Development and application of subfossil chironomid-based methods for late

Quaternary climate reconstructions in eastern Australia

Jie Christine Chang

BE (Chemical Engineering)

GDip NatResSt

A thesis submitted for the degree of Doctor of Philosophy at

The University of Queensland in 2015

School of Geography, Planning and Environmental Management

Abstract

The lack of tools to quantify past climate has hindered the development of our knowledge of climate change in the late Quaternary in eastern Australia. This thesis developed two methods using subfossil remains of non-biting midge larvae (Chironomid, Diptera: ) preserved in lake sediments to reconstruct past changes in the Australian climate system during and since the Last Glacial Maximum (LGM). Firstly a transfer function model for reconstructing past summer temperatures based on the temperature tolerance of modern Australian chironomid taxa was created. Secondly the stable oxygen and hydrogen isotope composition (δ18O, δ2H) of the chitinous head capsules was used to determine temperature changes from chironomids from the same lakes. Both techniques were tested and the complexities of applying these methods as independent temperature proxies in Australia were explored.

In order to develop the transfer function, forty-five lakes were examined in eastern Australia for eighteen physical and chemical parameters. The physical and chemical analysis of the lakes suggests that sub-humid and semi-arid sites are naturally eutrophic, whereas high altitude and actively managed sites were mostly mesotrophic. Thirty-four lakes were subsequently used in the transfer function. The transfer function first axis was Mean February Temperature which explains 9.5% of the variance and the secondary axis is pH which also explains 9.5% of the variance. 2 Despite these low values the function appears robust. The transfer function had an r jack-knifed of 0.69 and a root mean squared error of prediction (RMSEP) of 2.33°C.

The transfer function method was then applied to subfossil chironomids from a subtropical site on North Stradbroke Island, Australia that spans the LGM and the last deglaciation. This has provided the first quasi-continuous quantitative temperature reconstruction from mainland terrestrial Australia covering these critical periods (between c. ~ 23.2 and 15.5 cal ka BP). The results of this reconstruction show a maximum cooling of c. ~6.5°C had occurred at c. ~18.5 cal ka BP during the LGM. A rapid warming followed and temperatures reached near Holocene values by c. ~17.3 cal ka BP. The warming trend started at c. ~18.1 cal ka BP and is consistent with the start of deglaciation from Antarctic records. The records suggest high latitude influencing of subtropical climate during the LGM.

Stable isotopes (δ18O and δ2H) from a single genus of chironomid head capsules ( spp.) were evaluated to reconstruct past temperatures also. Results suggest that both δ18O and δ2H are

i | P a g e

potentially valuable tools for reconstructing temperature in humid and low nutrient regions of Australia (e.g. Tasmania and the southeastern highlands). The correlation between δ18O of chironomid head capsules and temperature is consistent with observations from Europe and indicates that there is potential for this technique. However, the collection of enough material from ancient sites is challenging and the use of stable isotopes is recommended for late Holocene and modern studies, perhaps focussed on nutrient changes rather than temperature.

ii | P a g e

Declaration by author This thesis is composed of my original work, and contains no material previously published or written by another person except where due reference has been made in the text. I have clearly stated the contribution by others to jointly-authored works that I have included in my thesis.

I have clearly stated the contribution of others to my thesis as a whole, including statistical assistance, survey design, data analysis, significant technical procedures, professional editorial advice, and any other original research work used or reported in my thesis. The content of my thesis is the result of work I have carried out since the commencement of my research higher degree candidature and does not include a substantial part of work that has been submitted to qualify for the award of any other degree or diploma in any university or other tertiary institution. I have clearly stated which parts of my thesis, if any, have been submitted to qualify for another award.

I acknowledge that an electronic copy of my thesis must be lodged with the University Library and, subject to the policy and procedures of The University of Queensland, the thesis be made available for research and study in accordance with the Copyright Act 1968 unless a period of embargo has been approved by the Dean of the Graduate School.

I acknowledge that copyright of all material contained in my thesis resides with the copyright holder(s) of that material. Where appropriate I have obtained copyright permission from the copyright holder to reproduce material in this thesis.

Jie Chang

iii | P a g e

Publications during candidature

Peer-reviewed Journal articles Incorporated into the Thesis

Chang, J., Woodward, C., Shulmeister, J. (2014). A snapshot of the limnology of eastern Australian water-bodies spanning the tropics to Tasmania: The land-use, climate, limnology nexus. Marine and Freshwater Research: 65, 872-883

Chang, J., Shulmeister, J., Woodward, C. (2015a). A chironomid based transfer function for reconstructing summer temperatures in south eastern Australia. Paleogeography, Paleoclimatology, Paleoecology: 423, 109-121

Chang, J., Shulmeister, J. Woodward, C., Steinberger, L., Tibby, J., Barr, C. (2015b). A chironomid-inferred summer temperature reconstruction from subtropical Australia during the last glacial maximum (LGM) and the last deglaciation. Quaternary Science Reviews: 122, 282-292

Chang, J., Shulmeister, J., Woodward, C., Michalski, G. (Submitted). Can stable oxygen and hydrogen isotopes from Australian subfossil chironomid head capsules be used as proxies for past environmental change? Limnology and Oceanography

iv | P a g e

Conference Presentations

Chang, J., Woodward, C., Shulmeister, J., Zawadzki, A., Jacobsen, G. 2012. New data from Little Llangothlin Lagoon, New England Tablelands, eastern Australia, indicate no significant post-European erosion. In: 15th Biennial Conference of the Australia and New Zealand Geomorphology Group (ANZGG), Bundanoon, Australia, 2nd -7th December, 2012

Chang, J., Shulmeister, J., Woodward, C. 2013. Developing methods that use Australian chironomid (non-biting midge) larvae as proxies for past climate change. In: Asia-Oceania Geoscience Society (AOGS) 10th Annual Meeting, Brisbane, Australia, 24th -28th June, 2013

Chang, J., Shulmeister, J., Theiling, B. 2013. Application of stable isotopes from Australian chironomid (non-biting midge) head capsules as proxies for past climate change. In: 12th Australasian Environmental Isotope Conference (AEIC), Perth, Australia, 10th -12th July, 2013 [awarded Australian Nuclear Science Technology Organization (ANSTO) Student Travel Grant]

Chang, J., Shulmeister, J., Woodward, C. 2014. Development and application of a chironomid based transfer function for reconstructing summer temperatures in south eastern Australia. In: Australasian Quaternary Biennial Conference 2014, Mildura, Australia, 30th June – 3rd July, 2014

Chang, J., Shulmeister, J., Woodward, C., Steinberger L., Tibby, J. 2014. Using chironomid-based transfer function and stable isotopes for reconstructing past climate in south eastern Australia. In: American Geophysical Union Fall Meeting 2014, San Francisco, USA, 15th – 19th Dec, 2014

Chang. J., Shulmeister, J., Woodward, C., Steinberger L., Tibby, J. Barr, C. 2015. (Accepted for oral presentation) A high resolution chironomid-based summer temperature reconstruction from North Stradbroke Island, Australia, during the last glacial maximum (LGM). In: XIX International Quaternary Association (INQUA) Congress 2015, Nagoya, Japan, 28th July – 1st Aug, 2015 [awarded INQUA Early Career Researchers Travel Grant]

v | P a g e

Chang, J., Shulmeister, J., Woodward, C. Michalski, G. (Accepted for poster presentation). The potential for applying stable oxygen and hydrogen isotopes in subfossil chironomids from southeastern Australia as proxy for past changes. In: XIX International Quaternary Association (INQUA) Congress 2015, Nagoya, Japan, 28th July – 1st Aug, 2015

vi | P a g e

Publications included in this thesis

Publication citation – incorporated as Chapter 3.

Chang, J., Woodward, C., Shulmeister, J. (2014). A snapshot of the limnology of eastern Australian water-bodies spanning the tropics to Tasmania: The land-use, climate, limnology nexus. Marine and Freshwater Research: 65 (10), 872-883

Contributor Statement of contribution Jie Chang (Candidate) Research design (70%) Field work (70%) Statistical analyses (85%) Wrote and edited the paper (70%) Craig Woodward Research design (15%) Field work (25%) Statistical analyses (15%) Wrote and edited the paper (15%) James Shulmeister Research design (15%) Field work (5%) Wrote and edited the paper (15%)

Publication citation – incorporated as Chapter 4.

Chang, J., Shulmeister, J., Woodward, C. (2015). A chironomid based transfer function for reconstructing summer temperatures in south eastern Australia. Paleogeography, Paleoclimatology, Paleoecology: 423, 109-121

Contributor Statement of contribution Jie Chang (Candidate) Research design (70%) Field work (60%) Laboratory work (100%) Statistical analyses (85%) Wrote and edited the paper (70%) James Shulmeister Research design (15%)

vii | P a g e

Field work (10%) Wrote and edited the paper (20%) Craig Woodward Research design (15%) Field work (30%) Statistical analyses (15%) Wrote and edited the paper (10%)

Publication citation – incorporated as Chapter 5.

Chang, J., Shulmiester, J., Woodward, C. Steinberger, L., Tibby, J., Barr, C. (2015). A chironomid- inferred summer temperature reconstruction from subtropical Australia during the last glacial maximum (LGM) and the last deglaciation. Quaternary Science Reivews: 122, 282-292

Contributor Statement of contribution Author Jie Chang (Candidate) Research design (80%) Laboratory work (80%) Statistical analyses (95%) Wrote and edited the paper (70%) Author James Shulmeister Research design (10%) Wrote and edited paper (15%) Author Craig Woodward Research design (5%) Statistical analyses (5%) Wrote and edited paper (5%) Author Lincoln Steinberger Laboratory work (20%) Author John Tibby Research design (5%) Field work (50%) Wrote and edited paper (5%) Author Cameron Barr Field work (50%) Wrote and edited paper (5%)

viii | P a g e

Publication citation – incorporated as Chapter 6.

Chang, J., Shulmeister, J., Woodward, C. Michalski, G. (submitted to Limnology and Oceanography). Can stable oxygen and hydrogen isotopes from Australian subfossil chironomid head capsules be used as proxies for past environmental change?

Contributor Statement of contribution Author Jie Chang (Candidate) Designed experiments (70%) Field work (60%) Laboratory work and analyses (90%) Statistical analyses (100%) Wrote the paper (70%) Author James Shulmeister Designed experiments (20%) Field work (10%) Wrote and edited paper (20%) Author Craig Woodward Designed experiments (10%) Field work (30%) Wrote and edited paper (5%) Author Greg Michalski Laboratory work and equipment training (10%) Wrote and edited paper (5%)

Contributions by others to the thesis

Standard contributions were made by the supervision team of James Shulmeister and Craig Woodward. This includes shared project design and standard editing and critical revision of work, which substantially improved the quality and clarity of the research. In addition, John Tibby and Cameron Barr were responsible for the collection and radiocarbon dating of the sediment core from Welsby Lagoon presented in Chapter 5 (Chang et al., 2015b). Lincoln Steinberger assisted in laboratory work for picking chironomids from Welsby Lagoon sediment. Greg Michalski was responsible for the training of using stable isotope equipment at Purdue Stable Isotope Facility.

