible environmental factors underlying amphibian decline in eastern Puerto Rico: Analysis of U.S. government data archives. Conservation Biology, 15: 943- 953.

Staub N.L., Brown C.W. and Wake D.B. (1995). Patterns of growth and movements in a population of Ensatina eschscholtzii platensis (Caudata: Plethodontidae) in the Sierra Nevada, California. Journal of Herpetology, 29: 593-599.

Stebbins R.C. and Cohen N.W. (1995). A Natural History of Amphibians. Princeton University Press, Princeton, New Jersey.

Steinberg E.K. and Jordon C.E (1994). Using molecular genetics to learn about the ecology of threatened species: the allure and the illusion of measuring genetic structure in natural populations. In: Fiedler P.L. and Kareiva P.M., (Eds.), Conservation Biology: For the Coming Decade, 2nd edn. Chapman and Hall, New York, pp. 440-460.

Stephenson H. (1980). Cattle and Huts of the High Plains. Stephenson Graphic Work, Armadale, Victoria.

Stirrat S.C., Lawson D., Freeland W.J. and Morton R. (2001). Monitoring Crocodylus porosus populations in the Northern Territory of Australia: a retrospective power analysis. Wildlife Research, 28: 547-554.

Storey K.B. (1990). Life in a frozen state: adaptive strategies for natural freeze tolerance in amphibians and reptiles. American Journal of Physiology, 258: R559-R568.

293 Storfer A. (2003). Amphibian declines: future directions. Diversity and Distributions, 9: 151-163.

Straughan I.R. (1975). An analysis of the mechanism of mating call discrimination in the frogs Hyla regilla and H. cadaverina. Copeia, 1975: 415-424.

Strijbosch H. (1980). Habitat selection by amphibians during their terrestrial phase. British Journal of Herpetology, 6: 93-98.

Sugalski M.T. and Claussen D.L. (1997). Preference for soil moisture, soil pH, and light intensity by the salamander, Plethodon cinereus. Journal of Herpetology, 31: 245-250.

Sullivan B.K. (1992). Sexual selection and calling behavior in the American toad (Bufo americanus). Copeia, 1992: 1-7.

Sullivan B.K. and Fernandez P.J. (1999). Breeding activity, estimated age-structure, and growth in sonoran desert anurans. Herpetologica, 55: 334-343.

Taigen T.L. and Wells K.D. (1985). Energetics of vocalization by an anuran amphibian (Hyla versicolor). Journal of Comparative Physiology, 155: 163-170.

Tárano Z. (2002). Vocal responses to conspecific call variation in the neotropical frog Physalaemus enesefae. Journal of Herpetology, 36: 615-620.

Thomas L. (1997). Retrospective power analysis. Conservation Biology, 11: 276-280.

VDSE (2003). Advisory List of Threatened Vertebrate Fauna in Victoria - 2003. Department of Sustainability and Environment, East Melbourne, Victoria.

Tinsley R.C. and Tocque K. (1995). The population dynamics of a desert anuran, Scaphiopus couchii. Australian Journal of Ecology, 20: 376-384.

Townsend D.S. and Stewart M.M. (1986). Courtship and mating behavior of a Puerto Rican frog, Eleutherodactylus coqui. Herpetologica, 42: 161-166.

Trenerry M.P. (1988). The ecology of tadpoles in a tropical rainforest stream. B. Sc. Honours Thesis. James Cook University.

Trenerry M.P., Laurance W.F. and McDonald K.R. (1994). Further evidence for the precipitous decline of endemic frogs from tropical Australia. Pacific Conservation Biology, 1: 150- 153.

Trenham P.C., Shaffer H.B., Koenig W.D. and Stromberg M.R. (2000). Life history and demographic variation in the California tiger salamander (Ambystoma californiense). Copeia, 2000: 365-377.

294 Tyler M.J. (1989). Australian Frogs. Viking O'Neil, Victoria, Australia.

Tyler M.J. (1991). Declining amphibian populations - a global phenomenon? an Australian perspective. Alytes, 9: 43-50.

Tyler M.J. (1992). Encylopedia of Australian Animals. Frogs. Angus and Robertson, Pymble, New South Wales.

Tyler M.J. (1997). The Action Plan for Australian Frogs. Wildlife Australia, Canberra. van Gelder J.J., Aarts H.M.J. and Staal H.J.W.M. (1986). Routes and speed of migrating toads (Bufo bufo L.): a telemetric study. Herpetological Journal, 1: 111-114.

Vatnick I., Brodkin M.A., Simon M.P., Grant B.W., Conte C.R., Gleave M., Myers R. and Sadoff M.M. (1999). The effects of exposure to mild acidic conditions on adult frogs (Rana pipiens and Rana clamitans): mortality rates and pH preferences. Journal of Herpetology, 33: 370-374.

Vences M., Galán P., Palanca A., Vieites D.R., Nieto S. and Rey J. (2000). Summer microhabitat use and diel activity cycles in a high altitude Pyrenean population of Rana temporaria. Herpetological Journal, 10: 49-56.

Vertucci F.A. and Corn P.S. (1996). Evaluation of episodic acidification and amphibian declines in the Rocky Mountains. Ecological Applications, 6: 449-457.

Vial J.L. and Saylor L. (1993). The Status of Amphibian Populations: Compilation and Analysis. IUCN/SSC Declining Amphibian Populations Taskforce, Work Doc. No. 1. von Bertalanffy L. (1938). A quantitative theory of organic growth. Human Biology, 10: 181-213. von Sacken A. (1998). Interior Habitat. In: Voller J. and Harrison S., (Eds.), Conservation Biology Principles for Forested Landscapes. UBC Press, British Columbia, pp. 131-145.

Wagner Jr., W.E. (1989). Social correlates of variation in male calling behavior in Blanchard's cricket frog, Acris crepitans blanchardi. Ethology, 82: 27-45.

Wake D.B. and Castanet J. (1995). A skeletochronological study of growth and age in relation to adult size in Batrachoseps attenuatus. Journal of Herpetology, 29: 60-65.

Wake D.B. and Morowitz H.J. (1990). Declining amphibian populations - a global phenomenon? Workshop report, National Research Council, Irvine, California.

Walker J. and Tunstall B.R. (1981). Field estimation of foliage cover in Australian woody vegetation. Memorandum 81/19, CSIRO Australian Land Use Research and Technology, Canberra.

295 Walsh N.G., Barley R.H. and Gullan P.K. (1984). The Alpine Vegetation of Victoria (Excluding the Bogong High Plains Region). Environmental Studies Series Report No. 376, Department of Conservation, Forests and Lands, Melbourne.

Walters C.J. (1986). Adaptive Management of Renewable Resources. Macmillan, New York.

Wardell-Johnson G. and Roberts J.D. (1993). Biogeographic barriers in a subdued landscape: the distribution of the Geocrinia rosea (Anura, Myobatrachidae) complex in south-western Australia. Journal of Biogeography, 20: 95-108.

Waters W.F. (1966). The Baw Baws - A short history of the Plateau, from its discovery to the present day. Melbourne Walker, 37: 33-45.

Watson G.F., Littlejohn M.J., Gartside D.F. and Loftus-Hills J.J. (1985). The Litoria ewingi complex (Anura: Hylidae) in south-eastern Australia. VIII. hybridisation between L. ewingi and L. verreauxi alpina in the area, south central Victoria. Australian Journal of Zoology, 33: 143-152.

Watson G.F., Littlejohn M.J., Hero J-M. and Robertson P. (1991). Conservation Status, Ecology and Management of the Spotted Tree Frog (Litoria spenceri). Arthur Rylah Institute Technical Report Series No. 116, Department of Conservation and Environment, Victoria.

Watson G.F. and Martin A.A. (1973). Life history, larval morphology and relationships of leptodactylid frogs. Transactions of the Royal Society of South Australia, 97: 33-45.

Weathers K.C., Cadenasso M.L. and Pickett S.T.A. (2001). Forest edges as nutrient and pollutant concentrators: potential synergisms between fragmentation, forest canopies, and the atmosphere. Conservation Biology, 15: 1506-1514.

Webb C.E. and Joss J. (1997). Does predation by fish Gambusia holbrooki (Atheriniformes: Poeciliidae) contribute to declining frog populations? Australian Zoologist, 30: 316-324.

Webb G.A. (1989). Notes on the biology and conservation of Philoria sphagnicolus (Moore 1958) (Anura: Myobatrachidae). Herpetofauna, 19: 1-6.

Wells K.D. (1977a). The courtship of frogs. In: D.H. Taylor and S.I. Guttman, (Eds.), The Reproductive Biology of Amphibians. Plenum Press, New York, pp. 233-262.

Wells K.D. (1977b). The social behaviour of anuran amphibians. Animal Behaviour, 25: 666-693.

Wells K.D. (1986). The effect of social interactions on calling energetics in the gray treefrog (Hyla versicolor). Ecology and Sociobiology, 106: 265-278.

296 Wells K.D. (1988). The effect of social interactions on anuran vocal behaviour. In: B. Fritzsch, M.J. Ryan, W. Wilczynski, T.E. Hetherington and W. Walkowiak, (Eds.), The Evolution of the Amphibian Auditory System. John Wiley and Sons, New York, pp. 433-454.

Wells K.D. and Taigen T.L. (1989). Calling energetics of a neotropical treefrog, Hyla microcephala. Behavioural Ecology and Sociobiology, 25: 13-22.

Wells K.D., Taigen T.L., Rusch S.W. and Robb C.C. (1995). Seasonal and nightly variation in glycogen reserves of calling gray treefrogs (Hyla versicolor). Herpetologica, 51: 359- 368.

Welsh Jr., H.H. and Lind A.J. (1995). Habitat correlates of the Del Norte salamander, Plethodon elongatus (Caudata: Plethodontidae), in northwestern California. Journal of Herpetology, 29: 198-210.

Welsh Jr., H.H. and Lind A.J. (1996). Habitat correlates of the southern torrent salamander, Rhyacotriton variegatus (Caudata: Rhyacotritonidae), in northwestern California. Journal of Herpetology, 30: 385-398.

Welsh Jr., H.H. and Lind A.J. (2002). Multiscale habitat relationships of stream amphibians in the Klamath-Siskiyou region of California and Oregon. Journal of Wildlife Management, 66: 581-602.

Welsh Jr., H. (1990). Relictual amphibians and old-growth forests. Conservation Biology, 4: 309- 319.

Werner J.K. and McCune M.B. (1979). Seasonal changes in anuran populations in a northern Michigan pond. Journal of Herpetology, 13: 101-104.

Weygoldt P. (1989). Changes in the composition of mountain stream frog communities in the Atlantic Mountains of Brazil: Frogs as indicators of environmental deteriorations? Studies on Neotropical Fauna and Environments, 24: 249-255.

Wheeler B.A., Prosen E., Mathis A. and Wilkinson R.F. (2003). Population declines of a long- lived salamander: a 20+-year study of hellbenders, Cryptobranchus alleganiensis. Biological Conservation, 109: 151-156.

White A.W. and Pyke G.H. (1996). Distribution and conservation status of the green and golden bell frog Litoria aurea in New South Wales. Australian Zoologist, 30: 177-189.

White G.C. and Garrott R.A. (1990). Analysis of Wildlife Radio-Tracking Data. Academic Press, Inc, San Diego, California.

Whiteman H.H., Wissinger S.A. and Brown W.S. (1996). Growth and foraging consequences of

297 facultative paedomorphosis in the tiger salamander, Ambystoma tigrinum nebulosum. Evolutionary Ecology, 10: 433-446.

Wilhere G.F. (2002). Adaptive management in habitat conservation plans. Conservation Biology, 16: 20-29.

Wilkins R.N. and Peterson N.P. (2000). Factors related to amphibian occurrence and abundance in headwater streams draining second-growth Douglas-fir forests in southwestern Washington. Forest Ecology and Management, 139: 79-91.

Williams D.F. and Baker R.M. (1979). Fire Protection Plan Baw Baw National Park. Forests Commission, Melbourne.

Williamson I. and Bull C.M. (1996). Population ecology of the Australian frog Crinia signifera: adults and juveniles. Wildlife Research, 23: 249-266.

Wilson L.D. and McCranie J.R. (1994). Rotenone hazards to amphibians and reptiles. Herpetological Review, 25: 150-153.

Woodgate P.W., Peel W.D., Ritman K.T., Coram J.E., Brady A., Rule A.J. and Banks J.C.G. (1993). A Study of the Old-growth Forests of East Gippsland. Department of Conservation and Natural Resources, Melbourne.

Woolbright L.L. (1985). Patterns of nocturnal movement and calling by the tropical frog Eleutherodactylus coqui. Herpetologica, 41: 1-9.

Woolbright L.L. (1997). Local extinctions of anuran amphibians in the Luquillo experimental forest of northeastern Puerto Rico. Journal of Herpetology, 31: 572-576.

Worrest R.C. and Kimeldorf D.J. (1976). Distortions in amphibian development induced by ultraviolet-B enhanced (290-315 nm) of a simulated solar spectrum. Photochemistry and Photobiology, 24: 377-382.

Worton B.J. (1989). Kernel methods for estimating the utilisation distribution in home-range studies. Zoology, 70: 164-168.

Wyman R.L. (1990). What's happening to the amphibians? Conservation Biology, 4: 350-352.

Zimmerman B.L. (1994). Audio Strip Transects. In: W. Heyer, M.A. Donnelly, R.W. McDiarmid, L-A .C. Hayek and M.S. Foster, (Eds.), Measuring and Monitoring Biological Diversity. Standard Methods for Amphibians. Smithsonian Institution Press, Washington, pp. 92-97.

Zug G.R. and Zug P.B. (1979). The marine toad, Bufo marinus: a natural history resume of native populations. Smithsonian Contribution to Zoology, 248: 1-58.

298 Appendices

Appendix 3.1: Census data of calling males recorded at survey transects at different elevations between 1983 and 2002.

The distribution of transects is classified into elevation groups, denoted by S = sub-alpine (> 1400 m), SM = sub-alpine-montane (1300 - 1400 m), M = montane (960 - 1299 m). For census data, ‘/’ separates census data collected on different days. Data from 1983 and 1984 are from Malone (1985a).

Survey Transect Elevation 1983 1984 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 Group Access Road 1 M 432112110 2 1102 Access Road 2 M 302600000 0 2202 Access Road 3 M 0000000 3 2003 Baragwanath Flat S 167245944011 0/5000 Barnies Creek SM-M-----00 8 3801 Bell Clear Creek SM-M------0 - ---- Bell Clear Creek (below rd) M ------0 - ---- Bell Clear Creek A SM-M------0 ---- Block 10a M ------9--0 Block 10b M ------16--7 Block 10c M ------3--- Block 10d M ------1--1 Block 10 Myrrhee CK M ------0---

299 Survey Transect Elevation 1983 1984 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 Group Block 10e M ------3--- Block 10f M ------0--- Block 10 section a M ------3--- Block 10 section b M ------2--- Block 10 section c M ------2--- Block 10 section d M ------1--- Cascade Creek (below 1300 m A) M ------1 1 0--- Cascade Creek (below 1300 m B) M ------0 - ---- Cascade Creek B M ------0 ---- Cascade Creek frost hollow S-SM-00---- 0 0--- Cascade Creek frost hollow-side branch S-SM------2--- Cascade Creek (side branch A) SM------4--- Chairlift + corner SM-M68----- 0 -525 Chairlift corner M 6801200 0 0000 Charity Creek (left branch) M -----265/03-02/6 Charity Creek (left branch-Ck A) M ------0 0 ---5 Charity Creek (left branch-top) SM-M------3 2 --02 Charity Creek (right branch-above rd) SM-M------5 8 4203 Charity Creek (right branch-below rd) M ------5 0 2322 Charity Creek (bog below sewage farm) M 2 3 - - 1 - Charity Creek (below sewage pond rd) M -----00 0 ---2 Charity Creek (side ck A) M -----12- - ---8 Charity Creek (side ck B) M ------0 2 ---0 Charity Creek (side ck C) M ------0 4 --00

300 Survey Transect Elevation 1983 1984 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 Group Charity Creek (side ck D) SM-M------0 3 2--7 Charity Creek (side ck E) M ------0 3 ---0 Charity Creek (side ck F) SM-M------0 4-00 Charity Creek (side ck G) M ------2--0/4 Charity Creek (up-stream sewage pond rd) M ------6 0 9--3 Charity Creek (right branch-below Ck Corner)M ------0 0 1803 Chinamans 1 M ------6 ---8 Chinamans 2 M ------11/6---- Chinamans 3 M ------8 ---- Chinamans 4 M ------1 ---- Chinamans 5 M ------1 ---- Chinamans 6 SM-M------3 ---- Creek Corner SM524900240 2 2401 Creek Corner 2 SM912--501 2 2409 Currawong Flat/Sandys Flat/Pauciflora Flat S 536-82520 1/1724/0/00/2 Currawong Flat S 174231210000 0000 Currawong Flat ds S ----31------East Tanjil SM71-10020 1 5700 Ellery Creek M -----13------Faith Creek M -----0- 5 ---- Faith Creek (800-1000 m) M ------0 - ---- Faith Creek (left branch) M -----1- 1 ---- Faith Creek (right branch) M -----2------Freemans Flat/Tullicoutty S 41 104 00000 1 210/01