Statement of parts of the thesis submitted to qualify for the award of another degree None

ix | P a g e

Acknowledgements There are so many people to thank throughout the course of my PhD candidature. I would like to start with thanking my principal advisor Jamie Shulmeister. Jamie has played more than just a supervisor role in the past 3 and half years to me. He has been also a great mentor and a wonderful friend to me since I started my postgraduate research project with him in early 2011. He has guided me all the way through till this point and it has not always been easy. Jamie got me interested in this field and has successfully converted me from being potentially a chemical engineer to a palaeoecologist. Today, I still think it was one of the best decisions I made. I deeply appreciate and respect Jamie for his enthusiasm, approachability and vision/insight of our research field. I particularly thank Jamie for his approach of training me up to a new level to become an independent researcher and enthusiastic scientist. I would also like to thank Jamie’s family, wife Val and daughter Niamh for their warm hospitality on our ‘get together’ BBQs/Mexican nights.

I would like to thank Craig Woodward as my other advisor for his substantial input into the project and paper ideas, but in particular, his input and training for me in the field and the lab. Together with Jamie, they are an absolutely great supervision team. I was inexperienced with field work when I started my PhD project, and when I say ‘inexperienced’, I mean ‘never been to any places out of the city’ to do any work. Craig helped and taught me the art of field work from scratch, from planning, equipment organisation and moreover, getting the boat out of a lake to get a nice sediment core out. I’m thankful for Craig’s patience when it comes to species . He trained me from ‘how to use the microscope’ and had to put up with questions such as ‘what is this one again? why does it have extra teeth?’ throughout the first at least six months of my PhD. I would also like to thank both Jamie and Craig for their thoughtful and careful feedback on papers, manuscript revisions, and thesis drafts.

I would like to thank John Tibby, Cameron Barr and Lincoln Steinberger for their input into the work related to Welsby lagoon. I particularly thank John and Cameron for giving me the opportunity to work on the site and their constructive feedback on my research related to the site. I’m grateful for the assistance in picking chironomids from Welsby lagoon provided by Lincoln. It saved me at least a few weeks of time in the lab. I also thank Greg Michalski from Purdue University for providing stable isotope equipment training to me during my visit at Purdue. I would also like to thank him and his family for letting me stay at their place and for their warm hospitality throughout my visit. I would also like to mention Tanya Katzman from Purdue for her assistance in the stable isotope laboratory. In particular, I would like to thank Matthew Jones from Nottingham for having fruitful discussions with me about stable isotopes during his visit here at UQ. I would x | P a g e

also like to thank Andrew Rees for sharing pictures of Tasmanian chironomids with me and this has greatly helped me to develop my knowledge on taxonomy.

I would like to thank the Department of Primary Industries, Water and Environment (DPIWE), the Department of Sustainability and Environment (DSE) and the Department of Environment and Heritage Protection (DEHP) for the permission of sample collection for the project. I have been very fortunate to have a few wonderful field assistants. I thank Abdollah (Ben) Jarihani, Lydia Mackenzie for their outstanding assistance during the field seasons in Tasmania and Victoria respectively. I would also like to mention Adrian Slee for his assistance in lake coring during the Tasmanian field season. I have been very lucky to have made some wonderful friends through working on my project also at GPEM and these include Abdollah (who is also a wonderful office mate), Lydia, Lincoln, Adrian, Dan (Sabrina), David, Nicola, Sui, Emily, Daniel, Dong, Min, Ming etc.

I am very grateful to have Tim Cohen and Patrick Moss as my academic referees and I would like to thank them for providing wonderful recommendation letters for me which helped me get a Postdoc job lined up. I would like to thank my future advisor Patricio Moreno for providing me an exceptional opportunity to continue my research work in northwestern Patagonia for the next couple of years. In addition, I would like to thank Martin Williams and Peter Cranston for their encouragement and interests in my on-going and future research. I would like to thank all the professional and technical staff at GPEM. They have been a very supportive team. I especially thank Michael Tobe, Alan Victor and Clwedd Burns for their assistance in setting up laboratory and field equipment.

I have been very lucky to receive a number of grants to fund my research and to attend conferences. The ARC discovery grant has provided adequate funding for me to do field work and attend a number of conferences, including Australia and New Zealand Geomorphology Group meeting in Bundanoon and Mount Tamborine, Australasian Quaternary Association (AQUA) meeting in Mildura, the International Quaternary Association (INQUA) early career researchers (ECR) meeting in Wollongong. The ARC discovery top-up scholarship and the University of Queensland Research Scholarship allowed me to fully focus on my research. I also received the Graduate School International Travel Award which allowed me to visit Purdue to do laboratory work. I also received the postgraduate travel grant provided through Australia Nuclear Science and Technology Organisation (ANSTO) to attend the Australasian Environmental Isotope Conference in Perth. I also received the GPEM school travel grant to attend the American Geophysical Union Meeting at xi | P a g e

San Francisco. More recently, I received the INQUA travel grant to attend the congress in Nagoya, Japan.

Last, I would like to thank my family. My parents have been very supportive and encouraging throughout my PhD. I am grateful for both of their emotional and financial backing. This PhD would not be possible without them. I would like to thank my fiancé David McCann and his family (Dale, Steve, Mel and Dave). I am very lucky to have David to share this journey with me. I believe this past 3 and half years will be a very special lifetime memory that we will talk about in the future. I would like to specially thank David for comforting and putting up with me when I was stressed and frustrated during my PhD. I particularly thank David for being supportive for me to pursue my career because he knows what makes me happy.

Finally, I would like to thank the two thesis examiners, Professors Ian Walker and Isabelle Larocque-Tobler for their useful and constructive suggestions.

xii | P a g e

Keywords Southeast Australia, freshwater lakes, subfossil chrionomids, transfer function, temperature, palaeoclimate reconstruction, last glacial maximum, last deglaciation, stable isotopes

Australian and New Zealand Standard Research Classifications (ANZSRC)

ANZSRC code: 060206, Paleoecology, 45% ANZSRC code: 040605, Paleoclimatology, 45% ANZSRC code: 040203, Isotope Geochemistry, 10%

Fields of Research (FoR) Classification

FoR code: 0406, Physical Geography and Environmental Science, 80% FoR code: 0402, Geochemistry, 10% FoR code: 0602, Ecology, 10%

xiii | P a g e

Contents 1 INTRODUCTION ...... 1

1.1 BACKGROUND ...... 1

1.2 RESEARCH OBJECTIVES ...... 5

1.3 OVERVIEW OF RESEARCH METHODOLOGY ...... 6

1.4 THESIS STRUCTURE AND SYNOPSIS ...... 9 2 THE BIOLOGY OF CHIRONOMIDS ...... 11

2.1 THE PRINCIPLES OF APPLYING THE TRANSFER FUNCTION AND STABLE ISOTOPE METHODS IN

PALAEOCLIMATE RECONSTRUCTIONS ...... 14 2.1.1 The transfer function method ...... 14 2.1.2 The stable isotope method ...... 14 3 A SNAPSHOT OF THE LIMNOLOGY OF EASTERN AUSTRALIAN WATER BODIES SPANNING THE TROPICS TO TASMANIA: THE LAND-USE, CLIMATE, LIMNOLOGY NEXUS ...... 16

3.1 SUMMARY ...... 16

3.2 ABSTRACT ...... 17

3.3 INTRODUCTION ...... 18 3.3.1 Previous water-chemistry studies of Australian lakes ...... 18 3.3.2 Artificial water bodies ...... 21

3.4 MATERIALS AND METHODS ...... 24 3.4.1 Study sites ...... 24 3.4.2 Sample collection and laboratory analyses ...... 27 3.4.3 Spatial data ...... 27 3.4.4 Statistical analyses ...... 28

3.5 RESULTS ...... 30 3.5.1 Lake-water chemistry ...... 31 3.5.2 PCA results ...... 33 3.5.3 RDA results ...... 38

3.6 DISCUSSION ...... 41 3.6.1 The effect of including reservoirs and other artificial waterways ...... 41 3.6.2 Salinity, pH and bedrock effects ...... 41 3.6.3 Nitrogen and phosphorus limitation ...... 42 3.6.4 The interaction of human impact and climatic effects on lake nutrient status ...... 42

3.7 CONCLUSIONS ...... 44

xiv | P a g e

4 A CHIRONOMID BASED TRANSFER FUNCTION FOR RECONSTRUCTING SUMMER TEMPERATURES IN SOUTH EASTERN AUSTRALIA ...... 45

4.1 SUMMARY ...... 45

4.2 ABSTRACT ...... 46

4.3 INTRODUCTION ...... 47

4.4 MATERIALS AND METHODS ...... 51 4.4.1 Study sites ...... 51 4.4.2 Chironomid collection and analysis ...... 51 4.4.3 Lake water chemistry and environmental variables ...... 52 4.4.4 Statistical analyses ...... 54

4.5 RESULTS ...... 56 4.5.1 Chironomid taxa ...... 56 4.5.2 Test for Tasmanian endemism and difference between natural and artificial lakes ...... 56 4.5.3 Selection of environmental variables and model development ...... 61 4.5.4 The transfer function ...... 65

4.6 DISCUSSION ...... 71 4.6.1 Can we include artificial lakes in temperature training sets? ...... 71 4.6.2 Endemism and other considerations ...... 71 4.6.3 Reconciliation and integration with Tasmanian transfer function ...... 72 4.6.4 Value of the transfer function ...... 73

4.7 CONCLUSIONS ...... 74 5 A CHIRONOMID-INFERRED SUMMER TEMPERATURE RECONSTRUCTION FROM SUBTROPICAL AUSTRALIA DURING THE LAST GLACIAL MAXIMUM (LGM) AND THE LAST DEGLACIATION ...... 75

5.1 SUMMARY ...... 75

5.2 ABSTRACT ...... 76

5.3 INTRODUCTION ...... 77

5.4 REGIONAL SETTING ...... 78 5.4.1 Climate ...... 79 5.4.2 Modern circulation of the Eastern Australian Current (EAC) ...... 79 5.4.3 Physical environment of Welsby Lagoon ...... 80

5.5 METHODOLOGY ...... 81 5.5.1 Chironomids sampling and analysis ...... 81 5.5.2 Chronology ...... 82 5.5.3 Statistical methods and reconstruction examination ...... 83 xv | P a g e

5.6 RESULTS ...... 84 5.6.1 Chironomid analysis ...... 84 5.6.2 The LGM (~23.2 cal ka BP – 18.1 cal ka BP) ...... 85 5.6.3 The deglacial period (~18.1 cal ka BP – 13 cal ka BP) ...... 86 5.6.4 The Holocene (~13 cal ka BP – present) ...... 86 5.6.5 Relationships between changes in chironomid assemblages and temperature and other variables ...... 89

5.7 DISCUSSION ...... 93 5.7.1 Precision and reliability of the reconstruction ...... 93 5.7.2 The deglacial and LGM temperatures of North Stradbroke Island and regional climate 97 5.7.3 Wider paleoclimatic considerations ...... 99

5.8 CONCLUSION ...... 100 6 CAN STABLE OXYGEN AND HYDROGEN ISOTOPES FROM AUSTRALIAN SUBFOSSIL CHIRONOMID HEAD CAPSULES BE USED AS PROXIES FOR PAST ENVIRONMENTAL CHANGE? ...... 101

6.1 SUMMARY ...... 101

6.2 ABSTRACT ...... 102

6.3 INTRODUCTION ...... 103

6.4 MATERIALS AND METHODS ...... 105 6.4.1 Study sites and sample collection ...... 105 6.4.2 Climatic and stable isotopes in precipitation data ...... 106 6.4.3 Stable isotope sample preparation ...... 106 6.4.4 Stable isotope analyses ...... 108 6.4.5 Statistical analyses ...... 109

6.5 RESULTS ...... 110 6.5.1 Stable oxygen isotope (δ18O) analyses results ...... 111 6.5.2 Stable hydrogen isotope (δ2H) results ...... 119

6.6 DISCUSSION ...... 124 6.6.1 Analytical considerations ...... 124 6.6.2 Relationship between δ18O of lake water and regional climate ...... 125 6.6.3 Lake trophic status related δ18O fractionation process and the metabolism and respiration of Chironomus spp...... 126 6.6.4 Relationship between δ18O of Chironomus spp. and regional climate ...... 129 6.6.5 The δ2H of precipitation, lake water, Chironomus spp. and temperature ...... 129 xvi | P a g e

6.7 CONCLUSIONS AND FUTURE WORK ...... 132 7 CONCLUSIONS AND FUTURE WORK ...... 134

7.1 SUMMARY OF RESEARCH AND MAJOR FINDINGS ...... 134 7.1.1 Connections of modern limnology to climate and land-use ...... 134 7.1.2 The transfer function ...... 135 7.1.3 The application of the transfer function: temperature reconstruction covering the Last Glacial Maximum and the last deglaciation from subtropical Australia ...... 137 7.1.4 The potential of applying stable oxygen and hydrogen isotopes from Australian subfossil chironomid head capsules as independent proxies for past change ...... 138

7.2 RESEARCH SIGNIFICANCE ...... 139

7.3 CRITICAL REFLECTIONS ...... 139 7.3.1 The debate and controversy about transfer function and its application ...... 140 7.3.2 The limitation of the stable isotope method ...... 141

7.4 RECOMMENDATIONS AND FUTURE WORK ...... 142 7.4.1 Extension and integration of the modern calibration dataset in Australia ...... 142 7.4.2 Application of the transfer function to quantify long-term (late-Pleistocene to Holocene) temperature changes from various sites ...... 143 7.4.3 Further exploration and application of the chironomid δ18O and δ2H method ...... 144 7.4.4 Regional patterns and hemispheric to global connections ...... 144 REFERENCES ...... 146 APPENDICES ...... 168

APPENDIX 1. A SUMMARY OF PREVIOUS MODERN CALIBRATION TRAINING SETS

DEVELOPED AND APPLIED FOR PALEO-RECONSTRUCTIONS USING CHIRONOMIDS ...... 168

APPENDIX 2. CHIRONOMID TAXA ENUMERATED FROM THE MODERN

CALIBRATION DATA SET OF SOUTHEASTERN AUSTRALIA ...... 175

APPENDIX 3. CHIRONOMID FOSSIL TAXA ENUMERATED FROM WELSBY LAGOON,

NORTH STRADBROKE ISLAND, AUSTRALIA FROM THE LAST GLACIAL MAXIMUM

TO NEAR PRESENT ...... 195

APPENDIX 4. PUBLISHED JOURNAL ARTICLES DURING CANDIDATURE ...... 203

xvii | P a g e

List of Figures Figure 1.1 Overview of the framework and project flow ...... 7 Figure 1.2 Overview of the thesis structure ...... 10 Figure 2.1 Key morphological characteristics of chironomid larvae (Brooks et al., 2007, Page 4) .. 11 Figure 2.2 Chironomid Life Cycle (Walker, 1987, Page 852) ...... 12 Figure 2.3 A fully developed chironomid head capsule (Chironomus sp.) from Lake Albina, Mt Kosciuszko, New South Wales, Australia (photo is taken by J. Chang at the Physical Geography laboratory, School of Geography, Planning and Environmental Management, University of Queensland) ...... 12 Figure 2.4 Chitin molecule structure...... 13 Figure 3.1 Map of eastern Australia, with the 45 study sites identified; the numbers correspond to the lake names and numbers in Table 3.1...... 19 Figure 3.2 Tri-plot of major cations, with lakes labelled by type, as follows: CL, coastal lakes (lowland lakes that are <100 km away from the sea); M, montane lakes (>1000 m asl); IL, inland lakes (lowland lakes that are >100 km away from the sea). The tri-plot shows cation concentration as a percentage. All 45 lakes are distributed closer to the world-average seawater ratio (WASW) than to the world-average freshwater ratio. On average, lowland coastal lakes are closer to the ratio than inland lakes or those at high elevation. This indicates that the primary source of the salts in eastern Australian lakes is seawater...... 32 Figure 3.3 Biplot of pH against conductivity. This demonstrates that pH co-varies with bedrock type and salinity. All saline lakes are strongly alkaline, whereas there is more variability in the pH of freshwater lakes. This means that the freshwater lakes are dominantly controlled by bedrock types. Lakes above log conductivity of 3 are classified as brackish...... 33 Figure 3.4 Initial PCA plot of a) the 45 water bodies constrained to b) the 19 environmental variables, with land-use, climate, and lake depth variables entered as passive variables to focus the analysis on in-lake (limnological variables). The first and second axes of this plot are predominantly specific conductance and pH gradients; it shows a large number of correlated water conductivity proxies (ion concentrations)...... 35 Figure 3.5 Revised PCA plot of a) the 45 water bodies constrained to b) the 12 environmental variables (ionic variables removed), with land-use, climate, and lake depth variables entered as passive variables to focus the analysis on the in-lake (limnological) variables. The first axis is now a nutrient gradient and the second axis is a pH gradient...... 36 Figure 3.6 Final principal components analysis (PCA) with all environmental data (excluding ion concentration data) entered as active variables. (a) The 45 water bodies, along with (b)

xviii | P a g e

morphology, environment, climate and land-use variation patterns. The primary y-axis is a trophic status (total nitrogen (TN), total phosphorus (TP), Chlorophyll a), conductivity and agriculture gradient; the second axis is aligned to temperature. The PCA also indicates that eutrophic saline lakes tend to be shallow lakes located in arid regions...... 37 Figure 3.7 Principal components analyses (PCAs) to compare the changes in limnology for natural lakes and artificial lakes along a gradient that reflects lake depth and human impact. The Axis 1 sample scores from plot (a) shows the amount of agricultural activity in the lake catchments. The Axis 1 sample scores from plot (b) shows the trophic status of the lakes. The Axis 1 sample scores for both PCAs (a and b) are used to visualise the relationship between land use, aridity and limnology. The regression for (c) all lakes and (d) artificial lakes only. The trend of the regression is similar between the two graphs, indicating that artificial lakes have traits similar to those of natural lakes. The lower slope on the artificial lakes may indicate lower residence times. Overall, these results demonstrate that lakes in agricultural catchments are more eutrophic; however, see the main text because this trend is overwritten by climatic effects and lake morphology...... 40 Figure 4.1 (a) Plot of observed species richness vs predicted species richness, (b) rarefaction curves for individual sites indicating estimated species richness with respect to increasing sub-sample size. Rarefaction curves begin to flatten once true species richness is achieved. Together, these plots indicate that sample sizes are adequate for most samples to capture all but the most rare species. Only one site (CTL) with a low head capsule count may possibly underestimate the true species richness. Minimum sample size varied from site to site and counts as low as 50 may be sufficient to capture the actual species richness (e.g. LPO)...... 53 Figure 4.2 PCA plots for exploring the difference between (a) mainland and Tasmanian lakes and (b) natural and artificial lakes. PCA axis 1 and axis 2 explains 16.9% and 10.8% of the variance in chironomid species data respectively. A t-test was performed on the sample score means for each (Tables 4.1a and 4.1b). The sample size for the Tasmania and artificial lakes is small (<10). There are no significant differences apparent between axis 1 scores for both tests. There are no differences in axis 2 scores either for Tasmania vs mainland lakes, but there is for natural vs artificial lakes. Warm taxa are over-represented in artificial lakes that are not reservoirs (Fig. 4.3)...... 57 Figure 4.3 This shows a PCA of chironomid species used to compare natural and artificial lakes. In a) PCA axis 2 show separation of warm stenotherms (low PCA axis 2 scores, e.g. Dicrotendipes, Kiefferulus, Cladopelma, Procladius, Paratanytarsus etc.) from cold stenotherms (high PCA axis 2 scores, e.g. Parakiefferiella). b) shows surface sediment samples plotted as circles where the circle size is relative to mean February temperature (MFT). Large xix | P a g e

circles = high MFT, small circles = low MFT. Values for MFT for each site are also provided. Open circles = natural water bodies, closed circles = artificial water bodies. Note that some artificial water bodies (not reservoirs) with low MFTs have low PCA axis 2 scores, which indicates that warm stenotherms are more common in these sites compared to natural sites with similar MFTs...... 58 Figure 4.4 CCA biplots of (a) sample and (b) species scores constrained to seven environmental variables that individually explain a significant (p ≤ 0.05) proportion of the chironomid species data. Sites codes correspond to site names in Tables 3.1 and 3.2 (lake numbers 1-34). Taxon numbers correspond to taxa in Table 4.5. Sites and taxa in warmer environments tend to plot in the upper left quadrant. Taxa typical of warmer sites include Harnischia spp., Cryptochironomus spp., Polypedilum spp., Coelopynia pruinosa type, Cladopelma spp., Paratanytarsus spp., Procladius spp., and Riethia spp., while sites and taxa in colder environments tend to plot in the lower right hand quadrant. Taxa typical of cold sites include Paralimnophyes morphotype 3, Parakiefferiella morphotype 1, Orthoclad type 1, Orthoclad type 4, Pseudosmittia type 2, and Botrycladius. Eutrophic sites and taxa typical of these environments tend to plot in the lower left hand quadrant, while oligotrophic sites and taxa typical of these environments tend to plot in the upper right hand quadrant...... 66 Figure 4.5 Taxon response curves for taxa that show a significant response to temperature (p < 0.05) using a Generalized Linear Model with Poisson distribution (ter Braak and Šmilauer, 2002). (a) Taxa which are more common at lower temperatures, such as Paralimnophyes morphotype 1 and Parakiefferiella morphotype 2 respond strongly to cooling in temperature (b) Taxa which are more common at higher temperatures demonstrate weaker but still significant responses. Examples include Procladius, Polypedilum spp. and Tanytarsus lactescens type...... 67 Figure 4.6 Stratigraphy diagram of the 38 non-rare taxa included in the final model, where observed mean February temperature is on the y-axis and taxon abundance is in percentage. Taxa such as Cladopelma, Dicrotendipes and Kiefferulus (*) show high abundance in lowland warm lakes but are also present in highland artificial lakes (Grey bars)...... 68 Figure 4.7 Performance of the three component PLS model where (a) shows the predicted versus observed mean February temperature and (b) displays residuals of the predicted versus observed mean February temperature. Note that the model has a potential to over predict temperatures from some very shallow high altitude lakes by up to ~6 °C. These lakes have increased mean water temperature in summer and chironomid assemblages may resemble lower elevation sites...... 73 Figure 5.1 Location of the study site and other sites discussed in this study. (a) A summary of the surface currents within the Tasman Sea. The denoted by A, B, C, and D shows the direction of xx | P a g e