301 Survey Transect Elevation 1983 1984 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 Group Gwinear Flat S 93200000 0 0000/0 Hope Creek S-SM-M-----00 213--2 Hope Creek (access rd 1 above & below rd) M ------2 3 4-46 Hope Creek (access rd 2 above rd) SM-M------0 0 1817 Hope Creek (access rd 2 below rd) M ------124 9011 Hope Creek (access rd 3-above rd) SM-M------4152415011 Hope Creek (access rd 3-below rd) M ------5 2 8003 Hope Creek (access rd 3-right branch) S-SM-M------2 1 -11-1 Hope Creek (left branch) SM-M-----08 2 9-110 Jeep Track 1 SM 0 0 0 ------Jeep Track 2 S 6 0 0 ------Jeep Track 3 S 3 0 0 ------Jeep Track 4 SM 0 0 0 ------La Trobe Plain S 206 - 32045 0 1202 Little Boys SM810----0 0 ---- Little Boys A S-SM6415----0 0 ---- Little Boys Creek (lower) M ------0 ---- Little Boys Creek (upper) M ------0 0 0--- Little Girls Creek M ------0 0 0--- Little Girls Creek A M ------0 0 0--- Little Girls Creek B SM-M------0 ---- Long Creek M -----953199143127 Long Creek (below road) M ------0 - ---- Long Creek (left branch side Ck) M ------21281781838

302 Survey Transect Elevation 1983 1984 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 Group Macallister Plain S 82-52010 0 0000 McMillians Flat S 64 - 00000 0 0-0/00 Moondarra Flat S 225-00000 0 1000 Mt Tyers 1 M -----3------Mt Tyers 2 M -----5------Mt Tyers 3 M -----0------Mt Tyers 4 M -----0------Mustering Flat S 57000000 0 2100 Neulyne Plain S 24-22101 0 0100 Newlands 1 M ------30--- Newlands 2 M ------60--- Newlands 3 M ------6--- Newlands 4 M ------11--- Newlands 5 M ------38--- North Cascade 1 M ------0 ---- North Cascade 2 SM-M ------0 ---- Pudding Basin S 101-21100 0 0000 Rum Creek SM-M------0 ---- Rum Creek (Ezards mid branch) M ------0 0 0--- South Cascade SM-M70--0-1 0 ---- South Cascade A SM 0 0 ----0 0 ---- South Cascade B SM ------0 ---- Swift Creek (above rd) M ------0 0 ---- Talbot Ck (below 1200 m) M ------0 - ----

303 Survey Transect Elevation 1983 1984 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 Group Talbot Ck frost hollow section S 00----0 - ---- Tanjil Plain S 120 - 93331 1 510/00 Tanjil River East Branch SM-M -----0- 2 ---0 Tanjil River Tributary 1 M -----9- 8 5--- Tanjil River Tributary 2 M -----35 2 --911 Tanjil River West Branch 1A (above aqueduct) M -----0- - --88 Tanjil River West Branch 1A (below aqueduct) M ------11/8-17-- Tanjil River West Branch 1B (above aqueduct) M ------6 -13124 Tanjil River West Branch 1B (below aqueduct) M ------76/65--- Tanjil River West Branch 2 (above aqueduct) M -----4183027284713 Tanjil River West Branch 2 (below aqueduct) M ------2 - ---- Tanjil River West Branch 3 (above aqueduct) M -----021124-290 Tanjil River West Branch 4 (above aqueduct) M -----0- 4 1-6- Tanjil River West Branch 5 (above aqueduct) SM-M -----0- 2 --4- Tanjil River West Branch 6 (above aqueduct) SM-M -----1-12---- Tanjil River West Branch 7 (below aqueduct) M ------118/2447--- Tanjil River West Branch 7b (below aqueduct) M ------1 - ---- Tanjil River West Branch 8 (above aqueduct) M ------1 --22- Tanjil River West Branch 9 (above aqueduct) M ------7 --10- Tanjil River West Branch 10 (above aqueduct) M ------1420 Tanjil River West Branch 11 (above aqueduct) M ------150- The Morass SM667-----414/15/156/2945-54 The Morass (with 2 sections not surveyed) SM - - 30 26 10 23 35 31 6/23 44 1/34 49 Thomson River 1 M ------11362---

304 Survey Transect Elevation 1983 1984 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 Group Thomson River 2 M ------3212/125--- Thomson River 3 M ------4 ---- Thomson River 4 M ------0 ---- Thomson River 5 M ------0 ---- Tyers River 1 (right branch) M -----00 1 1-273 Tyers River 1 frost hollow S-SM14900000 0 0000 Tyers River 2 (left branch) M -----0------Tyers River 2 (right branch) M -----0------Tyers River 2 frost hollow SM7300000 0 0000 Tyers River 3 M ------71---- Tyers River 4 M ------6 ---- Tyers River 5 M ------19---- Tyers River East Branch (tributary 1 ) M ------125-177 Tyers River East Branch (tributary 2 ) M ------121-0/00 Tyers River East Branch (tributary 3 ) M ------3 4873 Tyers River East Branch (tributary) M -----00 1 010/00 Tyers River West Branch (tributary A) M ------11---- Tyers River West Branch (tributary B) M ------32---- Tyers River West Branch (tributary C) M ------21---- Village Flat/Dam Valley/Big Hill S 183 149 10000 0 0000 Village Flat 2 S 24 - - - 2 1 0 0 1 0 - 0 Whitelaw 1 S 0-210------Whitelaw 10 SM17-1------Whitelaw 11 S --0------

305 Survey Transect Elevation 1983 1984 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 Group Whitelaw 12 SM-M--100------Whitelaw 13 SM--00------Whitelaw 2 S-SM67-2134------Whitelaw 3 SM-2802------Whitelaw 4 SM39-010------Whitelaw 5 S-SM53-112------Whitelaw 6 S-SM49-001------Whitelaw 6a SM---2------Whitelaw 7 S-SM96-010------Whitelaw 8 S 30-000------Whitelaw 9 S-SM90-0------Whitelaw Creek M ------0 0 ---- Whitelaw Creek (frost hollow) S-SM368-161------Whitelaw Creek (lower section) M ------0 - 0--- Whitelaw Creek (right branch A) SM-M------0 0 ---- Whitelaw Creek (right branch A-lower section)M ------0 - ---- Whitelaw Creek (right branch B) SM-M------0 0 5--- Whitelaw Creek (right branch C) M ------0--- Whitelaw Creek 1 SM-M------0 0--- Whitelaw Creek 2 M ------0--- Whitelaw Creek Middle Branch M ------0 0 ---- Whitelaw Ruins S164-100------Wombat Flat SM - 18 1 - - - 21 2 16 - - 5 Wombat Flat A S 0-----1 - ----

306 Survey Transect Elevation 1983 1984 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 Group X-Sandys Flat SM-----00 1 4000

307

Appendix 3.2: Seasonally-corrected census data of calling males (C prefix) and estimated level of participation (%) by calling males (P prefix) for survey transects censused between 1993 and 1999.

Year identifiers are abbreviated as: 93 = 1993, 94 = 1994, etc. ‘/’ separates data from different days. The distribution of transects is classified into elevational groups, denoted by S = sub-alpine (> 1400 m), SM = sub-alpine-montane (1300 - 1400 m) and M = montane (960 - 1299 m).

Survey Transect Elevation P93 C93 P94 C94 P95 C95 P96 C96 P97 C97 P98 C98 P99 C99 Group Access Road 1 M 28.0 3.6 45.0 4.4 27.0 3.7 42.5 2.4 48.0 0.0 51.0 3.9 60.5 1.7 Access Road 2 M 42.0 0.0 40.5 0.0 27.0 0.0 42.5 0.0 48.0 0.0 51.0 0.0 60.5 3.3 Access Road 3 M 42.0 0.0 31.0 0.0 27.0 0.0 42.5 0.0 48.0 0.0 44.5 6.7 60.5 3.3 Baragwanath Flat S 76.0 11.8 40.5 9.9 45.0 8.9 53.0 0.0 34.5 2.9 48.0 2.1 81.0 6.2 Barnies Creek SM-M ------49.0 0.0 67.5 0.0 60.0 13.3 41.0 7.3 Bell Clear Creek SM-M ------31.50.0- - - - Bell Clear Creek (below rd) M ------31.5 0.0 - - - - Bell Clear Creek A SM-M------60.50.0- - Block 10a M ------52.05.8 Block 10b M ------52.015.4 Block 10c M ------52.07.7 Block 10d M ------52.01.9 Block 10 Myrrhee CK M ------54.00.0 Block 10e M ------52.05.8 Block 10f M ------52.00.0 Block 10 section a M ------52.0 5.8

309 Survey Transect Elevation P93 C93 P94 C94 P95 C95 P96 C96 P97 C97 P98 C98 P99 C99 Group Block 10 section b M ------52.03.8 Block 10 section c M ------52.0 3.8 Block 10 section d M ------52.0 1.9 Cascade Creek (below 1300 m A) M ------10.0 10.0 60.5 1.7 11.0 0.0 Cascade Creek (below 1300 m B) M ------0.0- - - - Cascade Creek B M ------60.50.0- - Cascade Creek frost hollow S-SM 62.0 0.0 ------60.5 0.0 - - Cascade Creek frost hollow-side branch S-SM ------11.0 18.2 Cascade Creek (side branch A) SM ------11.0 36.0 Chairlift + corner SM-M ------44.50.0- - Chairlift corner M 42.0 0.0 25.0 4.0 43.5 4.6 42.5 0.0 48.0 0.0 44.5 0.0 60.5 0.0 Charity Creek (left branch) M ------49.5 4.0 69.0 8.7 59.5 8.4 60.5 5.0 Charity Creek (left branch-Ck A) M ------69.0 0.0 59.5 0.0 - - Charity Creek (left branch-top) SM-M ------69.0 4.4 59.5 3.4 - - Charity Creek (right branch-above rd) SM-M ------48.0 10.4 51.0 15.7 18.5 21.6 Charity Creek (right branch-below rd) M ------67.5 7.4 59.5 0.0 18.5 10.8 Charity Creek (bog below sewage farm) M ------67.53.060.56.6- - Charity Creek (below sewage pond rd) M ------52.5 0.0 5.0 0.0 51.0 0.0 - - Charity Creek (side ck A) M ------49.5 24.2 ------Charity Creek (side ck B) M ------69.0 0.0 59.5 3.4 - - Charity Creek (side ck C) M ------69.00.059.56.7- - Charity Creek (side ck D) SM-M ------69.0 0.0 59.5 5.0 60.5 3.3 Charity Creek (side ck E) M ------69.0 0.0 59.5 6.7 - - Charity Creek (side ck F) SM-M ------59.5 0.0 60.5 6.6

310 Survey Transect Elevation P93 C93 P94 C94 P95 C95 P96 C96 P97 C97 P98 C98 P99 C99 Group Charity Creek (side ck G) M ------60.53.3 Charity Creek (up-stream sewage pond rd) M ------48.0 12.5 51.0 0.0 60.5 14.9 Charity Creek (right branch-below Ck Corner) M ------58.0 0.0 69.5 0.0 60.5 1.7 Chinamans 1 M ------37.016.2- - Chinamans 2 M ------40.0/37.0 43.7 - - Chinamans 3 M ------40.020.0- - Chinamans 4 M ------35.02.9- - Chinamans 5 M ------35.02.9- - Chinamans 6 SM-M------23.013.0- - Creek Corner SM 28.0 0.0 36.5 0.0 40.0 5.0 50.0 8.0 62.0 0.0 51.0 3.9 18.5 10.8 Creek Corner 2 SM - - - - 40.0 12.5 50.0 0.0 62.0 1.6 51.0 3.9 18.5 10.8 Currawong Flat/Sandys Flat/Pauciflora Flat S 78.0 10.3 45.0 4.4 42.0 11.9 54.0 3.7 41.0 0.0 48.0/60.5 3.8 81.5 8.6 Currawong Flat S 76.0 2.6 45.0 2.2 43.5 0.0 54.0 0.0 41.0 0.0 48.0 0.0 81.5 2.5 Currawong Flat ds S - - - - 43.5 6.9 54.0 1.9 ------East Tanjil SM 78.0 1.3 48.0 0.0 44.5 0.0 54.0 3.7 57.5 0.0 60.0 1.7 80.0 6.3 Ellery Creek M ------53.024.5------Faith Creek M ------50.0 0.0 - - 13.5 37.0 - - Faith Creek (800-1000 m) M ------25.5 0.0 - - - - Faith Creek (left branch) M ------50.0 2.0 - - 13.5 7.4 - - Faith Creek (right branch) M ------53.0 3.8 ------Freemans Flat/Tullicoutty S 68.0 0.0 50.0 0.0 22.0 0.0 53.0 0.0 25.5 0.0 60.0 1.7 75.5 2.6 Gwinear Flat S 66.0 0.0 32.5 0.0 22.0 0.0 44.5 0.0 69.0 0.0 60.0 0.0 36.5 0.0 Hope Creek S-SM-M ------49.0 0.0 67.5 0.0 60.0 3.3 41.0 31.7 Hope Creek (access rd 1 above & below rd) M ------48.0 4.2 51.0 5.9 60.5 6.6

311 Survey Transect Elevation P93 C93 P94 C94 P95 C95 P96 C96 P97 C97 P98 C98 P99 C99 Group Hope Creek (access rd 2 above rd) SM-M ------48.0 0.0 44.5 0.0 60.5 1.7 Hope Creek (access rd 2 below rd) M ------48.0 25.0 51.0 7.8 60.5 14.9 Hope Creek (access rd 3-above rd) SM-M ------48.0 8.3 44.5 33.7 60.5 39.7 Hope Creek (access rd 3-below rd) M ------67.5 7.4 51.0 3.9 60.5 13.2 Hope Creek (access rd 3-right branch) S-SM-M ------67.5 3.0 60.5 1.7 - - Hope Creek (left branch) SM-M ------49.0 0.0 67.5 11.9 60.0 3.3 41.0 22.0 Jeep Track 1 SM 73.0 0.0 ------Jeep Track 2 S 73.0 0.0 ------Jeep Track 3 S 73.0 0.0 ------Jeep Track 4 SM 73.00.0------La Trobe Plain S 76.0 4.0 36.5 5.5 44.0 0.0 49.0 8.2 67.5 7.4 55.0 0.0 79.0 1.3 Little Boys SM ------62.0 0.0 48.0 0.0 - - Little Boys A S-SM ------62.0 0.0 48.0 0.0 - - Little Boys Creek (lower) M ------18.00.0- - Little Boys Creek (upper) M ------20.0 0.0 18.0 0.0 17.0 0.0 Little Girls Creek M ------25.5 0.0 58.5 0.0 17.0 0.0 Little Girls Creek A M ------25.5 0.0 58.5 0.0 17.0 0.0 Little Girls Creek B SM-M------58.50.0- - Long Creek M ------44.5 46.6 66.0 80.3 60.5 31.4 51.0 17.6 Long Creek (below road) M ------25.5 0.0 - - - - Long Creek (left branch side Ck) M ------66.0 31.8 60.5 46.3 51.0 33.3 Macallister Plain S 77.0 6.5 40.5 4.9 44.0 0.0 54.0 1.9 57.5 0.0 55.0 0.0 79.0 0.0 McMillians Flat S 78.0 0.0 48.0 0.0 40.0 0.0 44.0 0.0 57.5 0.0 48.0 0.0 43.0 0.0 Moondarra Flat S 77.0 0.0 38.0 0.0 44.0 0.0 54.0 0.0 34.5 0.0 55.0 0.0 79.0 1.3

312 Survey Transect Elevation P93 C93 P94 C94 P95 C95 P96 C96 P97 C97 P98 C98 P99 C99 Group Mt Tyers 1 M ------45.06.7------Mt Tyers 2 M ------45.0 11.1 ------Mt Tyers 3 M ------45.00.0------Mt Tyers 4 M ------45.00.0------Mustering Flat S 66.0 0.0 50.0 0.0 41.0 0.0 44.5 0.0 69.0 0.0 60.0 0.0 36.5 5.5 Neulyne Plain S 76.0 2.6 36.5 5.5 44.0 2.3 49.0 0.0 67.5 1.5 55.0 0.0 79.0 0.0 Newlands 1 M ------52.0 57.7 Newlands 2 M ------52.0 115.4 Newlands 3 M ------54.0 11.1 Newlands 4 M ------54.020.4 Newlands 5 M ------52.0 73.1 North Cascade 1 M ------25.0 0.0 - - North Cascade 2 SM-M ------25.0 0.0 - - Pudding Basin S 77.0 2.6 40.5 2.5 44.0 2.3 50.0 0.0 34.5 0.0 55.0 0.0 79.0 0.0 Rum Creek SM-M------40.00.0- - Rum Creek (Ezards mid branch) M ------7.0 0.0 40.0 0.0 46.0 0.0 South Cascade SM-M - - - - 20.0 0.0 - - 64.0 1.6 59.5 0.0 - - South Cascade A SM ------64.0 0.0 59.5 0.0 - - South Cascade B SM ------59.5 0.0 - - Swift Creek (above rd) M ------20.00.018.00.0- - Talbot Ck (below 1200 m) M ------10.00.0- - - - Talbot Ck frost hollow S ------64.0 0.0 - - - - Tanjil Plain S 78.0 11.5 50.5 5.9 40.0 7.5 54.0 5.6 34.5 2.9 55.0 1.8 81.0 2.5 Tanjil River East Branch SM-M ------53.0 0.0 - - 16.0 12.5 - -

313 Survey Transect Elevation P93 C93 P94 C94 P95 C95 P96 C96 P97 C97 P98 C98 P99 C99 Group Tanjil River Tributary 1 M ------49.5 18.2 - - 25.0 32.0 - - Tanjil River Tributary 2 M ------49.5 6.1 69.0 7.3 27.5 7.3 11.0 45.5 Tanjil River West Branch 1A (above aqueduct) M ------10.0 0.0 ------Tanjil River West Branch 1A (below aqueduct) M ------31.5 3.2 59.5/60.5 14.9 - - Tanjil River West Branch 1B (above aqueduct) M ------59.5 10.1 - - Tanjil River West Branch 1B (below aqueduct) M ------31.5 22.2 58.5/61.0 20.1 33.0 15.2 Tanjil River West Branch 2 (above aqueduct) M ------10.0 40.0 36.0 50.0 59.5 50.4 33.0 81.8 Tanjil River West Branch 2 (below aqueduct) M ------36.0 5.6 - - - - Tanjil River West Branch 3 (above aqueduct) M ------10.0 0.0 36.0 5.6 59.5 18.5 33.0 72.7 Tanjil River West Branch 4 (above aqueduct) M ------10.0 0.0 - - 59.5 6.7 33.0 3.0 Tanjil River West Branch 5 (above aqueduct) SM-M ------10.0 0.0 - - 57.0 3.5 - - Tanjil River West Branch 6 (above aqueduct) SM-M ------10.0 10.0 - - 57.0 21.1 - - Tanjil River West Branch 7 (below aqueduct) M ------31.5 34.9 60.5/61.0 52.6 33.0 142.4 Tanjil River West Branch 7b (below aqueduct) M ------31.5 3.2 - - - - Tanjil River West Branch 8 (above aqueduct) M ------16.0 6.3 - - Tanjil River West Branch 9 (above aqueduct) M ------57.0 12.3 - - Tanjil River West Branch 10 (above aqueduct) M ------Tanjil River West Branch 11 (above aqueduct) M ------The Morass SM 78.0 38.5 47.0 55.3 25.5 39.2 44.5 51.7 55.0 74.6 60/57/58.5 58.6 13.5/80.0 228.4 The Morass (with 2 sections not surveyed) SM 78.0 38.5 47.0 55.3 25.5 39.2 44.5 51.7 55.0 63.6 60/57/58.5 53.0 13.5/80.0 176.6 Thomson River 1 M ------36.0 30.6 35.0 8.6 54.0/80.0 114.8 Thomson River 2 M ------36.0 88.9 35.0/23.0 38.6 54.0/80.0 46.3 Thomson River 3 M ------23.0 17.4 - - Thomson River 4 M ------27.50.0- -

314 Survey Transect Elevation P93 C93 P94 C94 P95 C95 P96 C96 P97 C97 P98 C98 P99 C99 Group Thomson River 5 M ------27.50.0- - Tyers River 1 (right branch) M ------25.5 0.0 20.0 0.0 12.0 8.3 46.0 2.2 Tyers River 1 frost hollow S-SM 68.0 0.0 32.5 0.0 20.0 0.0 44.5 0.0 64.0 0.0 59.5 0.0 51.0 0.0 Tyers River 2 (left branch) M ------1.0 0.0 ------Tyers River 2 (right branch) M ------1.0 0.0 ------Tyers River 2 frost hollow SM 68.0 0.0 32.5 0.0 20.0 0.0 44.5 0.0 66.0 0.0 59.5 0.0 51.0 0.0 Tyers River 3 M ------61.0 116.4 - - Tyers River 4 M ------61.09.8- - Tyers River 5 M ------61.031.2- - Tyers River East Branch (tributary 1 ) M ------51.0 23.5 46.0 10.9 Tyers River East Branch (tributary 2 ) M ------51.0 23.5 46.0 2.2 Tyers River East Branch (tributary 3 ) M ------51.0 5.9 46.0 8.7 Tyers River East Branch (tributary) M ------25.5 0.0 20.0 0.0 12.0 8.3 46.0 0.0 Tyers River West Branch (tributary A) M ------51.0 21.6 - - Tyers River West Branch (tributary B) M ------51.0 62.8 - - Tyers River West Branch (tributary C) M ------57.0 36.8 - - Village Flat/Dam Valley/Big Hill S 42.0 2.4 38.0 0.0 40.0 0.0 52.5 0.0 41.0 0.0 44.5 0.0 80.0 0.0 Village Flat 2 S 29.0 6.9 52.5 1.9 57.5 0.0 51.0 0.0 80.0 1.3 Whitelaw 1 S 44.0 4.6 22.0 4.6 13.0 0.0 ------Whitelaw 10 SM 34.0 2.9 ------Whitelaw 11 S 44.00.0------Whitelaw 12 SM-M 44.0 2.3 22.0 0.0 15.0 0.0 ------Whitelaw 13 SM 39.0 0.0 22.0 ------Whitelaw 2 S-SM 39.0 5.1 22.0 59.1 13.0 30.8 ------

315 Survey Transect Elevation P93 C93 P94 C94 P95 C95 P96 C96 P97 C97 P98 C98 P99 C99 Group Whitelaw 3 SM 39.0 0.0 17.0 11.8 ------Whitelaw 4 SM 39.0 0.0 17.0 5.9 13.0 0.0 ------Whitelaw 5 S-SM 39.0 2.6 17.0 5.9 13.0 15.4 ------Whitelaw 6 S-SM 34.0 0.0 17.0 0.0 13.0 7.7 ------Whitelaw 6a SM - - 17.0 11.8 ------Whitelaw 7 S-SM 34.0 0.0 17.0 5.9 12.0 0.0 ------Whitelaw 8 S 34.0 0.0 11.0 0.0 12.0 0.0 ------Whitelaw 9 S-SM34.00.0------Whitelaw Creek M ------64.0 0.0 12.0 0.0 - - Whitelaw Creek (frost hollow) S-SM 34.0 2.9 11.0 54.6 12.0 8.3 ------Whitelaw Creek (lower section) M ------64.0 0.0 - - - - Whitelaw Creek (right branch A) SM-M ------69.0 0.0 61.0 0.0 - - Whitelaw Creek (right branch A-lower section) M ------66.0 0.0 - - - - Whitelaw Creek (right branch B) SM-M ------69.0 0.0 61.0 0.0 75.5 6.6 Whitelaw Creek (right branch C) M ------75.5 0.0 Whitelaw Creek 1 SM-M ------12.0 0.0 75.5 0.0 Whitelaw Creek 2 M ------75.50.0 Whitelaw Creek Middle Branch M ------62.0 0.0 27.5 0.0 - - Whitelaw Ruins S 39.0 2.6 17.0 0.0 12.0 0.0 ------Wombat Flat SM 68.0 1.5 ------66.0 31.8 59.5 3.4 36.5 43.8 Wombat Flat A S ------66.01.5- - - - X-Sandys Flat SM ------54.0 0 41.0 0.0 60.5 1.6 81.5 4.9

316 Appendix 3.3: Date in which each survey transect was censused between 1993 and 2002.

Survey Transect 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 Access Road 1 15-Nov-93 23-Nov-94 2-Nov-95 1-Nov-96 31-Oct-97 16-Oct-98 25-Oct-99 9-Nov-00 8-Oct-01 20-Oct-02 Access Road 2 17-Nov-93 5-Dec-94 2-Nov-95 1-Nov-96 31-Oct-97 16-Oct-98 25-Oct-99 9-Nov-00 28-Sep-01 16-Nov-02 Access Road 3 17-Nov-93 16-Nov-94 2-Nov-95 1-Nov-96 31-Oct-97 14-Oct-98 25-Oct-99 6-Nov-00 28-Sep-01 16-Nov-02 Baragwanath Flat 29-Nov-93 5-Dec-94 19-Nov-95 13-Nov-96 29-Oct-97 15-Oct-98 12-Oct-99 21-Oct-00 29-Sep-01 22-Oct-02 Barnies Creek - - - 6-Nov-96 4-Nov-97 24-Oct-98 2-Nov-99 30-Oct-00 3-Nov-01 22-Oct-02 Bell Clear Creek - - - - 20-Nov-97 - - - - - Bell Clear Creek (below rd) - - - - 20-Nov-97 - - - - - Bell Clear Creek A - - - - - 29-Oct-98 - - - - Block 10a ------29-Oct-99 - - 21-Nov-02 Block 10b ------29-Oct-99 - - 21-Nov-02 Block 10c ------29-Oct-99 - - - Block 10d ------29-Oct-99 - - 21-Nov-02 Block 10 Myrrhee Creek ------28-Oct-99 - - - Block 10e ------29-Oct-99 - - - Block 10f ------29-Oct-99 - - - Block 10 section a ------29-Oct-99 - - - Block 10 section b ------29-Oct-99 - - - Block 10 section c ------29-Oct-99 - - - Block 10 section d ------29-Oct-99 - - - Cascade Creek (below 1300 m A) - - - - 25-Nov-97 30-Oct-98 14-Nov-99 - - - Cascade Creek (below 1300 m B) - - - - 25-Nov-97 - - - - - Cascade Creek B - - - - - 30-Oct-98 - - - - Cascade Creek frost hollow 3-Dec-93 - - - - 30-Oct-98 - - - - Cascade Creek frost hollow-side branch ------14-Nov-99 - - - Cascade Creek (side branch A) ------14-Nov-99 - - - Chairlift + corner - - - - - 14-Oct-98 - 6-Nov-00 28-Sep-01 6-Nov-02 Chairlift corner 17-Nov-93 14-Nov-94 15-Nov-95 1-Nov-96 31-Oct-97 14-Oct-98 25-Oct-99 6-Nov-00 28-Sep-01 6-Nov-02

317 Survey Transect 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 Charity Creek (left branch) - - - 7-Nov-96 6-Nov-97 23-Oct-98 25-Oct-99 - 22-Oct-01 17-Oct/9-Nov-02 Charity Creek (left branch-Ck A) - - - - 6-Nov-97 23-Oct-98 - - - 23-Nov-02 Charity Creek (left branch-top) - - - - 6-Nov-97 23-Oct-98 - - 22-Oct-01 9-Nov-02 Charity Creek (right branch-above rd) - - - - 31-Oct-97 16-Oct-98 10-Nov-99 8-Nov-00 26-Sep-01 16-Nov-02 Charity Creek (right branch-below rd) - - - - 4-Nov-97 23/25-Oct-98 10-Nov-99 8-Nov-00 8-Oct-01 17-Oct-02 Charity Creek (bog below sewage farm) - - - - 4-Nov-97 23-Oct-98 - - 28-Sep-01 17-Oct-02 Charity Creek (below sewage pond Rd) - - - 12-Nov-96 31-Oct-97 16-Oct-98 - - - 6-Nov-02 Charity Creek (side ck A) - - - 7-Nov-96 - - - - - 6-Nov-02 Charity Creek (side ck B) - - - - 6-Nov-97 23-Oct-98 - - - 17-Oct-02 Charity Creek (side ck C) - - - - 6-Nov-97 23-Oct-98 - - 22-Oct-01 9-Nov-02 Charity Creek (side ck D) - - - - 6-Nov-97 23-Oct-98 25-Oct-99 - - 9-Nov-02 Charity Creek (side ck E) - - - - 6-Nov-97 23-Oct-98 - - - 9-Nov-02 Charity Creek (side ck F) - - - - - 23-Oct-98 25-Oct-99 - 22-Oct-01 9-Nov-02 Charity Creek (side ck G) ------25-Oct-99 - - 17-Oct/9-Nov-02 Charity Creek (up-stream sewage pond rd) - - - - 31-Oct-97 16-Oct-98 25-Oct-99 - 8-Oct-01 20-Oct-02 Charity Creek (right branch-below ck corner) - - - - 31-Oct-97 23-Oct-98 25-Oct-99 8-Nov-00 28-Sep-01 17-Oct-02 Chinamans 1 - - - - - 15-Nov-98 - - - 22-Nov-02 Chinamans 2 - - - - - 14/15-Nov-98 - - - - Chinamans 3 - - - - - 14-Nov-98 - - - - Chinamans 4 - - - - - 16-Nov-98 - - - - Chinamans 5 - - - - - 16-Nov-98 - - - - Chinamans 6 - - - - - 21-Nov-98 - - - - Creek Corner 15-Nov-93 6-Dec-94 10-Nov-95 8-Nov-96 10-Nov-97 16-Oct-98 10-Nov-99 7/6-Nov-00 28-Sep-01 16-Nov-02 Creek Corner 2 - - 10-Nov-95 8-Nov-96 10-Nov-97 16-Oct-98 10-Nov-99 6-Nov-00 28-Sep-01 16-Nov-02 Currawong Flat/Sandys Flat/Pauciflora Flat 25-Nov-93 23-Nov-94 22-Nov-95 14-Nov-96 30-Oct-97 15/25oct 16-Oct-99 1-Nov-00 29-Sep/5/10-Oct-02 10/22-Oct-02 Currawong Flat 29-Nov-93 23-Nov-94 21-Nov-95 14-Nov-96 30-Oct-97 15-Oct-98 16-Oct-99 1-Nov-00 5/15-Oct-01 - Currawong Flat ds - - 21-Nov-95 14-Nov-96 ------East Tanjil 26-Nov-93 25-Nov-94 20-Nov-95 14-Nov-96 12-Nov-97 4-Nov-98 13-Oct-99 6-Nov-00 29-Oct-01 28-Oct-02 Ellery Creek - - - 16-Nov-96 ------Faith Creek - - - 8-Nov-96 - 25-Nov-98 - - - -

318 Survey Transect 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 Faith Creek (800-1000 m) - - - - 21-Nov-97 - - - - - Faith Creek (left branch) - - - 8-Nov-96 - 25-Nov-98 - - - - Faith Creek (right branch) - - - 16-Nov-96 ------Freeman’s Flat/Tullicoutty 1-Dec-93 27-Nov-94 2-Dec-95 16-Nov-96 28-Oct-97 4-Nov-98 17-Oct-99 7-Nov-00 5/15-Oct-01 28-Oct-02 Gwinear Flat 2-Dec-93 7-Dec-94 2-Dec-95 26-Nov-96 5-Nov-97 4-Nov-98 3-Nov-99 31-Oct-00 5-Oct-01 11/12-Oct-02 Hope Creek - - - 6-Nov-96 4-Nov-97 24-Oct-98 2-Nov-99 - - - Hope Creek (access rd 1 above & below rd) - - - - 31-Oct-97 16-Oct-98 25-Oct-99 - 8-Oct-01 20-Oct-02 Hope Creek (access rd 2 above rd) - - - - 31-Oct-97 14-Oct-98 25-Oct-99 9-Nov-00 28-Sep-01 16-Nov-02 Hope Creek (access rd 2 below rd) - - - - 31-Oct-97 16-Oct-98 25-Oct-99 9-Nov-00 - 6-Nov-02 Hope Creek (access rd 3-above rd) - - - - 31-Oct-97 14-Oct-98 25-Oct-99 6-Nov-00 28-Sep-01 16-Nov-02 Hope Creek (access rd 3-below rd) - - - - 4-Nov-97 16-Oct-98 25-Oct-99 8-Nov-00 27-Sep-01 20-Oct-02 Hope Creek (access rd 3-right branch) - - - - 4-Nov-97 25-Oct-98 10-Nov-99 8/6-Nov-00 - 20-Oct-02 Hope Creek (left branch) - - - 6-Nov-96 4-Nov-97 24-Oct-98 2-Nov-99 - 27-Sep-01 10-Nov-02 Jeep Track 1 30-Nov-93 ------Jeep Track 2 30-Nov-93 ------Jeep Track 3 30-Nov-93 ------Jeep Track 4 30-Nov-93 ------La Trobe Plain 23-Nov-93 18-Nov-94 16-Nov-95 6-Nov-96 4-Nov-97 18-Oct-98 14-Oct-99 - 3-Nov-01 22-Nov-02 Little Boys - - - - 10-Nov-97 11-Nov-98 - - - - Little Boys A - - - - 10-Nov-97 11-Nov-98 - - Little Boys Creek (lower) - - - - - 23-Nov-98 - - - - Little Boys Creek (upper) - - - - 22-Nov-97 23-Nov-98 11-Nov-99 - - - Little Girls Creek - - - - 21-Nov-97 6-Nov-98 11-Nov-99 - - - Little Girls Creek A - - - - 21-Nov-97 6-Nov-98 11-Nov-99 - - - Little Girls Creek B - - - - - 6-Nov-98 - - - - Long Creek - - - 16/26-Nov-96 8-Nov-97 30-Oct-98 30-Oct-99 16-Nov-00 8-Nov-01 7-Nov-02 Long Creek (below road) - - - - 21-Nov-97 - - - - - Long Creek (left branch side Ck) - - - - 8-Nov-97 30-Oct-98 30-Oct-99 16-Nov-00 8-Nov-01 7-Nov-02 Macallister Plain 24-Nov-93 5-Dec-94 16-Nov-95 14-Nov-96 12-Nov-97 18-Oct-98 14-Oct-99 30-Oct-00 6-Oct-01 28-Oct-02 McMillians Flat 25-Nov-93 25-Nov-94 10-Nov-95 25-Nov-96 12-Nov-97 15-Oct-98 31-Oct-99 21-Oct-00 6-Oct-01 8-Nov-02