flow of the relevant part of the EAC along with where the SST records are mentioned. Note that B is associated with the Tasman Front (Ridgway and Dunn, 2003). The locations of marine cores PC-27 (Dunbar and Dickens, 2003), GC-12 (Bostock et al., 2006), GC-25 (Troadson and Davies, 2001), PC-9 (Troadson and Davies, 2001) along the EAC, terrestrial temperature estimates from Lake Victoria (Miller et al., 1997), the Snowy Mountains (Galloway, 1965), and Lake Selina (Fletcher and Thomas, 2010) are displayed (b) Location of North Stradbroke Island (NSI) and Lake Mackenzie, Fraser Island (c) Location on North Stradbroke Island of Welsby Lagoon and other sites mentioned in the text, including Tortoise Lagoon (Petherick et al., 2008; Moss et al., 2013), Blue Lake (Barr et al., 2013), Eighteen Mile Swamp (sampling location) and Native Companion Lagoon (McGowan et al., 2008; Moss et al., 2013)...... 81 Figure 5.2 Radiocarbon and calibrated ages for Welsby Lagoon with respect to depth. Open shapes represent calibrated ages calibrated using SHCal13 (Hogg et al., 2013)...... 82 Figure 5.3 Chironomid stratigraphy from Welsby Lagoon plotted with respect to the radiocarbon chronology. Chironomid inferred mean February temperatures (MFTs) are shown in (a) and the number of head capsules counted for each sample are shown in (b). The dashed line in (a) indicates the current mean February temperature at Welsby Lagoon and the dashed line in (b) indicates the minimum number of head capsules for a quantitative reconstruction from a sample. All fossil taxa present in the Welsby Lagoon record are included in the diagram. Age in calibrated years is on the y-axis and taxa were grouped according to their optima in the modern training sets based on Chapter 4. Grey areas represent sediment layers where chironomids were absent. Note that Dicrotendipes, a benthic taxon that does not have a significant correlation (p > 0.05) to MFT in the modern calibration data set, is dominant in all four Holocene samples. Each of the taxon was photographed and included in Appendix 3...... 88 Figure 5.4 Reconstruction diagnostics for Welsby Lagoon samples. (a) Reconstructed mean February temperatures (MFTs) with the RMSEP (2.2 °C) from the modern training set (Chapter 4) plotted in dashed horizontal lines for each sample. (b) Goodness-of-fit statistics where the fossil samples are passively fitted to the CCA ordination axis derived from the modern training set constrained to MFT. The vertical dashed line represents the 95th percentiles of modern squared residual lengths beyond which fossil samples are considered to have poor fits to MFT. (c) Rare taxa plot where closed circles represent fossil samples with well- represented taxa in the modern training set and open circles represent fossil samples with rare taxa (Hill's N2 ≤ 5) summing to greater than 10% abundance. (d) Non-analogue plot where closed circles indicate fossil samples with either close or good modern analogues, whereas open circles indicate fossil samples with no good modern analogues (at the 5th percentile cut- off level). The vertical lines represent the 2nd, 5th and 10th percentiles respectively...... 90 xxi | P a g e

Figure 5.5 The Welsby Lagoon chironomid record was split into three zones using the constrained incremental sums of squares (CONISS) function in Psimpoll (Bennett, 2002). W-1 represents the last glacial maximum (LGM), W-2 represents the deglacial and W-3 represents the Holocene. This zonation is consistent with the observed changes in chironomid assemblages shown in Fig. 5.3...... 91 Figure 5.6 Trajectory changes of trend through time of fossil samples (black circles) in the Welsby Lagoon record, passively plotted in a CCA of the training set lakes (open circles, where site 1– 33 are waterbodies from (Chapter 4, Table 4.1) and site 34 is Swallow Lagoon, a waterbody near Welsby Lagoon with similar physical characteristics) against significant variables (p < 0.05). (a) Samples covering the last glacial maximum (LGM) and deglacial period, where changes in trend are parallel to the mean February temperature (MFT) gradient, indicating that MFT is a primary controlling variable and drives the changes of chironomid assemblages in the samples from c. ∼23.2 to 16.6 cal ka BP. (b) samples covering the period of deglacial and Holocene from c. ∼16.3 to 0.2 cal ka BP. The trend of changes of chironomid assemblages in these fossil samples are parallel to the secondary gradient that is represented by nutrients, conductivity, pH and lake depth, indicating that non-climatic factors are likely dominant in Holocene changes...... 92 Figure 5.7 Chironomid mean February temperature (MFT) transfer function based reconstruction 18 from Welsby Lagoon (WEL) (in plot d) compared to (a) the δ O and CO2 records from Antarctica (Stenni et al., 2011; Pedro et al., 2012); (b) insolation curve for seasonality (Berger and Loutre, 1991); (c) δ18O measured from G. ruber from marine cores along the East Australian Current (EAC) and (d) temperature estimated based on branched glycerol dialkyl glycerol tetraether (GDGT) from Fraser Island (Woltering et al., 2014). The chironomid results from WEL (d) indicate a maximum cooling relative to the modern MFT along the Southeast Queensland coast of c. ∼6.5 °C with much lower cooling at other times during the late LGM and a rapid recovery at about 18.1 cal ka BP. The cooling is significantly less than the 9–12 °C for the adjacent uplands from periglacial landforms from Galloway (1965), but is in line with estimates from nearby marine cores shown in (c) and lacustrine record from Lake Mackenzie (LM), Fraser Island shown in (d) (note the reconstructed temperature from Fraser Island is lower than Welsby Lagoon because the GDGT is based on mean annual temperatures calibration whereas chironomid transfer is based on MFT). It is also noteworthy that the chironomids provide an estimate of summer temperatures whereas the periglacial landforms are indicators of winter temperature. This suggests a larger seasonal temperature contrast at the LGM which is consistent with the insolation curves (b). Finally the observed timing and pattern

xxii | P a g e

18 is very similar to both the δ O curve and CO2 record from Antarctica (a) (Stenni et al., 2011; Pedro et al., 2012)...... 96 Figure 6.1 Sample lakes are distributed across southeast Australia but there is a strong concentration (ten out of 16 lakes) in western Victoria. This reflects Chironomus spp. abundance in waterways with the tolerant Chironomus spp. strongly represented in the brackish and saline lakes of western Victoria...... 107 Figure 6.2 Chironomus spp. head capsules from (a) Freshwater Lake, Victoria, 20 ×. (b) Lake Albina, Mount Kosciuszko, New South Wales, 10 ×. Photos of head capsules were taken by J. Chang at the Physical Geography Laboratory of School of Geography, Planning and Environmental Management, University of Queensland ...... 109 Figure 6.3 (a)Plot of δ18O of precipitation against δ18O of lake water. The δ18O of precipitation and lake water is correlated but significant enrichment of lake water is observed at δ18O of precipitation of around -5 ‰ V-SMOW, which the offset is possibly due to high regional evaporation rate and prolonged residence time or local hydrological conditions (b) the plot of δ18O of lake water against Chironomus spp. showed no signifcant correlation (P = 0.13) between the two data sets, suggesting that not all oxygen stable isotopic composition of Chironomus spp. head capsules were reflecting changes of their ambient water ...... 110 Figure 6.4 (a) Plot of aridity index (AI) against δ18O of Chironomus spp.. AI has the best correlation and is the independent variable that explains the largest percent variance (Table 1) of Chironomus spp. δ18O data. At semi-arid and sub-humid areas (e.g. when AI <1), the variation in δ18O of Chironomus spp. data is large. (b) Total Nitrogen (TN) against fractionation factor

(α[δ18O(chiro-water)]). TN explains an independent and large percent of variance in the fractionation factor between Chironomus spp. and host lake water (Table 6.2), suggesting lake eutrophication is an important consideration for the interpretation of stable isotope data for Chironomus spp...... 111 Figure 6.5 (a)Plot of δ2H of precipitation against δ2H of lake water. The δ2H of precipitation and lake water is closely correlated, with r2 = 0.92 (b) Plot of δ2H of lake water aginst δ2H of Chironomus spp.. δ2H of lake water and Chironomus spp. showed a close correlation (r2= 0.69) suggesting that a large proportion of δ2H stable isotopic composition of Chironomus spp. head capsules that were directly derived from their ambient water or indirectly from feeding on planktonic algae, where algae also contain δ2H stable isotopes that are derived from lake water...... 112 Figure 6.6 (a) Plot of Mean annual temperature (MAT) against δ2H of Chironomus spp.. MAT had the best correlation and is the independent variable that explains the largest percent variance (Table 6.5) of Chironomus spp. δ2H data. The correlation could possibly be complicated by the xxiii | P a g e

2 2 fractionation between δ H of Chironomus spp. and δ H of lake water (α[δ2H(chiro-water)]), which was influenced by other environmental variables. (b) Plot of ln(COND) against fractionation

factor α[δ2H(chiro-water). Specific conductance (COND) explained an independent and large percent of variance in the fractionation factor between δ2H of Chironomus spp. and host lake water (Table 6.6), lake water COND had a close correlation (r2 = 0.695) with the fractionation factor (α) between δ2H of Chironomus spp. and lake water...... 116 Figure 6.7 Plot of mean annual temperature (MAT) against δ18O of precipitation for the 16 study lakes. There is a positive correlation between MAT and δ18O of precipitation, but since the differences within the spatial distribution for both study sites and precipitation sources are large, the data points are scattered...... 126 Figure 6.8 Plot of the experimental chironomid δ18O values against δ18O of host water in Wang et al (2009) compared to values derived from this study. Eight of the lakes from this study (CTL, BL, LA, WP, LEA, LLL, LSP, LTK), fall along the regression line of Wang et al (2009) these are oligotrophic and mesotrophic lakes from Kosciuszko and Tasmania, and two western Victorian lakes that are less enriched in δ18O, i.e. less affected by prolonged residence time. The other eight lakes (FWL, LEM, LCT, LTP, SWL, LFY, LMB, NGR) produced similar Chironomus spp. δ18O values...... 127 Figure 6.9 (a) Plot of mean annual temperature (MAT) against δ18O of Chironomus spp. with all sixteen lakes included (b) Plot of mean annual temperature (MAT) against δ18O of Chironomus spp. with only the eight lakes that have the δ18O enrichment of lake water less than ~8‰ VSMOW relative to precipitation δ18O are included. There is a close correlation (r2 = 0.78) between MAT and Chironomus spp. δ18O for this subset of study lakes. The slope and intercept values were very similar to those observed by Verbruggen et al. (2011)...... 130 Figure 6.10 Plot of ln(COND) against mean annual temperature (MAT). Eight out of ten lakes from Western Victoria are considered saline. Lake Fyans (LFY) and Nuggetty Gully Reservoir (NGR) were the only lakes that had a specific conductance value < 500 μs/cm from this region. Lake order effect and residence time are potentially affecting the observed relationship between temperature and specific conductance and the overall pattern is presumably due to that regional evaporation is a function of temperature ...... 131 Figure 6.11 Eight lakes that had a ln(COND) < ~ 6 were plotted for mean annual temperature against δ2H of Chironomus spp.. A close correlation was obtained with an r2 of 0.72. These include two lakes (Lake Fyans and Nuggetty Gully Reservoir) from western Victoria and six lakes from non-arid environments. These two western Victorian lakes are freshwater bodies presumably due to short residence times...... 132

xxiv | P a g e

List of Tables Table 3.1 Summarised information of the 45 water bodies sampled ...... 23 Table 3.2 Selected limnological and land-use/land cover variables for the forty-five water bodies . 29 Table 3.3 List of the 19 variables used, with mean, minimum and maximum values cited and the appropriate units highlighted ...... 31 Table 3.4 Partial redundancy analysis (RDA) results for trophic status as limnological variable. This table demonstrates that the effect of climate and water depth is more significant than the agricultural status of the basin. In short, shallow water bodies in semi-arid settings are more eutrophic than are other lakes. AI, aridity index; Chl a, Chlorophyll a; TN, total nitrogen; TP, total phosphorus ...... 38 Table 4.1 Selected climatic and environmental variables for the thirty-four water bodies sampled from southeast Australia used for transfer function development ...... 50 Table 4.2 Statistical t-Tests for two samples assuming unequal variances performed on (a) natural against artificial lakes and (b) Tasmania against mainland lakes. PCA scores of axis 1 and axis 2 were used from both datasets from Fig. 4.2a and Fig. 4.2b respectively...... 58 Table 4.3 Wright and Burgin (2007) identified 5 genera, 4 species and 14 morpho-species of endemic chironomids in Tasmania. These are arranged in alphabetic order, below. We performed a literature search and show that all the 5 genera and 4 species were previously recorded and described in mainland Australia, therefore endemism is not supported at these taxonomic levels. The 14 morpho-species are not described and we are unable to assess the reliability of the endemism claim for these taxa. (*) Indicates undescribed morpho-species. ... 60 Table 4.4 CCA summary of the seven significant variables including canonical co-efficients and t- values of the environmental variables with the ordination axes including 33 lakes and 38 non- rare species ...... 62 Table 4.5 Partial CCAs of the seven significant (p ≤ 0.05) environmental variables alone and with the effects of other significant variables partialled out for 33 lakes with 38 non-rare species included...... 63 Table 4.6 List of chironomid taxa enumerated in this study along with data on distribution and environmental significance...... 69 Table 4.7 Performance of partial least squares (PLS) model for reconstructing mean February temperature of southeast Australia using 33 lakes and 38 non-rare chironomid species. The bold indicates the model of choice...... 74

xxv | P a g e

Table 5.1 Radiocarbon Dates for Welsby Lagoon. Data shown includes sample depth, laboratory, materials used for dating, 14C age and calibrated age (SHCal13; Hogg et al., 2013) for each of the samples. (*) indicates samples excluded in the final age depth model...... 83 Table 6.1 Summary of information available for the 16 lakes included in this study. The complete data-set is available in Chapter 3...... 113 Table 6.2 Stable isotope analyses (δ18O and δ2H) results of southeastern Australian modelled precipitation, lake water, Chironomus spp. and the fractionation factor between chironomus head capsules and lake water (α)...... 115 Table 6.3 Redundancy Analysis (RDA) of δ18O of Chironomus spp. head capsules and environmental variables shows that Aridity Index (AI) has the strongest explanatory power (53.7%) among the five variables that have a significant correlation (P < 0.05). It is not independent of MAT but it is impossible that it would be independent as temperature plays a key role in potential evapotranspiration (PET). AI explains significantly more variance than MAT and MAT is reduced to very low explanatory power when AI is removed...... 117 Table 6.4 Redundancy Analysis (RDA) of the fractionation factor (α) (δ18O between Chironomus spp. and lake water) and environmental variables shows that Total Nitrogen (TN) is the most independent variable that explains the largest variance (64.4%) among all seven variables that are significantly correlated (P < 0.05) for the stable oxygen isotope (δ18O) fractionation factor

(α[δ18O(Chironomus spp - water)]) (Table 6.2) between Chironomus spp. and host lake water ...... 118 Table 6.5 Redundancy Analysis (RDA) of δ2H of Chironomus spp. head capsules and environmental variables shows that for δ2H there are no independent variables. MAT explains the largest variance (55.2%) in stable hydrogen isotope (δ2H) data of Chironomus spp. but interacts with aridity index (AI), specific conductance (COND) and total nitrogen (TN)...... 121 Table 6.6 Redundancy Analysis (RDA) of the fractionation factor (α) (δ2H between Chironomus spp. and lake water) and environmental variables shows that specific conductance (COND) is the most independent variable that explains the largest amount of variance (68.4%) among all seven variables that are significantly correlated (P < 0.05) in the hydrogen isotope (δ2H) fractionation factor (α) between Chironomus spp. and host lake water ...... 123

xxvi | P a g e

List of equations

Equation 6.1.

Equation 6.2.

xxvii | P a g e

List of abbreviations, acronyms, symbols and units

14C Radiocarbon a.s.l Above sea leve AAR Amino acid racemisation ACR Antarctic cold reversal AI Aridity index ArcGIS Arc Geographic Information System AUS-INTIMATE Australasian - INTegration of Ice-core, MArine and Terrestrial rEcords AveBiasjack Average bias (jack-knifed) BOM Bureau of Meteorology BP Before Present (1950) cal. Calender year CCA Canonical correspondence analysis CGIAR-CSI Consortium for Spatial Information Chl a Chlorophyll a COND Specific conductance CONISS Constrained incremental sums of squares DCA Detrended correspondence analysis DCCA Detrended canonical correspondence analysis DEM Digital elevation model DIN Dissloved inorganic nitrogen DO Dissolved oxygen DRP Dissolved reactive Phosphorus EAC Eastern Australian current ENSO El Niño Southern Oscillation GC Gas chromatography GDGT Glycerol dibiphytanyl glycerol tetraethers GF/F Glass fibre filters GLMs Generalized linear models GNIP Global Network of Isotopes in Precipitation G-ruber Globigerinoides ruber HydroSHEDS Hydrological data and maps based on SHuttle Elevation Derivatives at multiple Scales

xxviii | P a g e

IAEA International Atomic Energy Agency INQUA INternational QUaternary Association INTIMATE INTegration of Ice-core, MArine and Terrestrial rEcords IPO Interdecadal Pacific Oscillation IRMS Isotope ratio mass spectrometry ka Thousand years KOH Potassium hydroxide LGM Last glacial maximum LM Lake Mackenzie MAE Mean annual evaporation MAP Mean annual precipitation MAT Mean annual temperature MaxBiasjack Maximum bias (jack-knifed) MCR Mutual climatic range MFT Mean February temperature NOx Nitrogen oxides NSW New South Wales ORP Oxidation reduction potential P:EP Precipitation : Evapotranspiration PCA Principal component analysis pCCA Partial canonical correspondence analysis PET Potential evapotranspiration PLS Partial least square PMIP Paleoclimate Modelling Intercomparison Project PSI Purdue stable isotope RDA Redundancy analysis RMSEP Root Mean Square Error of Prediction SAL Salinity SAM Southern Annular Mode SHAPE Southern Hemisphere Assessment of PaleoEnvironment SHCal13 Southern Hemisphere Calibration curve for radiocarbon dates updated to 2013 SLAP Standard Light Antarctic Precipitation SqRL Squared residual length

xxix | P a g e

SST Sea surface temperature STAB Subtropical anti-cyclonic belt TALDICE TALos Dome Ice CorE TC/EA Temperature Conversion Elemental Analyzer TDS Total dissloved solids TN Total nitrogen TP Total phosphorous TURB Turbidity USGS U.S. Geological Survey VIF Variance inflation factor VSMOW Standard Mean Ocean water WAFW World average freshwater ratio WA-PLS Weighted Average - Partial least square WASW World average seawater ratio WEL Welsby Lagoon WMO World Meteorological Organization YD Younger Dryas

α Fractionation factor for stable isotopes β Model coefficient in Partial Least Square models λ Eigen values in multivariate statistical analyses °C Degree Celcius cm Centimeters km Kilometers μg Micrograms μg/L Micrograms per liter μs/cm Microsiemens per centimeter m Meters mg/L Milligrams per litre mm Millimeters δ2H Stable hydrogen isotope δ18O Stable oxygen isotope

xxx | P a g e

Background 1 INTRODUCTION

1.1 Background

Climate change is one of the most serious issues facing humanity (Heltberg et al., 2009). Continuing debates around this issue stem partly from our limited insight into the differences between natural and anthropogenically driven climate variability (Rosenzweig et al., 2008). There is also limited understanding on the possible impacts of future environmental change on ecosystems (Walther et al., 2002). In addition, climate change can be regional and in order to understand global climate, knowledge on regional climate change is essential (WMO, 2015). In order to address these gaps in knowledge, we need to understand past climate variations and their effects on ecosystems and the environment (Caseldine et al. 2010; Schmidt 2010). Most of the instrumental measurements of climate are available for no longer than a few hundred years and in many countries, such as Australia, these records are often only available for a few decades. Because of their limited duration, these data are not ideal to be used to determine long term climate change or to understand abrupt climate events that have occurred in the geological past. Instead, an alternative understanding of the climate system can be achieved from the development and application of proxies that have the potential to provide long term palaeoclimate records. These climate proxies are often derived from natural archives (Zachos et al., 2001) and are extremely diverse, including but not limited to lake sediment, speleothems, tree rings, sand dunes, ice cores, and corals.

Proxy based paleoclimate reconstructions with adequate temporal and spatial resolution will help us gain a more detailed understanding of climate change and ecosystem responses as a whole (e.g. Alloway et al., 2007). Proxy based paleoclimate reconstructions are also very useful to validate results from climate models and to improve the capability of global climate models to better predict future changes (e.g. Riebeek, 2006). One of the key variables that has been targeted for validating global climate modelling is temperature (Riebeek, 2006). In this thesis, proxies for inferring past climate change are developed based on the species distribution and stable isotopic composition of non-biting midges (Diptera: Chironomidae) from southeastern Australia. The method is applied to late Quaternary sedimentary records from the subtropics to cool mid-latitudes to generate mean annual and summer temperature reconstructions.