319 Survey Transect 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 Moondarra Flat 24-Nov-93 19-Nov-94 16-Nov-95 14-Nov-96 29-Oct-97 18-Oct-98 14-Oct-99 30-Oct-00 29-Sep/6-Oct-01 28-Oct-02 Mt Tyers 1 - - - 27-Nov-96 ------Mt Tyers 2 - - - 27-Nov-96 ------Mt Tyers 3 - - - 27-Nov-96 ------Mt Tyers 4 - - - 27-Nov-96 ------Mustering Flat 2-Dec-93 27-Nov-94 23-Nov-95 26-Nov-96 5-Nov-97 4-Nov-98 3-Nov-99 31-Oct-00 15-Oct-01 12-Oct-02 Neulyne Plain 23-Nov-93 18-Nov-94 16-Nov-95 6-Nov-96 4-Nov-97 18-Oct-98 14-Oct-99 29-Oct-00 3-Nov-01 11-Oct-02 Newlands 1 ------29-Oct-99 - - - Newlands 2 ------29-Oct-99 - - - Newlands 3 ------28-Oct-99 - - - Newlands 4 ------28-Oct-99 - - - Newlands 5 ------29-Oct-99 - - - North Cascade 1 - - - - - 20-Nov-98 - - - - North Cascade 2 - - - - 20-Nov-98 - - - - Pudding Basin 24-Nov-93 5-Dec-94 16-Nov-95 8-Nov-96 29-Oct-97 18-Oct-98 14-Oct-99 29-Oct-00 29-Sep-01 28-Oct-02 Rum Creek - - - - - 14-Nov-98 - - - - Rum Creek (Ezards mid branch) - - - - 24-Nov-97 14-Nov-98 1-Nov-99 - - - South Cascade - - 3-Dec-95 - 9-Nov-97 5-Nov-98 - - - - South Cascade A - - - - 9-Nov-97 5-Nov-98 - - - - South Cascade B - - - - - 5-Nov-98 - - - - Swift Creek (above rd) - - - - 22-Nov-97 23-Nov-98 - - - - Talbot Ck (below 1200 m) - - - - 25-Nov-97 - - - - - Talbot Ck frost hollow - - - 9-Nov-97 - - - - - Tanjil Plain 25-Nov-93 28-Nov-94 10-Nov-95 14-Nov-96 29-Oct-97 18-Oct-98 12-Oct-99 21-Oct-00 6-Oct/3-Dec-01 28-Oct-02 Tanjil River East Branch - - - 16-Nov-96 - 24-Nov-98 - - - - Tanjil River Tributary 1 - - - 7-Nov-96 - 20-Nov-98 - - - - Tanjil River Tributary 2 - - - 7-Nov-96 6-Nov-97 19-Nov-98 14-Nov-99 - 22-Oct-01 9-Nov-02 Tanjil River West Branch 1A (above aqueduct) - - - 11-Dec-96 - - - 22-Oct-00 21-Oct-01 19-Oct-02 Tanjil River West Branch 1A (below aqueduct) - - - - 20-Nov-97 29-Oct/5-Nov-98 - - - - Tanjil River West Branch 1B (above aqueduct) - - - - - 5-Nov-98 - 22-Oct-00 21-Oct-01 19-Oct-02

320 Survey Transect 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 Tanjil River West Branch 1B (below aqueduct) - - - - 20-Nov-97 2-6/11/98 4-Nov-99 - - - Tanjil River West Branch 2 (above aqueduct) - - - 11-Dec-96 19-Nov-97 5-Nov-98 4-Nov-99 22-Oct-00 - 19-Oct-02 Tanjil River West Branch 2 (below aqueduct) - - - - 19-Nov-97 - - - - - Tanjil River West Branch 3 (above aqueduct) - - - 11-Dec-96 19-Nov-97 5-Nov-98 4-Nov-99 - 9-Oct-01 19-Oct-02 Tanjil River West Branch 4 (above aqueduct) - - - 11-Dec-96 - 5-Nov-98 4-Nov-99 - 9-Oct-01 - Tanjil River West Branch 5 (above aqueduct) - - - 12-Dec-96 - 7-Nov-98 - - 21-Oct-01 - Tanjil River West Branch 6 (above aqueduct) - - - 12-Dec-96 - 7-Nov-98 - - - - Tanjil River West Branch 7 (below aqueduct) - - - - 20-Nov-97 29-Oct/2-Nov-98 4-Nov-99 - - - Tanjil River West Branch 7b (below aqueduct) - - - - 20-Nov-97 - - - - - Tanjil River West Branch 8 (above aqueduct) - - - - - 24-Nov-98 - - 21-Oct-01 - Tanjil River West Branch 9 (above aqueduct) - - - - - 7-Nov-98 - - 21-Oct-01 - Tanjil River West Branch 10 (above aqueduct) ------22-Oct-00 21-Oct-01 19-Oct-02 Tanjil River West Branch 11 (above aqueduct) ------22-Oct-00 - - The Morass 26-Nov-93 24-Nov-94 30-Nov-95 24-Nov-96 13-Nov-97 4/6/7-Nov-98 13-Oct/12-Nov-99 18-Nov-00 29-Oct/4-Nov-01 15-Nov-02 Thomson River 1 - - - - 19-Nov-97 16-Nov-98 28-Oct-99 - - - Thomson River 2 - - - - 19-Nov-97 16/21-Nov-98 28-Oct-99 - - - Thomson River 3 - - - - - 21-Nov-98 - - - - Thomson River 4 - - - - - 19-Nov-98 - - - - Thomson River 5 - - - - - 19-Nov-98 - - - - Tyers River 1 (right branch) - - - 5-Dec-96 22-Nov-97 26-Nov-98 1-Nov-99 17-Nov-00 10-Oct-01 21-Oct-02 Tyers River 1 frost hollow 1-Dec-93 7-Dec-94 3-Dec-95 26-Nov-96 9-Nov-97 5-Nov-98 30-Oct-99 7-Nov-00 15-Oct-01 14-Nov-02 Tyers River 2 (left branch) - - - 10-Dec-96 ------Tyers River 2 (right branch) - - - 10-Dec-96 ------Tyers River 2 frost hollow 1-Dec-93 7-Dec-94 3-Dec-95 26-Nov-96 8-Nov-97 5-Nov-98 30-Oct-99 7-Nov-00 15-Oct-01 14-Nov-02 Tyers River 3 - - - - - 2-Nov-98 - - - - Tyers River 4 - - - - - 2-Nov-98 - - - - Tyers River 5 - - - - - 2-Nov-98 - - - - Tyers River East Branch (tributary 1 ) - - - - - 10-Nov-98 1-Nov-99 17-Nov-00 10-Oct-01 21-Oct-02 Tyers River East Branch (tributary 2 ) - - - - - 10-Nov-98 1-Nov-99 - 9/29-Oct-01 21-Oct-02 Tyers River East Branch (tributary 3 ) - - - - - 10-Nov-98 1-Nov-99 17-Nov-00 10-Oct-01 21-Oct-02

321 Survey Transect 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 Tyers River East Branch (tributary) - - - 5-Dec-96 22-Nov-97 26-Nov-98 1-Nov-99 17-Nov-00 10/29-Oct-01 21-Oct-02 Tyers River West Branch (tributary A) - - - - - 10-Nov-98 - - - - Tyers River West Branch (tributary B) - - - - - 10-Nov-98 - - - - Tyers River West Branch (tributary C) - - - - - 7-Nov-98 - - - - Village Flat/Dam Valley/Big Hill 17-Nov-93 19-Nov-94 10-Nov-95 12-Nov-96 30-Oct-97 14-Oct-98 13-Oct-99 30-Oct-00 - 8-Nov-02 Village Flat 2 - - 28-Nov-95 12-Nov-96 12-Nov-97 16-Oct-98 13-Oct-99 7-Nov-00 - 6-Nov-02 Whitelaw 1 7-Dec-93 9-Dec-94 8-Dec-95 ------Whitelaw 10 9-Dec-93 ------Whitelaw 11 7-Dec-93 ------Whitelaw 12 7-Dec-93 9-Dec-94 7-Dec-95 ------Whitelaw 13 8-Dec-93 9-Dec-94 ------Whitelaw 2 8-Dec-93 9-Dec-94 8-Dec-95 ------Whitelaw 3 8-Dec-93 10-Dec-94 ------Whitelaw 4 8-Dec-93 10-Dec-94 8-Dec-95 ------Whitelaw 5 8-Dec-93 10-Dec-94 8-Dec-95 ------Whitelaw 6 9-Dec-93 10-Dec-94 8-Dec-95 ------Whitelaw 6a - 10-Dec-94 ------Whitelaw 7 9-Dec-93 10-Dec-94 9-Dec-95 ------Whitelaw 8 9-Dec-93 11-Dec-94 9-Dec-95 ------Whitelaw 9 9-Dec-93 ------Whitelaw Creek - - - - 9-Nov-97 26-Nov-98 - - - - Whitelaw Creek (frost hollow) 9-Dec-93 11-Dec-94 9-Dec-95 ------Whitelaw Creek (lower section) - - - - 9-Nov-97 - - - - - Whitelaw Creek (right branch A) - - - - 5-Nov-97 3-Nov-98 - - - - Whitelaw Creek (right branch A-lower section) - - - - 6-Nov-97 - - - - Whitelaw Creek (right branch B) - - - - 5-Nov-97 3-Nov-98 17-Oct-99 - - - Whitelaw Creek (right branch C) ------17-Oct-99 - - - Whitelaw Creek 1 - - - - - 26-Nov-98 11-Nov-99 - - - Whitelaw Creek 2 ------11-Nov-99 - - - Whitelaw Creek Middle Branch - - - - 10-Nov-97 19-Nov-98 - - - -

322 Survey Transect 1993 1994 1995 1996 1997 1998 1999 2000 2001 2002 Whitelaw Ruins 8-Dec-93 10-Dec-94 9-Dec-95 ------Wombat Flat 1-Dec-93 - - - 8-Nov-97 5-Nov-98 3-Nov-99 - - 14-Nov-02 Wombat Flat A - - - - 8-Nov-97 - - - - - X-Sandys Flat - - - 14-Nov-96 30-Oct-97 25-oct-98 16-Oct-99 1-Nov-00 - -

323

Appendix 4.1: Results of fitting vectors of maximum correlation for environmental variables in NMDS ordinations.

Table 1. Results of fitting vectors of maximum correlation for environmental variables in the 2-dimensional NMDS ordination of all 10 x 10 m scale sites.

Environmental Variable Ordination n Correlation p Axis 1 Axis 2 Axis 1 Axis 2 code direction direction scaled scaled cosine cosine coordinate coordinate Linear distance to nearest sub-alpine wet heathland LDNSWH 124 0.90 < 0.0001 0.9619 0.2736 1.30 0.37 Linear distance to nearest cool temperate mixed forest LDNCTMF 124 0.84 < 0.0001 -0.9445 -0.3284 -1.19 -0.41 Altitude ALTITUDE 124 0.84 < 0.0001 -0.9517 -0.3071 -1.20 -0.39 % cover of vegetation 0 – 1.8 m %CV 0 – 1.8 124 0.75 < 0.0001 -0.9985 0.0552 -1.12 0.06 Linear distance to nearest sub-alpine woodland LDNSW 124 0.69 < 0.0001 0.8158 0.5783 0.85 0.60 % cover of log/woody debris %CL/WD 124 0.68 < 0.0001 0.9945 0.1045 1.01 0.11 % cover of litter %CLITTER 124 0.65 < 0.0001 0.9245 -0.3813 0.90 -0.37 Total vegetation cover 5.0 – 8.0 m height class TVC 5 – 8 124 0.63 < 0.0001 0.8545 -0.5195 0.81 -0.49 Morphology - flat MORPH-F 124 0.57 < 0.0001 -0.6422 0.7665 -0.55 0.65 Total vegetation cover 2.5 – 5.0 m height class TVC 2.5 – 5 124 0.54 < 0.0001 0.9072 -0.4208 0.74 -0.34 Total vegetation cover 27.0 – 41.0 m height class TVC 27 – 41 124 0.53 < 0.0001 0.8547 0.5192 0.68 0.41 Total vegetation cover 1.0 – 2.5 m height class TVC 1 – 2.5 124 0.52 < 0.0001 0.3982 -0.9173 0.31 -0.72 Maximum exposed rock height MERH 124 0.52 < 0.0001 0.1007 -0.9949 0.08 -0.78 Average exposed rock height AERH 124 0.51 < 0.0001 0.0205 -0.9998 0.02 -0.76 Total vegetation cover 0.3 – 1.0 m height class TVC 0.3 – 1 124 0.47 < 0.0001 0.2075 -0.9782 0.15 -0.69 Aspect ASPECT 122 0.46 < 0.0001 0.4794 0.8776 0.33 0.60 Total vegetation cover 8.0 – 15.0 m height class TVC 8 – 15 124 0.46 < 0.0001 0.9356 -0.3532 0.64 -0.24 Sub-catchment drainage direction SDD 124 0.46 < 0.0001 0.9976 0.0688 0.68 0.05 Relief RELIEF 124 0.43 < 0.0001 0.5698 -0.8218 0.37 -0.53 Exposed rock volume ERV 124 0.40 < 0.0001 -0.0205 -0.9998 -0.01 -0.60

325 Environmental Variable Ordination n Correlation p Axis 1 Axis 2 Axis 1 Axis 2 code direction direction scaled scaled cosine cosine coordinate coordinate % cover of standing water %CSW 124 0.40 < 0.0001 -0.5672 0.8236 -0.34 0.49 % cover of bare ground %CBG 124 0.40 < 0.0001 0.4537 0.8912 0.27 0.53 Total vegetation cover 0.1 – 0.3 m height class TVC 0.1 – 0.3 124 0.38 < 0.0001 0.1628 -0.9867 0.09 -0.56 % cover of exposed rock %CER 124 0.36 < 0.0001 -0.0321 -0.9995 -0.02 -0.54 Total vegetation cover 15.0 – 27.0 m height class TVC 15 – 27 124 0.35 < 0.0001 0.9894 -0.145 0.52 -0.08 Number of calling males in 10 x 10 m quadrat NO.MALES 124 0.33 = 0.001 0.9829 0.1842 0.49 0.09 % cover of surface seepage %CSEEP 124 0.30 = 0.002 0.7663 0.6425 0.34 0.29 Surface curvature - flat SC-F 124 0.28 = 0.008 -0.917 0.3988 -0.38 0.17 Number of exposed rocks NO.ER 124 0.27 = 0.01 0.3866 -0.9223 0.16 -0.37 Total vegetation cover 41.0 – 55.0 m height class TVC 41 – 55 124 0.27 = 0.009 0.9946 -0.1033 0.40 -0.04 Morphology - slope MORPH-S 124 0.26 = 0.017 0.9785 -0.2061 0.38 -0.08 Drainage density250 DD250 124 0.23 = 0.038 0.7979 0.6027 0.28 0.21 Surface curvature - terrace SC-T 124 0.22 = 0.037 0.6352 0.7723 0.21 0.26 Morphology – open depression MORPH-O 124 0.22 = 0.048 0.2115 -0.9774 0.07 -0.32

326 Table 2. Results of fitting vectors of maximum correlation for environmental variables into the 2-dimensional NMDS ordination of sub-alpine 10 x 10 m scale sites.

Environmental Variable Ordination n Correlation p Axis 1 Axis 2 Axis 1 Axis 2 direction direction scaled scaled code cosine cosine coordinate coordinate

Total vegetation cover 0.3 – 1.0 m height class TVC 0.3 – 1 85 0.75 < 0.0001 0.1886 -0.9821 0.21 -1.11 Total vegetation cover 1.0 – 2.5 m height class TVC 1 – 2.5 85 0.71 < 0.0001 0.5472 -0.837 0.58 -0.89 Linear distance to nearest montane riparian thicket LDNMRT 85 0.69 < 0.0001 -0.9321 -0.3622 -0.96 -0.37 Total vegetation cover 2.5 – 5.0 m height class TVC 2.5 – 5 85 0.68 < 0.0001 0.9647 0.2632 0.98 0.27 Relief RELIEF 85 0.68 < 0.0001 0.9436 -0.3311 0.96 -0.34 Linear distance to nearest sub-alpine wet heathland LDNSWH 85 0.65 < 0.0001 0.8955 0.445 0.87 0.43 Maximum exposed rock height MERH 85 0.64 < 0.0001 0.83 -0.5577 0.80 -0.54 % cover of vegetation 0 – 1.8 m %CV 0 – 1.8 85 0.62 < 0.0001 -0.9726 -0.2326 -0.91 -0.22 Morphology - flat MORPH-F 85 0.62 < 0.0001 -0.9258 0.3781 -0.86 0.35 Total vegetation cover 0.1 - 0.3 m height class TVC 0.1 – 0.3 85 0.61 < 0.0001 0.2606 -0.9655 0.24 -0.88 Total vegetation cover 8.0 – 15.0 m height class TVC 8 – 15 85 0.59 < 0.0001 0.9998 0.0198 0.88 0.02 Total vegetation cover 5.0 – 8.0 m height class TVC 5 – 8 85 0.59 < 0.0001 0.9976 0.0688 0.88 0.06 Average exposed rock height AERH 85 0.58 < 0.0001 0.7032 -0.711 0.61 -0.62 Number of calling males in 10 x 10 m quadrat NO.MALES 85 0.53 < 0.0001 0.9343 0.3564 0.75 0.28 % cover of litter %CLITTER 85 0.53 < 0.0001 0.976 0.2177 0.77 0.17 Exposed rock volume ERV 85 0.48 < 0.0001 0.7387 -0.674 0.53 -0.49 Number of exposed rocks NO.ER 85 0.48 < 0.0001 0.9992 -0.0412 0.71 -0.03 % cover of exposed rock %CER 85 0.47 < 0.0001 0.887 -0.4617 0.63 -0.33 % cover of standing water %CSW 85 0.44 < 0.0001 -0.9486 0.3166 -0.62 0.21 % cover of log/woody debris %CL/WD 85 0.43 < 0.0001 0.955 0.2965 0.62 0.19 Aspect ASPECT 83 0.41 < 0.0001 0.2174 0.9761 0.13 0.61 Morphology - slope MORPH-S 85 0.40 = 0.002 0.9767 0.2147 0.58 0.13 Surface curvature - flat SC-F 85 0.38 = 0.002 -0.9713 -0.238 -0.55 -0.13 Altitude ALTITUDE 85 0.37 = 0.005 -0.8508 -0.5256 -0.47 -0.29

327 Environmental Variable Ordination n Correlation p Axis 1 Axis 2 Axis 1 Axis 2 direction direction scaled scaled code cosine cosine coordinate coordinate

% cover of surface seepage %CSEEP 85 0.35 = 0.002 0.7384 0.6743 0.39 0.35 Linear distance to nearest sub-alpine woodland LDNSW 85 0.35 = 0.008 -0.9322 0.3619 -0.48 0.19 Total vegetation cover 15.0 – 27.0 m height class TVC 15 – 27 85 0.34 = 0.002 0.9832 0.1823 0.50 0.09 Surface curvature - terrace SC-T 85 0.34 = 0.001 0.5647 0.8253 0.29 0.42 Morphology – open depression MORPH-O 85 0.27 = 0.044 0.2443 -0.9697 0.10 -0.39

328 Table 3. Results of fitting vectors of maximum correlation for environmental variables in the 3-dimensional NMDS ordination of all 1 x 1 m scale sites.

Variable Ordination code n Correlation p Axis 1 Axis 2 Axis 3 Axis 1 Axis 2 Axis 3 direction direction direction scaled scaled scaled cosine cosine cosine coordinate coordinate coordinate

Altitude ALTITUDE 124 0.76 < 0.0001 -0.9194 -0.106 0.3789 -1.04 -0.12 0.43 Relief RELIEF 124 0.69 < 0.0001 0.4606 -0.3508 0.8153 0.48 -0.36 0.85 Total vegetation cover 0.3 – 1.0 m height class TVC 0.3 – 1 124 0.63 < 0.0001 0.2759 0.094 0.9566 0.26 0.09 0.91 Value score in soil horizon 3 H3 VALUE 23 0.61 = 0.03 0.217 -0.3782 0.8999 0.20 -0.35 0.82 Total vegetation cover 0.1 – 0.3 m height class TVC 0.1 – 0.3 124 0.61 < 0.0001 0.3162 0.1067 0.9427 0.29 0.10 0.86 Peat texture for soil horizon 2 H2 PEAT 94 0.55 < 0.0001 -0.9349 0.35 0.0587 -0.76 0.29 0.05 Conductivity of water sample CONDUCTIVITY 119 0.52 < 0.0001 0.9184 0.173 -0.3559 0.71 0.13 -0.28 Average exposed rock height AERH 124 0.51 < 0.0001 0.1317 -0.4335 0.8915 0.10 -0.33 0.68 Sphagnum moss texture for soil horizon 1 H1 SPHAGMOSS 124 0.51 < 0.0001 -0.7808 -0.1779 -0.5989 -0.60 -0.14 -0.46 Number of calling males within 10 x 10 m breeding sites NO.MALES 124 0.49 < 0.0001 0.4556 -0.71 0.537 0.34 -0.52 0.40 Number of logs/woody debris 5 – 10 cm diametre NO.L 5 – 10 124 0.48 < 0.0001 0.5899 -0.7345 -0.3354 0.43 -0.53 -0.24 Total vegetation cover 2.5 – 5.0 m height class TVC 2.5 – 5 124 0.48 < 0.0001 0.5594 -0.0669 -0.8262 0.40 -0.05 -0.59 Chroma score for soil horizon 1 H1 CHROMA 124 0.48 < 0.0001 -0.941 -0.0328 -0.3368 -0.67 -0.02 -0.24 Soil saturation depth SSD 118 0.47 < 0.0001 -0.3541 0.586 -0.7289 -0.25 0.42 -0.52 Loam texture for soil horizon 1 H1 LOAM 124 0.46 < 0.0001 0.9372 0.33 0.1129 0.64 0.23 0.08 Sand texture for soil horizon 2 H2 SAND 94 0.45 < 0.0001 0.57 -0.8208 -0.038 0.38 -0.55 -0.03 Maximum exposed rock height MERH 124 0.43 < 0.0001 0.2123 -0.4213 0.8817 0.14 -0.27 0.57 Rock depth in soil profile ROCK DEPTH 120 0.43 < 0.0001 -0.389 0.053 -0.9197 -0.25 0.03 -0.59 Organic texture for soil horizon 2 H2 ORGANIC 94 0.42 = 0.001 -0.8403 0.5087 -0.1871 -0.52 0.32 -0.12 Value score in soil horizon 1 H1 VALUE 124 0.41 < 0.0001 -0.7425 -0.1629 -0.6497 -0.46 -0.10 -0.40 Total vegetation cover 5.0 – 8.0 m height class TVC 5 – 8 124 0.41 < 0.0001 0.5538 -0.1391 -0.8209 0.34 -0.08 -0.50 Total vegetation cover 27.0 – 41.0 m height class TVC 27 – 41 124 0.40 < 0.0001 0.7484 0.5281 0.4014 0.45 0.32 0.24 Number of logs/woody debris > 40 cm diametre NO.L > 40 124 0.39 < 0.0001 0.9484 -0.2942 0.1183 0.56 -0.17 0.07 Number of logs/woody debris 10 – 40 cm diametre NO.L 10 – 40 124 0.38 = 0.001 0.727 -0.6729 -0.1369 0.42 -0.39 -0.08 Total vegetation cover 1.0 – 2.5 m height class TVC 1 – 2.5 124 0.38 = 0.001 0.8288 -0.1805 0.5297 0.47 -0.10 0.30

329 Variable Ordination code n Correlation p Axis 1 Axis 2 Axis 3 Axis 1 Axis 2 Axis 3 direction direction direction scaled scaled scaled cosine cosine cosine coordinate coordinate coordinate

Ground surface temperature/male calling site temperature GS-BC TD 122 0.38 < 0.0001 -0.5593 0.3088 -0.7693 -0.31 0.17 -0.43 differential Ground surface temperature GST 122 0.37 < 0.0001 -0.5772 0.2507 -0.7772 -0.32 0.14 -0.44 Sand texture for soil horizon 1 H1 SAND 124 0.36 < 0.0001 0.6059 -0.7868 -0.1178 0.32 -0.42 -0.06 Clay loam texture for soil horizon 2 H2 CLAYLOAM 94 0.35 = 0.001 0.4889 0.7137 0.5016 0.26 0.37 0.26 Depth of soil horizon 2 H2 DEPTH 92 0.35 = 0.013 -0.1009 0.1568 -0.9825 -0.05 0.08 -0.51 Loam texture for soil horizon 2 H2 LOAM 94 0.34 = 0.007 0.7459 0.566 -0.3511 0.38 0.29 -0.18 PH of water sample PH 120 0.34 = 0.002 0.9865 -0.093 0.1348 0.50 -0.05 0.07 Hue score from soil horizon 1 H1 HUE 124 0.34 = 0.001 0.6035 -0.736 -0.3067 0.30 -0.37 -0.15 Peat texture for soil horizon 1 H1 PEAT 124 0.33 = 0.002 -0.7738 0.349 0.5286 -0.39 0.17 0.26 Organic texture for soil horizon 1 H1 ORGANIC 124 0.32 = 0.006 -0.881 0.367 -0.2987 -0.43 0.18 -0.14 Total vegetation cover 15.0 – 27.0 m height class TVC 15 – 27 124 0.32 = 0.004 0.7856 0.6116 -0.0935 0.38 0.29 -0.04 Site temperature-weather station temperature differential TD-TDW D 122 0.29 = 0.017 -0.9088 0.3903 0.2524 -0.40 0.17 -0.06

330 Table 4. Results of fitting vectors of maximum correlation for environmental variables in the 3-dimensional NMDS ordination of sub-alpine 1 x 1 m scale sites.

Variable Ordination code n Correlation p Axis 1 Axis 2 Axis 3 Axis 1 Axis 2 Axis 3 direction direction direction scaled scaled scaled cosine cosine cosine coordinate coordinate coordinate

Total vegetation cover 0.3 – 1.0 m height class TVC 0.3 – 1 85 0.76 < 0.0001 0.49 -0.3097 0.8149 0.49 -0.31 0.81 Relief RELIEF 85 0.75 < 0.0001 0.7535 0.1574 0.6383 0.75 0.16 0.64 Total vegetation cover 0.1 – 0.3 m height class TVC 0.1 –0.3 85 0.71 < 0.0001 0.6305 -0.0441 0.775 0.63 -0.04 0.78 Value score from soil horizon 3 H3 VALUE 20 0.63 = 0.044 0.052 0.1045 0.9932 0.05 0.10 0.99 Maximum exposed rock height MERH 85 0.62 < 0.0001 0.6207 0.0351 0.7833 0.62 0.04 0.78 Average exposed rock height AERH 85 0.62 < 0.0001 0.5949 0.0911 0.7986 0.59 0.09 0.80 Sphagnum moss texture for soil horizon 1 H1 SPHAGMOSS 85 0.57 < 0.0001 -0.6643 0.5649 -0.4895 -0.66 0.56 -0.49 Number of calling males within 10 x 10 m breeding sites NO.MALES 85 0.57 < 0.0001 0.8343 0.2511 0.4908 0.83 0.25 0.49 Total vegetation cover 1.0 – 2.5 m height class TVC 1 – 2.5 85 0.56 < 0.0001 0.8249 -0.4544 0.3361 0.82 -0.45 0.34 Altitude ALTITUDE 85 0.54 < 0.0001 -0.2956 -0.4071 0.8642 -0.30 -0.41 0.86 Peat texture for soil horizon 2 H2 PEAT 68 0.54 < 0.0001 -0.9679 -0.019 0.2506 -0.97 -0.02 0.25 Chroma score from soil horizon 1 H1 CHROMA 85 0.54 < 0.0001 -0.717 0.5078 -0.4776 -0.72 0.51 -0.48 Total vegetation cover 2.5 – 5.0 m height class TVC 2.5 – 5 85 0.52 < 0.0001 0.6393 0.1142 -0.7604 0.64 0.11 -0.76 Value score from soil horizon 1 H1 VALUE 85 0.50 < 0.0001 -0.5941 0.5046 -0.6265 -0.59 0.50 -0.63 Rock depth in soil profile ROCK DEPTH 82 0.48 < 0.0001 -0.5803 0.3257 -0.7464 -0.58 0.33 -0.75 Organic texture for soil horizon 2 H2 ORGANIC 68 0.47 = 0.002 -0.9506 -0.1883 0.2468 -0.95 -0.19 0.25 Sand texture for soil horizon 2 H2 SAND 68 0.47 < 0.0001 0.9993 0.0341 -0.0116 1.00 0.03 -0.01 Peat texture for soil horizon 1 H1 PEAT 85 0.45 < 0.0001 0.169 -0.8735 0.4565 0.17 -0.87 0.46 Number of logs/woody debris 5 – 10 cm diametre NO.L 5 – 10 85 0.45 < 0.0001 0.6574 0.3245 -0.6801 0.66 0.32 -0.68 Depth of soil horizon 2 H2 DEPTH 66 0.43 = 0.007 -0.6676 0.2977 -0.6824 -0.67 0.30 -0.68 Total vegetation cover 5.0 – 8.0 m height class TVC 5 – 8 85 0.42 < 0.0001 0.4092 0.1983 -0.8906 0.41 0.20 -0.89 Soil saturation depth SSD 84 0.41 = 0.001 -0.4418 -0.2125 -0.8716 -0.44 -0.21 -0.87 Ground surface temperature GST 84 0.41 = 0.002 -0.7159 0.1673 -0.6779 -0.72 0.17 -0.68 Ground surface temperature/male calling site temperature GS-BC TD 84 0.39 = 0.003 -0.7219 -0.1997 -0.6625 -0.72 -0.20 -0.66 differential

331 Variable Ordination code n Correlation p Axis 1 Axis 2 Axis 3 Axis 1 Axis 2 Axis 3 direction direction direction scaled scaled scaled cosine cosine cosine coordinate coordinate coordinate

PH of water sample PH 84 0.38 = 0.006 0.8948 0.0793 -0.4394 0.89 0.08 -0.44 Male calling site temperature BC-TEMP 84 0.37 = 0.009 -0.5097 0.7497 -0.422 -0.51 0.75 -0.42 Sand texture for soil horizon 1 H1 SAND 85 0.36 = 0.008 0.8449 0.4653 0.264 0.84 0.47 0.26 Chroma score for soil horizon 2 H2 CHROMA 67 0.36 = 0.028 -0.6707 -0.0555 -0.7397 -0.67 -0.06 -0.74 Hue score for soil horizon 1 H1 HUE 85 0.34 = 0.016 0.7936 0.5278 -0.3027 0.79 0.53 -0.30 Organic texture for soil horizon 1 H1 ORGANIC 85 0.32 0.03 -0.8563 -0.4661 0.2223 -0.86 -0.47 0.22 Number of logs/woody debris 10 – 40 cm diametre NO.L 10 – 40 85 0.31 0.026 0.5048 0.8178 -0.2764 0.50 0.82 -0.28 Clay loam texture for soil horizon 1 H1 CLAYLOAM 85 0.30 0.013 0.2763 0.2131 -0.9372 0.28 0.21 -0.94

332 Table 5. Results of fitting vectors of maximum correlation for environmental variables in the 2-dimensional NMDS ordination of all sites using biophysical data collected at 10 x 10 m (10), 1 x 1 m (1) and sub-catchment (S) scales. Variables relating to both 10 x 10 and 1 x 1 m scales have no identifier (e.g., Altitude).

Variable Ordination code n Correlation p Axis 1 Axis 2 Axis 1 Axis 2 direction direction scaled scaled cosine cosine coordinate coordinate Linear distance to nearest sub-alpine wet heathland(s) LDNSWHS 124 0.89 < 0.0001 0.956 0.2935 1.70 0.52 Linear distance to nearest cool temperate mixed forest(s) LDNCTMFS 124 0.88 < 0.0001 -0.9279 -0.3728 -1.63 -0.65 Altitude ALTITUDE 124 0.83 < 0.0001 -0.9351 -0.3544 -1.55 -0.59 % cover of vegetation 0 – 1.8 m (10) %CV 0 – 1.810 124 0.75 < 0.0001 -0.9975 -0.0705 -1.50 -0.11 Linear distance to nearest sub-alpine woodland(S) LDNSWS 124 0.74 < 0.0001 0.891 0.4539 1.32 0.67 % cover of log/woody debris (10) %CL/WD10 124 0.71 < 0.0001 0.9356 0.353 1.32 0.50 Relief(10) RELIEF10 124 0.69 < 0.0001 0.4763 -0.8793 0.65 -1.21 Peat texture for soil horizon 2(1) H2 PEAT1 94 0.65 < 0.0001 -0.8325 0.554 -1.08 0.72 Total vegetation cover 1 – 2.5 m height class(10) TVC 1 – 2.510 124 0.63 < 0.0001 0.3743 -0.9273 0.48 -1.18 % cover of litter(10) %CLITTER10 124 0.62 < 0.0001 0.8977 0.4406 1.12 0.55 Organic texture for soil horizon 2(1) H2 ORGANIC1 94 0.61 < 0.0001 -0.6694 0.7429 -0.82 0.91 Peat texture for soil horizon 1(1) H1 PEAT1 124 0.61 < 0.0001 -0.5902 -0.8072 -0.72 -0.99 Average exposed rock height(1) AERH1 124 0.57 < 0.0001 0.0227 -0.9997 0.03 -1.14 Sphagnum moss texture for soil horizon 1(1) H1 SPHAGMOSS1 124 0.57 < 0.0001 -0.6154 0.7882 -0.70 0.90 Total vegetation cover 27 – 41 m height class(10) TVC 27 - 4110 124 0.56 < 0.0001 0.9998 0.0177 1.12 0.02 Chroma score in soil horizon 1(1) H1 CHROMA1 124 0.56 < 0.0001 -0.5775 0.8164 -0.64 0.91 Morphology – slope(10) MORPH-S10 124 0.56 < 0.0001 0.4645 -0.8856 0.52 -0.98 Morphology – flat(10) MORPH-F10 124 0.55 < 0.0001 -0.7945 0.6073 -0.87 0.67 Loam texture for soil horizon 1(1) H1 LOAM1 124 0.55 < 0.0001 0.8773 0.4799 0.96 0.53 Conductivity of water sample(1) CONDUCTIVITY1 119 0.54 < 0.0001 0.9934 0.115 1.07 0.12 Total vegetation cover 8 – 15 m height class(10) TCV 8 - 1510 124 0.54 < 0.0001 0.7747 0.6323 0.83 0.68 Total vegetation cover 0.3 – 1 m height class(10) TCV 0.3 - 110 124 0.52 < 0.0001 0.2289 -0.9734 0.24 -1.02 Maximum exposed rock height(1) MERH1 124 0.52 < 0.0001 0.0538 -0.9986 0.06 -1.03