The INTIMATE project (INTegration of Ice-core, Marine and Terrestrial rEcords) initiated by the International Union for Quaternary Science (INQUA) in 1995 aims to generate high quality paleoclimate records from various regions around the globe. It also aims to integrate these regional

1 | P a g e

Background records into climate stratigraphies, to gain a better understanding of the changes of the global climate system through time. A number of synthesis articles of the INTIMATE project have been published throughout the years of 1998 to 2014 (e.g. Blockley et al., 2012). The project has significantly advanced the knowledge of paleoclimate and paleoenvironmental changes especially in the North Atlantic where the project is primarily focussed. Compared to the Northern Hemisphere, the development of proxies and proxy based paleoclimate records in the Southern Hemisphere has received less attention and has only started to emerge in the last decade or so. This was partly associated with the establishment of the Australasian INTIMATE (AUS-INTIMATE) project in 2003 (Alloway et al., 2007; Turney et al., 2007; Reeves et al., 2013; Petherick et al., 2013). More recently, the Southern Hemisphere Assessment of PaleoEnvironment (SHAPE) project has been initiated to further improve the palaeoclimate records from the Southern Hemisphere. This thesis was initiated as part of a project titled ‘The last glacial maximum (LGM) conundrum and environmental responses of the Australian continent to altered climate states’ funded through an Australian Research Council (ARC) Discovery grant. That project was framed in the context of both AUS-INTIMATE and the SHAPE project. This thesis is intended to contribute to the following objectives of the SHAPE project:  provide robust interpretations of proxies  generate regional reconstructions of past environmental and climatic change and if possible;  integrate the results into a hemispheric-wide story for key time slices, highlighting the changes and testing hypotheses for their causes

Geographically, the distribution of proxy based records is rather sparse and non-uniform in Australia. However, studies (Reeves et al., 2013; Turney et al., 2007; Williams et al., 2009) suggested that Australia can provide high quality information on likely patterns and directions of the future changes in the Southern Hemisphere. Firstly, it covers a large climate range from the cool temperate zone to the tropics and incorporates broad humid, sub-humid, arid and semi-arid zones. Secondly, it is situated between two regions that are critical to global climate; the Indonesian Maritime Continent to the north, and Antarctica and the Southern Ocean to the south. The former is considered one of the engines of global climate, while the latter plays a major role in inter- hemispheric climate teleconnections. In addition, the Australian continent contains a diversity of Quaternary landforms and sediments in a wide range of environments which provides a great opportunity to create an integrated regional record of climate and environmental change (Turney et al., 2007). Proxies that can be used to generate long-term records and also to quantify the changes, from Australia are sparse. Therefore, further investigation of past climate changes in Australia is 2 | P a g e

Background urgently required (Turney et al., 2007) and paleoecological research is one of the key components (Birks, 2008).

Most of the commonly applied palaeoclimate proxies in Australia such as tree rings and borehole measurements are unable to provide records that extend as far back as the LGM. The LGM (traditionally 21-17 ka yr) is one of the key time intervals for climate reconstruction as this interval represents the most significantly different global climate in the recent geological past. Understanding long-term climate change that extends back to the LGM will help us separate anthropogenic impacts on climate from natural internal variability and radiative forcing of climate change. From these insights we can further investigate the long-term effect of climate change on ecosystems.

Australian proxy based reconstructions for the time interval from the LGM to the start of the present interglacial (c. ~11,000 years ago) include temperature reconstructions based on pollen (e.g. Fletcher and Thomas 2010), periglacial landforms (Galloway 1965), Amino Acid Racemization (AAR) from Emu eggs (Miller et al., 1997) and more recently branched glycerol dialkyl glycerol tetraether (GDGT) (Woltering et al., 2014). With the exception of pollen these proxies are relatively unconventional for the application of temperature inferences and these records have provided limited or contradicting estimates of temperature change in Australia. These results may reflect either regional variability, or the unreliability of one or more of the techniques, or both.

In summary, despite efforts through the AUS-INTIMATE project to improve palaeoclimate data sets, a critical gap remains in techniques that are available to determine past changes in temperature in mainland Australia. Further developing and refining both conventional and novel proxy based methods that can reconstruct the LGM temperatures is required. Records derived from applying these methods will provide valuable information on the connection between Australian terrestrial temperature changes and adjacent oceans and Polar Regions during the LGM. Such data will advance our knowledge on hemispheric and global climate teleconnections.

Transfer function technique has been widely applied for quantitative reconstructions of past climate changes using a number of biological proxies (Birks, 1998). Proxies that have been applied in Australia include mainly diatoms (Tibby, 2004), pollen (Cook and van der Kaars, 2006; Fletcher and Thomas, 2010) and chironomids (Rees et al., 2008). Among these, diatoms are primarily used for salinity and other in-lake variable reconstructions. Pollen and chironomids are essentially

3 | P a g e

Background commonly applied climate indicators but with respective to the transfer function technique, more successful applications are demonstrated for the use of chironomids.

A number of studies show that chironomids can also be used as a proxy for nutrient status (Woodward and Shulmeister, 2006), pH (Rees and Cwynar, 2010b), hypoliminia conditions (Broderson et al., 2001) and salinity (Zhang et al., 2007) inferences. However, the most widely and traditionally applied chironomid based transfer functions are for palaeo-temperature reconstructions (Walker 1987; Olander et al., 1999; Lotter et al., 1999). A number of long term, high resolution temperature records have been derived using the transfer function method from high latitudes of the Northern Hemisphere (e.g. Brooks and Birks, 2000; Larocque-Tobler and Finsinger, 2008; Larocque-Tobler et al., 2010). In the last decade, chironomid based transfer functions have been developed and successfully applied from New Zealand (Woodward and Shulmeister, 2006, 2007; Dieffenbacher-Krall et al., 2007), Tasmania (Rees et al., 2008; Rees and Cwynar, 2010a) and South America (Massaferro et al., 2014). All of these applications have provided sensible quantitative temperature reconstructions. This thesis will present the development of the first subfossil chironomid based transfer function for mainland Australia and trial its application to an LGM record.

More recently, Wooller et al. (2004, 2008) and Verbruggen et al. (2010, 2011) showed that oxygen isotopes measurements on chironomid head capsules are a promising proxy that can be used to independently reconstruct past temperatures. The method of using stable isotope on chironomid head capsules has not been applied in the Australasian region and has limited investigation (Mayr et al., 2014) in the Southern Hemisphere. Many aspects associated with the use and interpretations of this new proxy are insufficiently understood. This thesis will include the first investigation on the potential of applying stable oxygen and hydrogen isotopes in subfossil chironomids from Australia as a proxy for past climate and environmental changes.

The paleoecological record generated from this study from the Australian continent will (hopefully) provide critical quantitative data concerning climate change in the region during the late Quaternary. The proxy based method developed in this thesis will have wide future applications, for example, long-term quantitative, high resolution climate reconstructions will be possible from different climate zones of eastern Australia.

4 | P a g e

Research objectives 1.2 Research objectives

The general aim of this thesis is to understand the past climate system and changes in eastern Australia with a focus on the LGM and late glacial by developing and applying proxy (subfossil chironomid) based methods for past climate reconstructions.

Objective 1: To provide an overview of the physical and chemical properties of water bodies in eastern Australia, extending from the tropics to Tasmania to provide the background data for the chironomid transfer function and stable isotope investigations.  To acquire and analyse water chemistry data of samples collected from 45 water bodies  To acquire and analyse remote sensed data of the catchment land-use and climate variables of the water bodies in ArcGIS  To examine the differences and similarities of the responses to stressors in the catchments between natural lake systems and reservoirs  To examine the inter-relationship of the catchment land-use, climate and environmental variables within eastern Australia freshwater systems  To summarize the current status and identify the major environmental controls on the Australian freshwater bodies

Objective 2: To create a transfer function(s) based on the chironomid assemblages across southeastern Australia that can be used to reconstruct past climate or other environmental proxies.  To extract chironomids from the water – sediment surface samples  To perform taxonomic analyses on the extracted chironomids  To determine the key environmental and climatic controls on chironomid species distribution  To develop transfer function(s) that model the relationship between chironomid species assemblages using the key variable(s) determined  To examine the effects of including artificial waterways for the temperature transfer function development  To examine the previously claimed endemic taxa effects of including Tasmanian lakes

Objective 3: To apply method(s) for down-core analyses of a selected paleosite(s) with records that extend back to the last glacial maximum (LGM) for climate reconstruction

5 | P a g e

Research objectives  To extract chironomid head capsules from the targeted section of a well-dated sediment core of a selected site  To perform taxonomic analyses on extracted subfossil chironomids throughout the sediment core  To perform a reconstruction applying the transfer function  To perform the reconstruction diagnostics and interpret the results while consulting other climatic and environmental reconstructions from adjacent sites  To infer changes in the climate system covering the LGM of eastern Australia based on the results obtained

Objective 4: To explore the opportunity of using the δ18O of subfossil chironomid head capsules for reconstructing past changes.  To obtain sufficient chironomid head capsules (minimum dry weight of 100 µg) from a selected genus of modern water – sediment lake samples for stable isotope analyses  To analyse the stable oxygen and hydrogen isotopes from chironomid head capsules samples and from lake water samples  To examine the relationships of the stable isotopes of chironomid head capsules against stable isotopes of lake water and precipitation, and other significant climatic and environmental variables  To calculate the fractionation factor of stable isotopes of chironomid head capsules and lake water  To examine and understand the environmental controls on the chironomid stable isotope fractionation  To determine if stable isotopic composition of chironomid head capsules has the potential to be used as a proxy for paleoclimate or other paleoenvironmental reconstructions in Australia

1.3 Overview of research methodology

Australia is generally a dry continent and therefore, there are a limited number of water bodies to choose from. The aim was to sample as many natural lakes as possible from the southeast Australian mainland. However, artificial lakes and reservoirs are also included due to the scarcity of permanent water bodies in some critical climate and elevation spaces. The effect of including artificial water-bodies in the dataset is investigated in Chapter 3 and 4, respectively. An overview of the flow and framework of this project and a project timeline is shown in Figure 1.1.

6 | P a g e

Research objectives

Data screening Statistical Field Work 1 Collected water analyses – To understand the (10 Lakes) samples to be modern limnology, climate analysed for stable and land-use relationships isotopes and chemistry Obj 1. Develop Obj 2. transfer

Field Work 2a Chironomid function(s) If Primary applicable (10 Lakes) Field Data taxonomy Site analyses Selection Down-core application and initial to selected site for Obj 3. information Field Work 2b Collected reconstruction Chironomids collection (14 Lakes) sediment samples to be processed in extracted and If Develop GPEM labs pre-treated applicable relationships based on chironomid Obj 4. stable isotopic composition Chironomid stable isotopes Field Work 3 All other site (Paleo-site) analyses Statistical analyses related data for chironomid

collected stable oxygen isotopes data

Figure 1.1 Overview of the framework and project flow 7 | P a g e

Research objectives Project timeline

8 | P a g e

Thesis structure and synopsis

1.4 Thesis structure and synopsis

This thesis uses ‘Thesis by publication’ format. Fig. 1.2 shows the overall thesis structure and visualizes how each of the three objectives is addressed in relevant publications or submitted manuscripts.

At the time of thesis submission, Chapter 3 is published in Marine and Freshwater Research (Chang et al., 2014); Chapter 4 is published in Palaeogeography, Paleoclimatology, Paleoecology (Chang et al., 2015a); Chapter 5 is published in Quaternary Science Reviews (Chang et al., 2015b) and Chapter 6 is submitted for publication in Limnology and Oceanography (Chang et al., submitted).

As opposed to a traditional thesis, each chapter (paper) contains a relevant literature review, detailed methodology, results, discussion and a conclusion sections. The final chapter comprises conclusions from this thesis drawing together the overall outcomes and contributions to the research context. Recommendations for future work are also discussed in this chapter. In addition, appendices incorporate supporting materials (Appendix 1) that are not presented in publications, and photographs of modern southeastern Australian chironomid head capsules (Appendix 2) and for the fossil taxa from the Welsby Lagoon palaeo-record (Appendix 3). An identification key based on the photographic materials will be submitted for publication as an electronic resource in the near future along with work from other colleagues. Published journal articles during the candidature are incorporated as Appendix 4.