333 Variable Ordination code n Correlation p Axis 1 Axis 2 Axis 1 Axis 2 direction direction scaled scaled cosine cosine coordinate coordinate Value score in soil horizon 1(1) H1 VALUE1 124 0.51 < 0.0001 -0.4798 0.8774 -0.49 0.90 Chroma score in soil horizon 3(1) H3 CHROMA1 23 0.51 = 0.051 0.089 0.996 0.09 1.01 Peat texture for soil horizon 3(1) H3 PEAT1 23 0.50 = 0.052 -0.796 0.6054 -0.79 0.60 Relief(1) RELIEF1 124 0.50 < 0.0001 0.5826 -0.8127 0.58 -0.81 Depth of soil horizon 2 H2 DEPTH1 92 0.49 < 0.0001 0.084 0.9965 0.08 0.98 Total vegetation cover 5 – 8 m height class(10) TVC 5 - 810 124 0.49 < 0.0001 0.9852 0.1715 0.96 0.17 Total vegetation cover 15 – 27 m height class(10) TVC 15 - 2710 124 0.47 < 0.0001 0.9765 0.2154 0.91 0.20 Organic texture for soil horizon 1(1) H1 ORGANIC1 124 0.46 < 0.0001 -0.8722 -0.4892 -0.81 -0.45 Sand texture for soil horizon 2(1) H2 SAND1 94 0.46 < 0.0001 0.4689 -0.8833 0.43 -0.81 Total vegetation cover 41 – 55 m height class(10) TVC 41 - 5510 124 0.45 < 0.0001 0.8198 0.5727 0.74 0.52 Exposed rock volume(10) ERV10 124 0.45 < 0.0001 -0.115 -0.9934 -0.10 -0.89 Total vegetation cover 2.5 – 5 m height class(10) TVC 2.5 - 510 124 0.44 < 0.0001 0.9505 -0.3108 0.84 -0.28 Aspect(10) ASPECT10 122 0.44 < 0.0001 0.84 0.5426 0.74 0.48 Aspect(s) ASPECTS 124 0.44 < 0.0001 0.9894 0.1452 0.87 0.13 Rock depth in soil profile(1) ROCK DEPTH1 120 0.44 < 0.0001 -0.5376 0.8432 -0.47 0.74 Surface curvature – flat(10) SC-F10 124 0.44 < 0.0001 -0.5596 0.8287 -0.49 0.73 Soil saturation depth (1) SSD1 118 0.43 < 0.0001 -0.6562 0.7546 -0.56 0.65 PH of water sample(1) PH1 120 0.42 < 0.0001 0.7934 0.6087 0.67 0.51 Loam texture for soil horizon 2(1) H2 LOAM1 94 0.42 < 0.0001 0.7742 0.633 0.65 0.53 %cover of bare ground(10) %CBG10 124 0.42 < 0.0001 0.9191 -0.3939 0.77 -0.33 Number of logs/woody debris > 40 cm diametre(1) NO.L >401 124 0.42 < 0.0001 0.9804 0.1972 0.82 0.16 % cover of exposed rock(10) %CER10 124 0.40 < 0.0001 -0.0803 -0.9968 -0.06 -0.79 Number of calling males (10) NO.MALES10 124 0.40 < 0.0001 0.5772 -0.8166 0.46 -0.65 Ground surface temperature(1) GST1 122 0.39 < 0.0001 -0.6021 0.7984 -0.47 0.63 Number of logs/woody debris 5 - 10 cm diametre(1) NO.L 5 - 101 124 0.38 < 0.0001 0.9127 0.4087 0.69 0.31 Ground surface temperature/ calling male site temperature GS-BC TD1 122 0.37 = 0.001 -0.6618 0.7497 -0.49 0.55 differential(1) Total vegetation cover 0.1 – 0.3 m height class(10) TVC 0.1 – 0.310 124 0.36 = 0.001 0.1995 -0.9799 0.15 -0.71

334 Variable Ordination code n Correlation p Axis 1 Axis 2 Axis 1 Axis 2 direction direction scaled scaled cosine cosine coordinate coordinate Clay loam texture for soil horizon 1(1) H1 CLAY LOAM1 124 0.35 = 0.001 0.9811 0.1934 0.69 0.14 Clay loam texture for soil horizon 2(1) H2 CLAY LOAM1 94 0.35 = 0.003 0.8783 -0.4781 0.62 -0.34 Number of logs/woody debris 10 - 40 cm diametre(1) NO.L 10 - 401 124 0.35 < 0.0001 0.9 0.4359 0.63 0.30 Maximum exposed rock height(10) MERH10 124 0.34 = 0.001 0.1496 -0.9888 0.10 -0.67 Aspect(1) ASPECT1 114 0.34 = 0.003 0.7718 0.6358 0.52 0.43 Site temperature-weather station temperature differential TD-TDW D 122 0.33 = 0.001 -0.9999 -0.0123 -0.65 -0.01 Morphology – open depression(10) MORPH-O10 124 0.32 = 0.002 -0.0057 1 0.00 0.64 % cover of standing water(10) %CSW10 124 0.31 = 0.002 -0.8895 0.4569 -0.55 0.29 Value score in soil horizon 2(1) H2 VALUE1 93 0.31 = 0.014 0.9994 0.0344 0.61 0.02 Drainage density250(s) DD250S 124 0.30 = 0.004 0.5677 0.8232 0.34 0.49 Sand texture for soil horizon 1(1) H1 SAND1 124 0.29 = 0.003 0.986 0.1665 0.58 0.10 Surface curvature – terrace(10) SC-T10 124 0.29 = 0.005 0.9192 -0.3939 0.53 -0.23 Chroma score in soil horizon 2(1) H2 CHROMA1 93 0.28 = 0.027 0.3687 0.9296 0.20 0.51 Average exposed rock height(10) AERH10 124 0.27 = 0.011 -0.078 -0.997 -0.04 -0.54 Male calling site temperature(1) BC TEMP1 122 0.27 = 0.009 -0.5183 0.8552 -0.28 0.46 % cover of surface seepage (10) %CSEEP10 124 0.27 = 0.012 0.8071 -0.5905 0.43 -0.32 % cover of flowing water(10) %CFW10 124 0.27 = 0.01 -0.2042 0.9789 -0.11 0.52 Hue score in soil horizon 1(1) H1 HUE1 124 0.26 = 0.014 0.9895 -0.1448 0.51 -0.07 Surface curvature – concave(10) SC-CC10 124 0.24 = 0.036 0.1006 -0.9949 0.05 -0.47

335 Table 6. Results of fitting vectors of maximum correlation for environmental variables in the 3-dimensional NMDS ordination of sub-alpine sites using biophysical data collected at 10 x 10 m (10), 1 x 1 m (1) and sub-catchment (s) scales. C = correlation coefficient. Variables relating to both 10 x 10 m and 1 x 1 m scales have no identifier (e.g., Dry bulb temperature).

Variable Ordination code n C p Axis 1 Axis 2 Axis 3 Axis 1 Axis 2 Axis 3 direction direction direction scaled scaled scaled cosine cosine cosine coordinate coordinate coordinate Relief(10) RELIEF10 85 0.83 < 0.0001 0.9941 0.1069 0.0159 1.24 0.13 0.02 Dry bulb temperature TD 84 0.83 < 0.0001 -0.0489 -0.5155 -0.8555 -0.06 -0.64 -1.07 Relative humidity RH 84 0.81 < 0.0001 0.1124 0.4262 0.8976 0.14 0.52 1.09 Dry bulb temperature/male calling site differential(1) TD-BC TD1 84 0.81 < 0.0001 -0.0427 -0.4707 -0.8812 -0.05 -0.57 -1.07 Linear distance to nearest montane riparian thicket(s) LDNMRTS 85 0.74 < 0.0001 -0.5893 0.79 -0.1689 -0.66 0.88 -0.19 Total vegetation cover 1 – 2.5 m height class(10) TVC 1 – 2.510 85 0.73 < 0.0001 0.6755 0.2921 -0.6771 0.74 0.32 -0.75 Peat texture for soil horizon 2(1) H2 PEAT1 68 0.73 < 0.0001 -0.8565 0.02 -0.5157 -0.94 0.02 -0.57 Peat texture for soil horizon 1(1) H1 PEAT1 85 0.73 < 0.0001 0.2521 0.8939 -0.3706 0.27 0.97 -0.40 Sand texture for soil horizon 2(1) H2 SAND1 68 0.72 < 0.0001 0.9598 0.119 0.2544 1.04 0.13 0.27 Value score in soil horizon 1(1) H1 VALUE1 85 0.71 < 0.0001 -0.5888 -0.5461 0.5959 -0.63 -0.58 0.64 Sphagnum moss texture for soil horizon 1(1) H1 SPHAGMOSS1 85 0.70 < 0.0001 -0.7169 -0.4654 0.5191 -0.75 -0.49 0.54 Chroma score in soil horizon 1(1) H1 CHROMA1 85 0.70 < 0.0001 -0.7239 -0.5452 0.4228 -0.76 -0.57 0.44 Organic texture for soil horizon 2(1) H2 ORGANIC1 68 0.67 < 0.0001 -0.9183 0.2238 -0.3266 -0.93 0.23 -0.33 Total vegetation cover 2.5 – 5 m height class(10) TVC 2.5 - 510 85 0.67 < 0.0001 0.7952 -0.6048 0.0431 0.80 -0.61 0.04 Relief1 RELIEF1 85 0.66 < 0.0001 0.9329 0.3344 -0.1337 0.92 0.33 -0.13 Morphology – flat(10) MORPH-F10 85 0.66 < 0.0001 -0.9009 0.2747 0.336 -0.89 0.27 0.33 Number of calling males(10) NO.MALES10 85 0.65 < 0.0001 0.9408 -0.0692 -0.3318 0.92 -0.07 -0.32 Total vegetation cover 0.3 – 1 m height class(10) TVC 0.3 - 110 85 0.65 < 0.0001 0.4094 0.5422 -0.7338 0.40 0.53 -0.71 Maximum exposed rock height(10) MERH10 85 0.64 < 0.0001 0.9948 0.0477 0.0895 0.96 0.05 0.09 Ground surface temperature/male calling site temperature GS-BC TD1 84 0.63 < 0.0001 -0.5154 -0.4455 -0.7321 -0.49 -0.42 -0.70 differential(1) Average exposed rock height(1) AERH1 85 0.62 < 0.0001 0.6911 0.6964 -0.1935 0.64 0.65 -0.18

336 Variable Ordination code n C p Axis 1 Axis 2 Axis 3 Axis 1 Axis 2 Axis 3 direction direction direction scaled scaled scaled cosine cosine cosine coordinate coordinate coordinate Maximum exposed rock height(1) MERH1 85 0.62 < 0.0001 0.6646 0.7094 -0.2347 0.61 0.66 -0.22 Linear distance to nearest sub-alpine wet heathland(s) LDNSWHS 85 0.61 < 0.0001 0.5587 -0.8212 -0.1163 0.51 -0.76 -0.11 Total vegetation cover 8 – 15 m height class(10) TVC 8 - 1510 85 0.61 < 0.0001 0.6988 -0.6068 0.3787 0.64 -0.56 0.35 Organic texture for soil horizon 1(1) H1 ORGANIC1 85 0.60 < 0.0001 -0.3023 0.9375 -0.1723 -0.27 0.84 -0.15 % cover of vegetation 0 – 1.8 m (10) %CV 0 – 1.810 85 0.59 < 0.0001 -0.895 0.4432 -0.05 -0.80 0.40 -0.04 Ground surface temperature(1) GST1 84 0.58 < 0.0001 -0.6864 -0.5524 -0.4729 -0.60 -0.48 -0.41 Morphology – slope(10) MORPH-S10 85 0.58 < 0.0001 0.851 -0.3809 0.3616 0.74 -0.33 0.32 % cover of exposed rock(10) %CER10 85 0.57 < 0.0001 0.9933 0.1153 -0.0066 0.85 0.10 -0.01 Sand texture for soil horizon 1(1) H1 SAND1 85 0.57 < 0.0001 0.358 -0.9001 0.2484 0.30 -0.76 0.21 Rock depth in soil profile(1) ROCK DEPTH1 82 0.55 < 0.0001 -0.9118 -0.3397 -0.2305 -0.75 -0.28 -0.19 Total vegetation cover 5 – 8 m height class(10) TVC 5 - 810 85 0.54 < 0.0001 0.8339 -0.5236 -0.1742 0.68 -0.42 -0.14 Soil saturation depth(1) SSD1 84 0.54 < 0.0001 -0.922 0.096 -0.3752 -0.74 0.08 -0.30 Exposed rock volume(10) ERV10 85 0.53 < 0.0001 0.9359 0.3254 0.1347 0.74 0.26 0.11 Surface curvature – terrace(10) SC-T10 85 0.53 < 0.0001 0.4614 -0.34 0.8195 0.36 -0.27 0.65 Number of exposed rocks(10) NO.ER10 85 0.50 < 0.0001 0.8697 -0.4935 -0.006 0.66 -0.37 0.00 Chroma score in soil horizon 2(1) H2 CHROMA1 67 0.50 < 0.0001 -0.5959 -0.5979 0.5361 -0.45 -0.45 0.40 Aspect(10) ASPECT10 83 0.49 < 0.0001 0.2169 -0.5117 0.8313 0.16 -0.37 0.61 Average exposed rock height(10) AERH10 85 0.48 < 0.0001 0.9685 0.1373 0.2079 0.70 0.10 0.15 Surface curvature – flat(10) SC-F10 85 0.47 < 0.0001 -0.8815 0.3634 -0.3014 -0.62 0.25 -0.21 Male calling site temperature(1) BC TEMP1 84 0.46 < 0.0001 -0.5609 -0.2659 0.784 -0.39 -0.18 0.54 Aspect(1) ASPECT1 75 0.45 = 0.002 0.0362 -0.3484 0.9366 0.02 -0.23 0.63 Altitude ALTITUDE 85 0.43 = 0.001 -0.1877 0.9818 -0.0281 -0.12 0.64 -0.02 Relative humidity/weather station relative humidity differential RH-RHW D 84 0.42 = 0.001 -0.3216 -0.2467 -0.9142 -0.20 -0.16 -0.58 Total vegetation cover 0.1 – 0.3 m height class(10) TVC 0.1 – 0.310 85 0.42 = 0.003 0.5533 0.6068 -0.5707 0.35 0.38 -0.36 % cover of litter(10) %CLITTER10 85 0.41 = 0.002 0.8447 -0.5244 0.1068 0.53 -0.33 0.07 Total vegetation cover 15 – 27 m height class(10) TVC 15 - 2710 85 0.41 = 0.002 0.6538 -0.6149 0.4409 0.40 -0.38 0.27 Depth of soil horizon 2(1) H2 DEPTH1 66 0.41 = 0.009 -0.9746 -0.2238 -0.0021 -0.60 -0.14 0.00

337 Variable Ordination code n C p Axis 1 Axis 2 Axis 3 Axis 1 Axis 2 Axis 3 direction direction direction scaled scaled scaled cosine cosine cosine coordinate coordinate coordinate Dry bulb temperature/weather station temperature differential TD-TDW D 84 0.41 = 0.001 -0.4228 -0.215 -0.8804 -0.26 -0.13 -0.54 Huescore in soil horizon 1(1) H1 HUE1 85 0.40 = 0.004 0.6434 -0.4999 0.5798 0.38 -0.30 0.35 value score in soil horizon 2(1) H2 VALUE1 67 0.37 = 0.027 0.1687 -0.2225 0.9602 0.09 -0.12 0.53 PH from water sample(1) PH1 84 0.36 = 0.011 0.7946 -0.5149 -0.3218 0.43 -0.28 -0.18 Morphology – open depression(10) MORPH-O10 85 0.36 = 0.008 -0.1254 0.2197 -0.9675 -0.07 0.12 -0.52 % cover of standing water(10) %CSW10 85 0.36 = 0.006 -0.8816 0.4358 0.1815 -0.47 0.23 0.10 Clay loam texture for soil horizon 1(1) H1 CLAY LOAM1 85 0.36 = 0.003 0.232 -0.7766 -0.5858 0.12 -0.41 -0.31 Clay loam texture for soil horizon 2(1) H2 CLAY LOAM1 68 0.34 = 0.05 0.1983 -0.3895 0.8994 0.10 -0.20 0.46 % cover log/woody debris(10) %CL/WD10 85 0.32 = 0.039 0.7596 -0.6496 0.0326 0.36 -0.31 0.02 % cover of bare ground(10) %CBG10 85 0.32 = 0.028 0.0574 -0.5151 -0.8552 0.03 -0.24 -0.40 Linear distance to nearest sub-alpine woodland(s) LDNSWS 85 0.31 = 0.038 -0.8546 -0.3494 0.3842 -0.39 -0.16 0.18 Aspect(s) ASPECTS 85 0.31 = 0.046 0.6782 -0.1573 0.7179 0.31 -0.07 0.33 % cover of surface seepage(10) %CSEEP10 85 0.30 = 0.049 0.8121 -0.5432 0.2133 0.36 -0.24 0.10

338 Table 7. Results of fitting vectors of maximum correlation for environmental variables in the 2-dimensional NMDS ordination of montane sites using biophysical data collected at 10 x 10 m (10), 1 x 1 m (1) and sub-catchment (s) scales. C = correlation coefficient. Variables relating to both 10 x 10 m and 1 x 1 m scales have no identifier (e.g., Altitude).

Variable Ordination code n Correlation p Axis 1 Axis 2 Axis 1 Axis 2 direction direction scaled scaled cosine cosine coordinate coordinate Chroma score in soil horizon 2(1) H2CHROMA1 26 0.80 < 0.0001 0.78 -0.62 0.95 -0.75 Organic texture for soil horizon 1(1) H1 ORGANIC1 39 0.77 < 0.0001 -0.66 0.75 -0.77 0.87 total vegetation cover 0.1 – 0.3 m height class(10) TVC 0.1 – 0.310 39 0.75 < 0.0001 1.00 0.07 1.12 0.08 Chroma score in soil horizon 1(1) H1CHROMA1 39 0.75 < 0.0001 0.80 -0.60 0.89 -0.68 Total vegetation cover 0.3 – 1 m height class(10) TVC 0.3 - 110 39 0.71 < 0.0001 0.96 0.29 1.02 0.31 PH of water sample(1) PH1 36 0.71 < 0.0001 0.06 -1.00 0.06 -1.07 Morphology – open depression(10) MORPH-O10 39 0.71 < 0.0001 -0.48 -0.88 -0.51 -0.93 Morphology – slope(10) MORPH-S10 39 0.71 < 0.0001 0.48 0.88 0.51 0.93 Value score in soil horizon 1(1) H1 VALUE1 39 0.71 < 0.0001 0.86 -0.52 0.91 -0.55 Average exposed rock height(10) AERH10 39 0.70 < 0.0001 -0.48 -0.88 -0.51 -0.93 Value score in soil horizon 2(1) H2 VALUE1 26 0.70 < 0.0001 0.99 -0.13 1.04 -0.14 Total vegetation cover 5 – 8 m height class(10) TVC 5 - 810 39 0.69 < 0.0001 -0.98 0.20 -1.01 0.20 Total vegetation cover 1 – 2.5 m height class(10) TVC 1 – 2.510 39 0.69 < 0.0001 0.97 0.24 1.00 0.24 Clay loam texture for soil horizon 2(1) H2 CLAYLOAM1 26 0.67 < 0.0001 0.95 -0.32 0.96 -0.32 Male calling site temperature(1) BC TEMP1 38 0.62 = 0.001 -0.78 0.63 -0.72 0.59 % cover of litter(10) %CLITTER10 39 0.61 < 0.0001 -0.83 -0.55 -0.77 -0.51 Clay loam texture for soil horizon 1(1) H1 CLAYLOAM1 39 0.59 = 0.001 0.94 -0.35 0.83 -0.31 % cover of flowing water(10) %CFW10 39 0.58 < 0.0001 -0.65 -0.76 -0.56 -0.66 Maximum exposed rock height(10) MERH10 39 0.58 = 0.001 -0.64 -0.77 -0.56 -0.67 Exposed rock volume(10) ERV10 39 0.57 < 0.0001 -0.18 -0.98 -0.15 -0.85 Total vegetation cover 27 – 41 m height class(10) TVC 27 - 4110 39 0.57 = 0.001 0.95 0.32 0.81 0.28 Sand texture for soil horizon 2(1) H2 SAND1 26 0.57 = 0.013 -0.69 0.73 -0.58 0.62

339 Variable Ordination code n Correlation p Axis 1 Axis 2 Axis 1 Axis 2 direction direction scaled scaled cosine cosine coordinate coordinate % cover of bare ground(10) %CBG10 39 0.55 = 0.001 1.00 0.02 0.83 0.02 Conductivity of water sample(1) CONDUCTIVITY1 35 0.55 = 0.003 0.01 1.00 0.01 0.83 Surface curvature – flat(10) SC-F10 39 0.54 = 0.002 -0.56 -0.83 -0.45 -0.67 % cover of exposed rock(10) %CER10 39 0.52 = 0.001 -0.40 -0.92 -0.32 -0.72 Linear distance to nearest sub-alpine woodland(s) LDNSWS 39 0.52 = 0.003 -0.22 0.98 -0.17 0.76 Peat texture for soil horizon 1(1) H1 PEAT1 39 0.52 = 0.007 -0.87 0.50 -0.67 0.38 Loam texture for soil horizon 2(1) H2 LOAM1 26 0.51 = 0.031 0.37 -0.93 0.28 -0.70 Ground surface temperature(1) GST1 38 0.50 = 0.009 -0.51 0.86 -0.38 0.65 Linear distance to nearest cool temperate mixed forest(s) LDNCTMFS 39 0.48 = 0.005 -0.40 0.92 -0.29 0.67 % cover of vegetation 0 – 1.8(10) %CV 0 – 1.810 39 0.48 = 0.012 0.39 0.92 0.28 0.66 Total vegetation cover 41 – 55 m height class(10) TVC 41 - 5510 39 0.47 = 0.012 -0.47 -0.88 -0.33 -0.62 Total vegetation cover 15 – 27 m height class(10) TVC 15 - 2710 39 0.46 = 0.014 0.20 -0.98 0.14 -0.68 Relief(10) RELIEF10 39 0.46 = 0.012 0.84 -0.54 0.58 -0.37 Depth of soil horizon 1(1) H1 DEPTH1 39 0.45 = 0.014 0.89 -0.46 0.60 -0.31 Hue score in soil horizon 1(1) H1 HUE1 39 0.45 = 0.012 -0.62 0.78 -0.42 0.53 Surface curvature – concave(10) SC-CC10 39 0.43 = 0.03 0.30 0.95 0.20 0.61 Altitude ALTITUDE 39 0.43 = 0.018 0.39 -0.92 0.25 -0.59 Ground surface temperature/male calling site temperature GS-BC TD1 38 0.41 = 0.037 -0.15 0.99 -0.09 0.61 differential Number of logs/woody debris 10 - 40 cm diametre(1) NO.L 10 - 401 39 0.39 = 0.045 -0.21 -0.98 -0.12 -0.57

340 Appendix 4.2: Alphabetic list of plant taxa recorded in 1 x 1 m and 10 x 10 m quadrats, including generic groups where individuals were unidentifiable to species level during the study.

Nomenclature follows NRE (2001). * denotes introduced taxa.

Taxa Amalgamated Taxa Acacia dealbata Acaena novae-zelandiae. *Acetosella vulgaris Agrostis spp. A. parviflora, A. venusta, A.meionectes, *A. capillaris, A. sp. aff. hiemalis *Anthoxanthum odoratum Asperula gunnii Asperula spp. A. conferta, A. pusilla Asplenium bulbiferum ssp. gracillimum Asplenium flabellifolium Astelia alpina var. novae-hollandiae Asterolasia trymalioides Atherosperma moschatum Australina pusilla ssp. muelleri Baeckea gunniana Baeckea latifolia Blechnum fluviatile Blechnum nudum Blechnum penna-marina ssp. alpina Blechnum wattsii Brachyscome obovata Bracteantha subundulata Callistemon pityoides Callitriche spp. *C. stagnalis, C. sp. Caltha introloba Carex alsophila Carex appressa Carex breviculmis Carex gaudichaudiana Carex incomitata Carex spp. C. blakei, C. jackiana, C. hebes, C. canescens Carpha spp. C. alpina, C. nivicola Celmisia asteliifolia spp. agg. *Cerastium vulgare Chiloglottis valida Chionogentias spp. C. muelleriana ssp. muelleriana, C. bawbawensis Clematis aristata Coprosma granadensis Coprosma hirtella Coprosma moorei Coprosma nitida Coprosma perpusilla ssp. perpusilla Coprosma reptans Craspedia glauca spp. agg. Deyeuxia spp. D. brachyathera, D. monticola var. monticola, D. innominata, D.

341 Taxa Amalgamated Taxa carinata Dianella tasmanica Dicksonia antarctica Drosera arcturi Empodisma minus Epacris paludosa Epacris petrophila Epilobium spp. E. billardierianum, E. gunnianum Erigeron bellidioides Erigeron paludicola Erigeron tasmanicus Eucalyptus delegatensis ssp. delegatensis Eucalyptus glaucescens Eucalyptus nitens Eucalyptus pauciflora ssp. acerina Eucalyptus regnans Euchiton spp. E. involucratus s.l., E. collinus s.l. Euphrasia spp. E. collina, E. gibbsiae ssp. subglabrifolia Exocarpus nanus Fungi Galium gaudichaudii Gaultheria appressa Geranium potentilloides Gonocarpus micranthus Gonocarpus spp. G. humilis, G. montanus, G. tetragynus Grammitis billardierei Grevillea australis Hierochloe redolens Histiopteris incisa *Holcus lanatus Huperzia australiana Hydrocotyle spp. H. hirta, H. algida, H. laxiflora Hymenophyllum australe Hymenophyllum peltatum Hypericum japonicum *Hypochoeris radicata Hypolepis spp. H. rugosula, H. amaurorachis Isolepis crassiuscula Isolepis spp. I. aucklandica, I. habra, I. subtilissima *Juncus articulatus *Juncus bulbosus Juncus sandwithii Juncus alexandri ssp. alexandri Lagenophora stipitata Leptinella filicula Leptospermum grandifolium Leucopogon gelidus Leucopogon maccraei Libertia pulchella Liverwort (Marchantia) Liverwort spp. Lobelia pedunculata s.l. Luzula spp. L. meridionalis var. flaccida, L. meridionalis var. meridionalis, L. modesta Lycopodium fastigiatum

342 Taxa Amalgamated Taxa Lycopodium scariosum Mitrasacme serpyllifolia Monotoca oreophila Montia fontana ssp. fontana Moss & Lichen Myosotis australis Nothofagus cunninghamii Olearia algida Olearia phlogopappa var. flavescens Oreobolus distichus Oreobolus oxycarpus ssp. oxycarpus Oreobolus pumilio ssp. pumilio Oreomyrrhis spp. O. ciliata, O. eriopoda Orites lancifolia Oxalis magellanica Ozothamnus sp. 1 (formerly hookeri) Ozothamnus secundiflorus Pentachondra pumila Persoonia arborea Pimelea axiflora ssp. axiflora Pimelea alpina Pittosporum bicolor Plantago spp. P. alpestris, P. euryphylla *Poa annua Poa hiemata Poa spp.P. ensiformis, P. costiniana, P. fawcettiae, P. helmsii, P. labillardierei, P. sieberiana Poa tenera Podocarpus lawrencei Podolepis robusta Polyscias sambucifolia Polystichum proliferum Polytrichum alpinum Poranthera microphylla Prasophyllum spp. P. alpestre, P. alpinum Prostanthera cuneata Prostanthera lasianthos *Prunella vulgaris Pterostylis spp. P. alpina s.l., P. decurva Pultenaea muelleri var. muelleri Ranunculus gunnianus Ranunculus spp. R. collinus, R. muricatus, *R. repens Richea continentis Richea victoriana *Rubus fruticosus spp. agg. *Rumex obtusifolius spp. obtusifolius Rytidosperma nivicolum Rytidosperma nudiflorum *Salix cinerea Scaevola hookeri Schoenus calyptratus Senecio gunnii Senecio linearifolius Senecio pectinatus var. major Sphagnum cristatum

343 Taxa Amalgamated Taxa Sphagnum novozelandicum Stylidium graminifolium s.l. Tasmannia lanceolata Tasmannia vickeriana Thelymitra cyanea Trochocarpa clarkei Uncinia spp. U. flaccida, U. tenella Urtica incisa Viola hederacea s.l. Wittsteinia vacciniacea

344 Appendix 4.3: Matrix of taxa and structural attributes that best discriminated between random and breeding sites at 10 x 10 and 1 x 1 m scales from sub-alpine elevations, as identified using the step- wise, variant of ANOSIM procedure.

Original cover abundance values are presented for each pseudo-attribute; 1 = < 5%, 2 = 5 – 50%, 3 = > 50%, - = absent. * denotes exotic species.

10 x 10 m quadrat scale

Random Sites Breeding Sites

Carex gaudichaudiana 0 - 0.1m ----11--1-1--211------12-1------1------1------Carex gaudichaudiana 0.1 - 0.3m ------1------2-----12------1------Celmisia spp. 0 - 0.1m ---11--11-1-2--1111--111--11--11-1--1 --111-----111--1----1---11------1-1---1--1---11- Celmisia spp. 0.3 - 1.0m ------1------1--- Celmisia spp. 1 - 2.5m ------1------Chionogentias spp. 0 - 0.1m -111111------21--21---1111-1---1------11----1-1----1----1-----1------1-1-1------Chionogentias spp. 0.1 - 0.3m --11-1------1----1------Epacris petrophila 0 - 0.1m ------11------1------Epacris petrophila 0.1 - 0.3m ------2------1------Gonocarpus micranthus ssp. Mic.0 - 0.1m -----1----1--11-1------1------Olearia algida 0 - 0.1m ----1-1---1---21211---11------1------11---1---2------1-----1----1------Olearia algida 0.1 - 0.3m -1111----21---21-2121-1-1------11------11------11------1------1112--1------Olearia algida 0.3 - 1.0m --111------2--11---111111------31------2------1-1------Olearia algida 1 - 2.5m ------11------1------Pimelea axillaris 0 - 0.1m ------1------Pimelea axillaris 0.1 - 0.3m ------1------Poa hiemata 0 - 0.1m ----11-----2-111--12----2------21-1 ----2------1------Poa hiemata 0.1 - 0.3m ------1------3------Poa hiemata 0.3 - 1.0m ------3------Podolepis robusta 0 - 0.1m ---1-1------1------Acetosella vulgaris 0 - 0.1m ------1------Anthoxanthum odoratum 0 - 0.1m ------1------

345 Random Sites Breeding Sites Blechnum penna-marina 0 - 0.1m 1--1---1-1-11--11--1111--111----12121 111112111121111-111-111221111221111-1-1111111111 Blechnum penna-marina 0.1 - 0.3m ------1------1-1----211------11-- Blechnum penna-marina 0.3 - 1.0m ------1------1------Carex incomitata 0 - 0.1m ------1------Coprosma perpusilla ssp.perpus. 0 - 0.1m ---1-1------1------211------11--1--11-1--11---1----1------11--- Deyeuxia spp. 0 - 0.1m -1-1211-1--1-11-111------1-11-11-1 --111-----1----1-1-----1-1-1-1-1--1-----12-1111- Deyeuxia spp. 0.1 - 0.3m ------1---1-11-1------1------1------111--- Deyeuxia spp. 0.3 - 1.0m ------1------Deyeuxia spp. 1 - 2.5m ------1------Deyeuxia spp. 5 - 8.0m ------1------Dicksonia antarctica 0.1 - 0.3m ------1------1------Epilobium spp. 0 - 0.1m 1------1------1---1--1---1 2----1----211-----1---1-111-1111211---111-----11 Epilobium spp. 0.1 - 0.3m ------1 Erigeron bellidioides 0 - 0.1m ------11------1------Hierochloe redolens 0 - 0.1m ------1------1---1-----1- ----1-1-111------1-----11---1------1------1 Hierochloe redolens 0.1 - 0.3m -----1------1------1------1-11------11--1-1--1------1-1-----11-1 Hierochloe redolens 0.3 - 1.0m ------1------1------Hierochloe redolens 1 - 2.5m ------1------Hierochloe redolens 2.5 - 5.0m ------1------Holcus lanatus 0 - 0.1m ------1------Huperzia australiana 0 - 0.1m ------1------1--111------1-1------11-1111--1------Huperzia australiana 2.5 - 5.0m ------2------Juncus sandwithii 0 - 0.1m ------1- Leucopogon gelidus 0 - 0.1m ------1------1------Leucopogon gelidus 0.1 - 0.3m ------11 -11------1------1------1---1------1-1-- Leucopogon gelidus 0.3 - 1.0m ------1------1------1---- Leucopogon gelidus 1 - 2.5m ------2------1-1------Liverwort (Marchantia) 0 - 0.1m ------12------Liverwort spp. (other) 0 - 0.1m 1------12------11-2 1-1--22---2----1-11--111-1---1--2121-1--2-1111-- Liverwort spp. (other) 0.1 - 0.3m ------2------11-2 ---1-2----2------22------1-1-11-1-----12--- Liverwort spp. (other) 0.3 - 1.0m ------1-1 -----2----1------1-1--1-1-----12--- Liverwort spp. (other) 1 - 2.5m ------2------1-1------1---- Liverwort spp. (other) 2.5 - 5.0m ------1------Log 0 - 0.1m ------2------1------1--1 ------2------1----11-1--2-2---1-1------Log 0.1 - 0.3m ------1--1------1 ------2------2---211----2-----1-1------Log 0.3 - 1.0m ------1------1 ------2------211------1-1------Log 1 - 2.5m ------1------21------Log 2.5 - 5.0m ------1------Luzula spp. 0 - 0.1m ------1------1------1------111-----1---1-111------1----1-1-1--1- Luzula spp. 0.1 - 0.3m ------1------1---