9 | P a g e

Thesis structure and synopsis

Chapter 1 and 2: Introduction and background

Objective 1

Objective 2

Chapter 3: A snapshot of the Chapter 4: A chironomid limnology of eastern Australian based transfer function for water bodies spanning the tropics reconstructing summer to Tasmania: the land-use, temperatures in south climate, limnology nexus eastern Australia

Objective 3 Objective 4

Chapter 5: A chironomid-inferred summer temperature reconstruction from subtropical Australia during the last glacial maximum (LGM) and the last deglaciation Chapter 6: Can stable oxygen and hydrogen isotopes from Australian subfossil chironomid head capsules be used as proxies for past environmental change?

Chapter 7: Conclusions and future work

Figure 1.2 Overview of the thesis structure 10 | P a g e

The Biology of Chironomid

2 THE BIOLOGY OF CHIRONOMIDS

Chironomids (Chironomidae, known commonly as non-biting midges) are a family of two-wing (Insecta: Diptera) (Cranston and Martin, 1989). More than 5,000 species have been identified world-wide, however, it is estimated that up to 15,000 species may exist in total (Cranston 1995). Chironomid larvae range from 2-30 mm in length and most are pale yellow, green or red in colour. The larvae are worm-like, having three thoracic segments and a nine-segmented abdomen (Fig. 2.1) (Brooks et al., 2007). Most chironomid larvae are detritivores, i.e. obtain nutrients from grazing bacteria and algae (Cranston and Martin, 1989).

Figure 2.1 Key morphological characteristics of chironomid larvae (Brooks et al., 2007, Page 4)

The chironomid eggs are laid in a hygrophilous jelly mass dropped into the water or in a jelly- string, attached by one end to the substrate. The hatching rate and survivability are temperature dependant (Cranston and Martin, 1989). Most of the larvae are tube-dwelling, apart from subfamily Tanypodniiae, which are free-living. The silken tube is constructed from silk-spinning organs located in the ventro-mentum and protects larvae from predators, and also performs a respiratory function.

There are four larval stages of chironomids (from 1st instar to 4th instar) (Fig. 2.2). The development time is influenced by the combination of temperature and food availability (Johannsson, 1980). Most of the temperate species are multivoltine, univoltine or bivoltine (voltinism: used in biology to indicate the number of broods or generations of an organism in a year) (Tokeshi, 1995), for example, large species living in deep lakes may take more than one year to develop, whereas small species living in warm, shallow water bodies may complete 3-4 generations in one year (Tokeshi, 1995). In southeast Australia, apart from the high mountain lakes, it is reasonable to assume most

11 | P a g e

The Biology of Chironomid species may complete 2-3 generations in one year from lowland lakes. The pupal stage usually lasts 3-4 days and the timing of emergence is dependent on water temperature and light intensity (Kureck, 1979). Adults of most species live for just a few days, at most for few weeks (Oliver, 1971).

Figure 2.2 Chironomid Life Cycle (Walker, 1987, Page 852)

Figure 2.3 A fully developed chironomid head capsule (Chironomus sp.) from Lake Albina, Mt Kosciuszko, New South Wales, Australia (photo is taken by J. Chang at the Physical Geography laboratory, School of Geography, Planning and Environmental Management, University of Queensland) 12 | P a g e

The Biology of Chironomid

Chironomid head capsules (Fig. 2.3) are derived from dead larvae and also largely from head capsule moults between larval instars. They are usually the most abundant remains in lake sediments, in most freshwater situations, from the tropics to the arctic, as the chironomid larvae occupy a huge range of aquatic habitats (Cranston and Martin, 1989). The head capsules of chironomid larvae are made of chitin (Fig. 2.4), usually fully developed, globular, and pigmented dark brown or black. The subfamily has a more elongate and un-pigmented head capsule. The head capsule of all chironomids cannot be retracted into the thorax (Brooks et al., 2007). Chironomid head capsule samples are usually dominated by third and fourth instar capsules as the earlier instars (first and second) are less durable.

Figure 2.4 Chitin molecule structure

Chironomid larvae have several qualities that make them useful as a proxy in environmental and climatic palaeo-studies. They are stenotopic, which means different elements of chironomid assemblage respond in a distinctive way to particular environmental perturbations (Brooks, 2003). They are also abundant and ubiquitous, therefore, there is usually a large number of them that can be found in practically all permanent or semi-permanent freshwater bodies. Chironomids are also rich in species that make the assemblages sensitive to environmental changes (Brooks, 2003). The response of chironomids to any change in conditions is local and fast due to their short life cycle. They are also identifiable as the head capsules are usually well-preserved, although, it is still challenging to identify most of them beyond genus or species-group level.

More recently, Wooller et al (2004, 2008) and Verbruggen et al. (2011) have demonstrated that the chitin contained in the head capsules can be analysed for stable isotopes (δ18O and δ2H). Chitin is a long-chain polymer of an N-acetylglucosamine, a derivative of glucose (Fig. 2.4), and is the main component of the exoskeletons of chironomids. This has provided new opportunities for paleoecologists to investigate the novel methods to be used in paleoclimate and paleoenvironmental reconstructions. 13 | P a g e

The Biology of Chironomid

Chironomid taxa have been used as a proxy in paleoclimate and paleoenvironmental studies since 1980s (Boubee, 1983; Hofmann, 1986; Walker, 1987; Walker and Mathewes, 1989). Using chironomid taxa as quantitative climate indicators were derived from various parts of the world as early as 1990s (e.g. Walker et al., 1991a, b; Wilson et al., 1993; Levesque et al., 1993; Walker et al., 1995; Heinrichs et al., 1997; Olander et al 1997; Walker et al., 1997), but all focused on Northern Hemisphere. The first chironomid based transfer function from the Southern Hemisphere was developed by Woodward and Shulmeister (2006) from New Zealand. A detailed literature review of the development and uses of transfer functions is presented in Chapter 4 (Section 4.1). The current available modern calibration data sets and transfer functions is also summarized in Appendix 1.

2.1 The principles of applying the transfer function and stable isotope methods in palaeoclimate reconstructions

2.1.1 The transfer function method

As discussed above, chironomid-based transfer function, as a technique to quantify past climate and environmental variables has been developed and applied as long as thirty to forty years. The basic and simplified idea of the technique following Sachs et al. (1977) is: Assume that summer lake temperature (Ts) (which is close to the mean air temperature) will be a function of the chironomid taxa abundance. If the abundances measured at a particular point were represented by the probability series (for instance, there are 30 different taxa in the modern training set) p1, p2, p3…p30 (where Ʃp = 1), here ‘p’ represents the probability of occurrence of a taxon and ‘a’ represents the coefficient factor, then we can express the equation as follows:

Ts = a1p1 + a2p2 + a3p3 + . . . + a30p30 + a31p1p1 + a32 p1p2 + . . . + a60p1p60 + a61p1p2p1 + a62p1p2p2 ...

If we used all first and second-order terms, we would have to develop 496 parameters in the above equation. So to simplify, instead of using all of the chironomid taxa abundance data, factor analysis was used to group taxa which behave in a similar fashion together into a single factor. The minimum number of factors to represent an acceptable amount of variance will be used. Then the expression for temperature is simplified and becomes:

14 | P a g e

The Biology of Chironomid Ts = a1f1 + a2f2 + a3f3 + . . . + a6f6 + a7f1f1 + a8f1f2 + . . . +a27f6f6 + a28. (for instance, in this example it comes down to 6 factors) Values for the coefficient factors (a1 to a28) were then found by multiple regression (usually by applying a Monte Carlo simulation).

After the modern calibration set is established, fossilised chironomids found at different levels of a dated core can be examined. Different taxa of chironomids from each level should be counted and then the assemblages converted into factors determined by the modern calibration data set. Then we apply the factor loading scores to the transfer function and calculate the summer temperatures at different levels at the site of the dated core.

2.1.2 The stable isotope method

The basic idea for using the δ18O of subfossil chironomid head capsules to infer mean annual air temperature (MAAT) changes in the past is that we assume the δ18O values analysed on chitin from the chironomid head capsules gathered from the surface sediments have a strong linear relationship with the MAAT. This premise is based on the observed strong correlation between δ18O of chironomids and δ18O of lake water, δ18O of lake water and δ18O of local precipitation, as well as δ18O of precipitation and MAAT. A similar principle is applied to the hydrogen isotope (δ2H).

Verbruggen et al’s (2011b) study from northern Europe has shown strong linear relationships between MAAT and δ18O of Precipitation (r2=0.94), similar to δ18O of chironomids and δ18O of lake water (r2=0.9), however, the linear correlation between δ18O of precipitation and δ18O of lake water is weaker (r2=0.64). This is likely due to the δ18O of the lake water reflects both the source of precipitation and evaporative processes in the lake. In addition, the δ18O of precipitation can also reflect changes in source and seasonality among other variations as well as MAAT. In this study, these correlations will be tested from southeastern Australian lakes and we should expect complexities among these relationships. Same principle is applied to the hydrogen isotope (δ2H).

For the application of the stable isotope method to a dated core, firstly we convert the measured δ18O/ δ2H values from chironomid fossils to temperature (analogous to modern δ18O/ δ2H chironomid inferred temperature values) using the model developed based on δ18O/ δ2H values measured on modern chironomids from the lake surface sediment. Secondly, it is assumed that the MAAT reconstructed using this method represent and quantify the major temperature variations and fluctuations in the past.

15 | P a g e

Land-use, climate and limnology nexus of eastern Australia

3 A SNAPSHOT OF THE LIMNOLOGY OF EASTERN AUSTRALIAN WATER BODIES SPANNING THE TROPICS TO TASMANIA: THE LAND-USE, CLIMATE, LIMNOLOGY NEXUS

Chapter 3 is published in Marine and Freshwater Research (Chang et al., 2014)

3.1 Summary

The present study is the first comprehensive overview of the physical and chemical properties of lakes in eastern Australia, extending from the tropics to Tasmania. The results showed that climate is the dominant control on the lakes. The results suggested that high nutrient concentrations in lakes in drier areas may be the result of natural processes of concentrating nutrients from high evaporative flux in eastern Australia, rather than due to human activities.