346 Random Sites Breeding Sites Luzula spp. 2.5 - 5.0m ------1------Monotoca oreophila 0.3 - 1.0m -1------1------Monotoca oreophila 0.1 - 0.3m ------2------1------111------Moss & Lichen 0 - 0.1m ------121-2---12--111--1--22-1--2-2--11----1222221--- Moss & Lichen 0.1 - 0.3m ------1------1--- 11-1-2----2--11-1-----22---2111--1----2122111--- Moss & Lichen 0.3 - 1.0m ------1------1--- 11-1-2----1--11-1-----22--22111--1----1122111--- Moss & Lichen 1 - 2.5m ------1------1----2------1-1-----22--221-1--1----1-22111--- Moss & Lichen 2.5 - 5.0m ------1------1------Myosotis australis 0.1 - 0.3m ------1------Olearia phlogopappa 0 - 0.1m ------11------Olearia phlogopappa 0.1 - 0.3m ------1-1------1-1 11--11-2-----11--1---111--1-1------1-1-11--112-1 Olearia phlogopappa 0.3 - 1.0m ------1-1------1------11-1 11--1--2--1---1-1-----1-1-1-1-11-1--1-1-2-1111-1 Olearia phlogopappa 1 - 2.5m ------1------1-----1---- 11------1---1-1-----11----1------11--1--1 Olearia phlogopappa 2.5 - 5.0m ------1------1------Olearia phlogopappa 5 - 8.0m ------1------Podocarpus lawrencei 0 - 0.1m ------11------1-1------Podocarpus lawrencei 0.1 - 0.3m ------1--11------1------2-1------Podocarpus lawrencei 0.3 - 1.0m ------1--11------1------2-1------Podocarpus lawrencei 1 - 2.5m ------1--12------1------2-1------Polytrichum alpina 0 - 0.1m ----11--1-1-2-----1------1---3- ---1111-21-11------1----22111-----1------1--2 Polytrichum alpina 0.1 - 0.3m ------1------21------1--- Polytrichum alpina 0.3 - 1.0m ------2------1------1--- Polytrichum alpina 1 - 2.5m ------2------1------Pterostylis spp. 0 - 0.1m ------1------Pterostylis spp. 0.1 - 0.3m ------1------*Salix cinerea 1 - 2.5m ------1------1------*Salix cinerea 2.5 - 5.0m ------1------Schoenus calyptratus 0 - 0.1m ------1------Senecio gunnii 0 - 0.1m ------1------Senecio gunnii 0.1 - 0.3m ------1------Senecio linearifolius 0 - 0.1m ------1------1------1--1--1-- Senecio linearifolius 0.1 - 0.3m ------1------1------Senecio linearifolius 0.3 - 1.0m ------1------1------1--1--- Senecio linearifolius 1 - 2.5m ------1------1------Surface Seepage 0 - 0.1m ------1------21------1111-11-111111-1--2-111------1-1-2-12 Surface Seepage 0.1 - 0.3m ------2------Tea tree above ground roots/stump 0.1m ------2------1-11-- Tea tree a-g roots/stump 0.1 - 0.3m ------2------1------1--1-- Tea tree a-g roots/stump 0.3 - 1.0m ------1------Uncinia spp. 0 - 0.1m -----1------1-12-1-11------1-----1------11---1--

347 Random Sites Breeding Sites Uncinia spp. 0.1 - 0.3m ------1------1-----1------1------

348 1 x 1 m quadrat scale

Random Sites Breeding Sites

Baeckea gunniana 0 - 0.1m ---1------2------1------Baeckea gunniana 0.1 - 0.3m --21------1------2------Baeckea gunniana 0.3 - 1.0m --2------2------Baeckea utilis var. latifolia 0 - 0.1m -----1-1---1---1------1------1------Baeckea utilis var. latifolia 0.1 - 0.3m -----3-2----2------1---2-2---1------2------1------1- Baeckea utilis var. latifolia 0.3 - 1.0m ------2------1------3------1------2---2----1- Baeckea utilis var. latifolia 1 - 2.5m ------12--- -3----2-3------1------22--2---2------Baeckea utilis var. latifolia 2.5 - 5.0m ------2------Carpha spp. 0 - 0.1m -1--32------21---1------1------1------Carpha spp. 0.1 - 0.3m ------1------2------Carpha spp. 0.3 - 1.0m ------2------Empodisma minus 0 - 0.1m -3221-2---3--322-23--1-212-2--32------1-21-----1------1------111------Empodisma minus 0.1 - 0.3m --22--2-3------2-----2-2-2--2------2-3------1------312------2- Empodisma minus 0.3 - 1.0m --1------2-3------2------211------Hypericum japonicum 0 - 0.1m ------1------Lycopodium fastigiatum 0 - 0.1m -1--1------2---1------1------1------1---1----1------1---1------1- Lycopodium fastigiatum 0.1 - 0.3m ------11------1------Lycopodium fastigiatum 0.3 - 1.0m ------1------Olearia algida 0 - 0.1m ------2------11-1----21------Olearia algida 0.1 - 0.3m --1------1--1-2-1------1------Olearia algida 0.3 - 1.0m ------1------31------1------Olearia algida 1 - 2.5m ------1------Poa hiemata 0 - 0.1m ----1------2-1------1------1------Poa hiemata 0.1 - 0.3m ------3------1------Poa hiemata 0.3 - 1.0m ------1------13------Richea victoriana 0.3 - 1.0m ------3------Richea victoriana 1 - 2.5m ------3------Running water 0 - 0.1m ------2------1------Sphagnum cristatum 0 - 0.1m ---2-232--2-2-23-3-----2-----1212---3 ------1-----2---2------1------3- Sphagnum cristatum 0.1 - 0.3m ------3------1------2------Sphagnum cristatum 0.3 - 1.0m ------2------Tasmannia vickeriana 0 - 0.1m ------1------Tasmannia vickeriana 0.1 - 0.3m ------2----1-----2------2------1------Tasmannia vickeriana 0.3 - 1.0m ------2------1------3---- Tasmannia vickeriana 1 - 2.5m ------2------

349 Random Sites Breeding Sites Asperula spp. 0 - 0.1m ------1------1------1-- Bare Ground 0 - 0.1m 1-2-1--1-1------212------2-132------2-2--2-1221-1--12-2-22--23------2222 Bare Ground 0.1 - 0.3m ------3------2------2------Bare Ground 0.3 - 1.0m ------1------Callistemon pityoides 0 - 0.1m ------1------Callistemon pityoides 0.1 - 0.3m ---1--1------2------Callistemon pityoides 0.3 - 1.0m --1------2------2-1------Callistemon pityoides 1 - 2.5m ------1------2------1------Coprosma nitida 0 - 0.1m ------2-1------1------Coprosma nitida 0.1 - 0.3m ------1------1------11----1-1-1----1--2------1--2--2------Coprosma nitida 0.3 - 1.0m ------1------1------22------1111--3------Coprosma nitida 1 - 2.5m ------1------11------Exposed Rock 0 - 0.1m ----1------1--- 12-2-2---22--32-3-----2-----3--3-33-22-332--3--- Exposed Rock 0.1 - 0.3m ------32-2-2----2--32-32--2-2-----3--3-33-22-332--3--- Exposed Rock 0.3 - 1.0m ------32------1--32-3------3--3-33--2-322--3--- Exposed Rock 1 - 2.5m ------32-3------3----3------Fungi 0 - 0.1m ------1------Huperzia australiana 0 - 0.1m ------1---1------Isolepis spp. 0 - 0.1m ------1--1------1-- 1-1------11111-2121-1------22--- Isolepis spp. 0.1 - 0.3m ------2------Isolepis spp. 0.3 - 1.0m ------1--- Leptospermum grandifolium 0 - 0.1m ------1------1------Leptospermum grandifolium 0.3 - 1.0m ------2------2-1------3-22------1----3- Leptospermum grandifolium 1 - 2.5m ------3------3 -----32--233-----2-3-3-3---3-3332-----3--2--22-- Leptospermum grandifolium 2.5 - 5.0m ------23------3--3------23----3---3--3------3-1-23-- Leptospermum grandifolium 5 - 8.0m ------3------3--3------Libertia pulchella 0 - 0.1m ------3-21------11------1------1---1 Libertia pulchella 0.1 - 0.3m ------1------Libertia pulchella 0.3 - 1.0m ------2------1------Libertia pulchella 1 - 2.5m ------1------Liverwort spp. (other) 0 - 0.1m ------2------21-- 1-1--12---2------1----1-----2--311------1-2--- Liverwort spp. (other) 0.1 - 0.3m ------2------1------2------2------1------2--- Liverwort spp. (other) 0.3 - 1.0m ------1------1------2--- Luzula spp. 0 - 0.1m ------1-111------1------1--1- Luzula spp. 0.1 - 0.3m ------1------Moss & Lichen 0 - 0.1m ------1---1---12-2-1------1-1----2--2------122-2---2 Moss & Lichen 0.1 - 0.3m ------2--- -1------2---1------1----1-----122------Moss & Lichen 0.3 - 1.0m ------2--- -1------1------2----1-----1-2------Moss & Lichen 1 - 2.5m ------2----1------1------Podocarpus lawrencei 0.1 - 0.3m ------1------

350 Random Sites Breeding Sites Podocarpus lawrencei 0.3 - 1.0m ------1------1------Surface Seepage 0 - 0.1m ------2------1---1--21111----2-112------2-1-2--1 Surface Seepage 0.1 - 0.3m ------2-2------Wittsteinia vacciniacea 0 - 0.1m ------1-1------2---1-----1-2-----2-- -2--12-3-22---1--221-1-2--321-2-2--2--3-232----- Wittsteinia vacciniacea 0.1 - 0.3m ------11-1--2---1---3------23--3-3212----223-23-- Wittsteinia vacciniacea 0.3 - 1.0m ------3---1-----31---3--

351

Appendix 5.1: Distribution of seasonal and diel calling (calls/min, pulses/call/min) and growling (growls/min) data recorded at sub-alpine and montane breeding sites from 1994 - 1999.

Box plots show median, upper and lower quartiles, interquartile range and outside values. N.B. For the 1994 seasonal distribution, calling activity recorded at 15:00 h was used as a surrogate for 13:00 for the period 8 - 15 December (see methods). For diel calling activity, 0 = 24:00, 1 = 01:00, 7 = 07:00 h etc.

Seasonal Rates of Calling

1994

14 12 10 8 Montane 6 4 2

f 0 calls/min 2 4 6 Sub-alpine 8 10 12 14 Number o Number 13/11 20/11 27/11 4/12 11/12 18/12 Date (day/month)

1994 35 30 25 20 15 Montane 10 5 0 5 /call/min 10 15 Sub-alpine 20 25 30 35 13/11 20/11 27/11 4/12 11/12 18/12

Number of pulses of Number Date (day/month)

353 1995 16 14 12 10 8 6 Montane 4 2 0 calls/min 2 4 6 8 Sub-alpine 10 12 14 16

Number of 22/10 29/10 5/11 12/11 19/11 26/11 3/12 10/12 17/12 24/12

Date (day/month)

1995 25 20 15 10 Montane 5 0 5 /call/min 10 15 Sub-alpine 20 25 22/10 29/10 5/11 12/11 19/11 26/11 3/12 10/12 17/12 24/12 Number of pulses of Number Date (day/month)

354 1996 14 12 10 8 6 Montane 4 2 0 calls/min 2 4 6 Sub-alpine 8 10 12 14 Number of 20/10 27/10 3/11 10/11 17/11 24/11 1/12 8/12 15/12

Date (day/month)

28 1996 24 20 16

12 Montane 8 4 0 4 /call/min 8 Sub-alpine 12 16 20 24 28 20/10 27/10 3/11 10/11 17/11 24/11 1/12 8/12 15/12 Number of pulses of Number Date (day/month)

355 1997 12 10

8 Montane 6 4 2 0 calls/min 2 4 Sub-alpine 6 8 10 12

Number of 19/10 26/10 2/11 9/11 16/11 23/11 30/11 Date (day/month)

1997 30 25 20 15 Montane 10 5 0 5 /call/min 10 Sub-alpine 15 20 25 30 19/10 26/10 2/11 9/11 16/11 23/11 30/11

Number of pulses Date (day/month)

356 1998 8 7 6 5 4 Montane 3 2 1 00 calls/min 1 2 3 4 Sub-alpine 5 6 7 8

Number of 4/10 11/10 18/10 25/10 1/11 8/11 15/11 22/11 29/11 6/12

Date (day/month)

1998 90 80 70 60 50 40 Montane 30 20 10 0 10

/call/min 20 30 Sub-alpine 40 50 60 70 80 90 4/11 11/10 18/10 25/10 1/11 8/11 15/11 22/11 29/11 6/12 Number of pulses of Number Date (day/month)

357 1999 12 10 8 Montane 6 4 2 0 calls/min 2 04 Sub-alpine 6 8 10 12

Number of 5/9 12/9 19/9 26/9 3/10 10/10 17/10 24/10 31/10 7/11 14/11 21/11

Date (day/month)

1999 45 40 35 30 25 Montane 20 15 10 5 0 5

/call/min 10 15 Sub-alpine 20 25 30 35 40 45 5/9 12/9 19/9 26/9 3/10 10/10 17/10 24/10 31/10 7/11 14/11 21/11 Number of pulses Date (day/month)

358 Seasonal Rates of Growling

1994

20

15

10 Montane 5

0 growls/min 5 Sub-alpine

10

15

20 13/11 20/11 27/11 4/12 11/12 18/12 25/12 Number of Date (day/month)

1995

10 8 6 4 Montane 2 0 growls/min 2 4 Sub-alpine 6 8 10

Number of 22/10 29/10 5/11 12/11 19/11 26/11 3/12 10/12 17/12 24/12

Date (day/month)

359 1996

3.0 2.5 2.0 1.5 Montane 1.0 0.5 0.00

growls/min 0.5

1.00 Sub-alpine 1.5 2.0 2.5 3.0 20/10 27/10 3/11 10/11 17/11 24/11 1/12 8/12 15/12 Number of

Date (day/month)

1997

7 6 5 4

3 Montane 2 1 00 f

growls/min 1 Sub-alpine 2 3 4 5 6 7 19/10 26/10 2/11 9/11 16/11 23/11 30/11 Number o Date (day/month)

360 1998

1.2 1.0 0.8 0.6 0.4 Montane 0.2 0.00

growls/min 0.2 0.4 Sub-alpine 0.6 0.8 Number of 1.0 1.2 4/10 11/10/ 18/10 25/10 1/11 8/11 15/11 22/11 29/11 6/12

Date (day/month)

1999

5 4 3

2 Montane 1 0 growls/min 1 Sub-alpine 2 3 4 5

Number of 5/9 19/9 3/10 17/10 31/10 14/11 Date (day/month)

361 Diel Rates of Calling

1994 35 30 25 20 15 Montane 10 5 0 5 /call/min 10 Sub-alpine 15 20 25 30 35 0 1 2 3 4 5 6 7 8 9 1011121314151617181920212223

Number of pulses Time (hours)

1998 10 9 8 7 6 5 4 Montane 3 2 1 0 calls/min 1 2 3 4 5 Sub-alpine 6 7 8 9 10 Number of 0 1 2 3 4 5 6 7 8 9 1011121314151617181920212223 Time (hours)

362 60 1998 50 40 30 20 10 0 10 /call/min 20 30 40 50 60 0 1 2 3 4 5 6 7 8 9 1011121314151617181920212223

Number of pulses Time (hours)

Diel Rates of Growling

1994 20

15

10 Montane 5 0 growls/min

5 Sub-alpine 10

15

20 0 1 2 3 4 5 6 7 8 9 1011121314151617181920212223 Number of Time (hours/minutes)

363 1998 0.7 0.6 0.5 0.4 0.3 montane 0.2 0.1 0.00 growls/min 0.1 0.2 Sub-alpine 0.3 0.4 0.5 0.6 0.7 0 1 2 3 4 5 6 7 8 9 1011121314151617181920212223 Number of Time (hours/minutes)

364

Minerva Access is the Institutional Repository of The University of Melbourne

Author/s: Hollis, Gregory J.

Title: Ecology and conservation biology of the Baw Baw frog Philoria frosti (Anura: Myobatrachidae): distribution, abundance, autoecology and demography

Date: 2004-01

Citation: Hollis, G. J. (2004). Ecology and conservation biology of the Baw Baw frog Philoria frosti (Anura: Myobatrachidae): distribution, abundance, autoecology and demography. PhD thesis, Department of Zoology, University of Melbourne.

Publication Status: Unpublished

Persistent Link: http://hdl.handle.net/11343/39019

File Description: Ecology and conservation biology of the Baw Baw frog Philoria frosti (Anura: Myobatrachidae)

Terms and Conditions: Terms and Conditions: Copyright in works deposited in Minerva Access is retained by the copyright owner. The work may not be altered without permission from the copyright owner. Readers may only download, print and save electronic copies of whole works for their own personal non-commercial use. Any use that exceeds these limits requires permission from the copyright owner. Attribution is essential when quoting or paraphrasing from these works.