Highlights:  It provides an insight into how lake water chemistry, climate and human impact factors interact and correlated with each other from eastern Australian lakes in the modern age. It is critical to understand these correlations and connections before we try to use any proxies (not limited to chironomids) from lake sediment to reconstruct past lake conditions, environment or climate of the region or to use these data to predict any future changes.  This study highlights that lake responses to stresses can vary in predictable ways across the Australian landscape, and that management and restoration would benefit from considering factors such as geology, climate, etc..  This study will help to interpret the transfer function results, the transfer function based paleo-reconstructions and stable isotope results from modern lakes

16 | P a g e

Land-use, climate and limnology nexus of eastern Australia 3.2 Abstract

The present study investigates 45 natural and artificial water bodies extending across the whole of eastern Australia from the tropics to Tasmania. A broad variety of physical and chemical, land-use and climatic parameters were measured. Reservoirs and other artificial water bodies responded to stressors in their catchments in a similar fashion to natural lakes, but tended to be less nutrient rich, possibly because of shorter residence times and active management. Salinity and pH were strongly correlated in the dataset. Bedrock had a strong influence on pH in freshwater lakes, whereas all highly saline lakes were alkaline, irrespective of bedrock. High concentrations of anions in saline lakes precluded the existence of acid conditions by binding available hydrogen ions. Almost all lakes fell on salinity axes that indicated marine origin for their salts. An assessment of the total nitrogen to total phosphorus molar ratios from the lakes in the present dataset indicated that productivity in Australian lakes could be limited by both nitrogen and phosphorus. Future research using macro-nutrient enrichment experiments should be pursued to confirm this preliminary observation. There was a strong positive correlation between regional aridity and lake eutrophication. This is typical of semi-arid and seasonally arid environments and reflects the concentration of nutrients owing to evaporative flux in shallow basins with high residence times.

Key words: eutrophication, evapotranspiration, freshwater, lakes, reservoirs, lithology.

17 | P a g e

Land-use, climate and limnology nexus of eastern Australia 3.3 Introduction

Apart from a review of Australian lake geomorphology by Timms (1992), which focused primarily on the formation of lakes, and broad overviews by Brookes and Hamilton (2009) and Bridgman and Timms (2012) of all Australian lakes, which contained little detail, there is no summary of physical and chemical data from mainland eastern Australian lakes. The aim of the present paper was to provide baseline physical and chemical, land-use and climatic data for eastern Australian lakes, and to provide a preliminary evaluation of the primary controls on lake chemistry in eastern Australia. The paper emerged from a study designed to determine whether subfossil chironomid head capsules could be used as a basis for reconstructing temperature and nutrient status in Australian water bodies. The study questions related to whether and how climate, geology and land use affect lake water chemistry in eastern Australia. Given that there is little summary work in this area, the present basic descriptive survey was expected to generate findings that can lead to testable hypotheses about controls on lake water chemistry.

Globally, freshwater lakes are concentrated at high latitudes. Canada has the largest number of lakes, with an estimated 2 million water bodies, of which 31 752 are larger than 3 km2 (Renwick 2009). Northern Europe also has abundant lakes. In the Australasian region, New Zealand has 3820 lakes and Tasmania has more than 4000 lakes (Brookes and Hamilton 2009). In comparison, the Australian mainland has very few natural permanent freshwater bodies (Bridgman and Timms 2012). This is despite the fact that an area of ~1.5 million km2 (roughly six times the size of New Zealand) is humid (precipitation ≥ evaporation). Therefore, the lack of permanent freshwater lakes is not simply explained by a lack of precipitation. Geomorphology and, in particular, the limited extent of past glaciations in Australia appear to be the primary reason for this paucity of lakes (Bridgman and Timms 2012).

3.3.1 Previous water-chemistry studies of Australian lakes

Previous studies have recognised eight types of lakes in Australia, on the basis of their mode of formation (Timms 1992). These are glacial lakes, volcanic crater lakes, coastal lagoons, coastal dune lakes, tectonic lakes, fluvial lakes, solution lakes (sink-hole lakes) and lakes formed by landslides. We chose not to review tectonic lakes because most of the lakes in this category are not permanent (e.g. Lake George) and landslide-blocked lakes are rare, and we added an upland-lagoon category to cover shallow lakes on the tablelands (Fig. 3.1).

18 | P a g e

Land-use, climate and limnology nexus of eastern Australia

Figure 3.1 Map of eastern Australia, with the 45 study sites identified; the numbers correspond to the lake names and numbers in Table 3.1. 19 | P a g e

Land-use, climate and limnology nexus of eastern Australia Glacial lakes in mainland Australia are limited to alpine areas of the Australian Alps (Fig. 3.1). Williams et al. (1970) demonstrated that the Kosciuszko lakes (Fig. 3.1) were ‘extremely fresh’ and that salinity, and nitrate and phosphate concentrations were very low. Green (2012) summarised the seasonal chemical characteristics of five Kosciuszko lakes that are ice-covered in winter. He suggested that the salts and nutrients of the five Kosciuszko lakes were mainly derived from bedrock erosion and Aeolian inputs and nutrient deposition from precipitation was low. Most lakes in the Tasmanian highlands and central plateau (Fig. 3.1) are of glacial origin (Jennings and Banks 1958). The majority of these lakes have ionic composition consistent with aerosol sources deposited in the lake through both dry and wet deposition and originally relate to a seawater source (Buckney and Tyler 1973; Vanhoutte et al. 2006). The major water-chemistry gradients are conductivity, pH and gilvin (i.e. Chloraphyll) (Vanhoutte et al. 2006) and these relate primarily to bedrock geology and climate.

The second major lake type is the volcanic crater lake. Two regions in Australia have significant concentrations of crater lakes, namely, the Atherton Tablelands of northern Queensland and the basalt province of western Victoria (Fig. 3.1). Some isolated crater lakes (e.g. Coalstoun Lakes in southeastern Queensland) occur outside these main fields. In northern tropical Queensland, the paleoecology of these lakes is rather better studied than the modern lake chemistry, with extensive work on human settlement history, palaeoclimate and palaeoenvironmental reconstructions (e.g.Dimitriadis and Cranston 2001; Turney et al. 2001; Haberle 2005; Kershaw et al. 2007). In western Victoria and adjacent regions of South Australia, the crater lakes are more saline. These lakes were heavily investigated and the results of these studies were summarised in De Deckker (1983) and De Deckker and Williams (1988). These studies showed that sodium is the most common cation, nitrogen is the limiting plant nutrient and that the lakes are turbid because of high algal concentrations.

Coastal lagoons are widely distributed around the eastern and southeastern coasts of Australia. The literature on lake water chemistry of Australian coastal lagoons is scarce, or at least not fully in the public domain, largely comprising government monitoring data and reports. Scanes et al. (2007) evaluated 22 coastal New South Wales (NSW) lagoons and concluded that controls on the lagoons were highly variable and that the lagoons were minimally affected by human activity.

Dune lakes are the most common lake type in eastern Australia. The greatest concentration of dune lakes occurs on Fraser Island (Fig. 3.1), which has over 100 freshwater bodies and the limnology of these lakes has been heavily investigated since the 1960s (Bayly 1964; Bayly et al. 1975; Bowling 20 | P a g e

Land-use, climate and limnology nexus of eastern Australia 1988; Hembrow and Taffs, 2012). They occur as either perched or window lakes. Overall, these lakes are oligotrophic. Dune lakes in Gippsland, western Victoria and northeastern NSW were investigated by Timms (1973, 1977, 1982).

We present upland ‘lagoons’ of the NSW tablelands (Fig. 3.1) and particularly northern NSW tablelands as a distinct lake category (e.g. Little Llangothlin Lagoon, Mother of Ducks). There is no consensus on the origin of these shallow (< 3 m deep) lakes and they may have formed by sapping or deflation, although some are clearly created by drainage impoundment (Bell et al. 2008). As with other lakes, sodium is the dominant cation and the ionic concentration relates to the precipitation : evapotranspiration (P : EP) ratio and the season of precipitation (Briggs 1980).

In lowland Australia, there are numerous lakes that are formed by meander cut-offs (Constantine and Dunne 2008) and also intermittent lakes that form as part of a chain-of-ponds style drainage. We do not consider these types of lakes. In limestone areas, e.g. parts of northwestern and southern Tasmania, sink-hole lakes are occasionally present, such as e.g. Lake Chisholm, Tasmania (Bowling and Tyler 1988). Lakes formed by landslides, such as e.g. Lake Tali Karng (Timms 1974), are not covered in our dataset.

3.3.2 Artificial water bodies

Because natural water bodies do not occur in all regions, the present study included several reservoirs and other artificial water bodies. The construction of Australian reservoirs began in the early 1880s. Reservoirs are concentrated in the eastern and southeastern coastal regions of Australia where the population density is highest. Hydro-electric schemes are the other major source of artificial lakes. Tasmania has ~60 major reservoirs across the state for this purpose and other reservoirs are located in southeastern Australia as part of the Snowy Hydro Scheme. A few lakes have been created for recreational purposes, notably for fishing, and two of these lakes were included in the dataset.

Many studies have addressed the water quality issues that are important in reservoirs, including turbidity problems (Chanson and James 1999), problems with eutrophication and floating macrophytes (Harris 2001; Smith 2003; Davis and Koop 2006), deep water anoxia (Beutel and Horne 1999; Townsend 1999) and the presence of toxic cyanobacteria (Baker and Humpage 1994; Jones and Poplawski 1998). Eutrophication from nutrient rich rivers (e.g. Murray River) may

21 | P a g e

Land-use, climate and limnology nexus of eastern Australia also be a risk (Baker and Humpage 1994). The biggest difference between reservoirs and natural lakes is the shorter residence time for water in reservoirs (Søballe and Kimmel 1987).

22 | P a g e

Land-use, climate and limnology nexus of eastern Australia Table 3.1 Summarised information of the 45 water bodies sampled

23 | P a g e

Land-use, climate and limnology nexus of eastern Australia

3.4 Materials and methods

3.4.1 Study sites

The present dataset comprises 45 lakes and artificial water bodies located in the eastern and southeastern Australia (Table 3.1, Fig. 3.1). The transect covers a distance of 3700 km along the eastern coast of Australia, from Crater Lakes National Park, Queensland, to Mount Field National Park, Tasmania (17.24°S to 42.67°S, 140.18°E to 153.26°E; Table 3.1, Fig. 3.1). The climate ranges from monsoonal tropical in the north, to cool temperate in the south and, hence, there are large temperature (20.4°C for mean annual temperature) and precipitation (2476 mm for mean annual rainfall) gradients in the datasets. Elevation of the sites ranges from sea level to 2048 m above mean sea level (asl; Table 3.1). The dataset comprised 29 natural lakes and 16 artificial water bodies.

Climate

The climate system of eastern Australia is governed by the major global zonal circulation systems, from the monsoon tropics to the southern hemisphere westerly belt (Sturman and Tapper 2006). Climate regions can be summarised as follows: 1. The quasi-monsoonal (Gentilli 1971) north. This region, which extends from coastal northern Queensland to northern NSW, is dominated by southeastern trade winds in the winter and by the northern Australian monsoon in the summer. The climate in the vicinity of the coast is significantly modified by persistent moisture-laden onshore winds and orographic rainfall as these winds rise over the eastern highlands (Fig. 3.1). 2. Coastal southern NSW and Victoria are dominated by the subtropical anti-cyclonic belt (STAB) in summer and by a westerly flow in the winter. These areas are dry in summer, wet in winter, but the intermittent influence of easterlies and their interaction with topography modifies the coastal and near coastal climates. 3. A true Mediterranean climate exists in coastal South Australia and western Victoria. Rainfall predominantly occurs during winter in this area, under the influence of the westerlies, and there is sparse rainfall during summer under the STAB; however, this region is typically unmodified by eastern coast lows or other moisture sources. 4. Tasmania is the only dominantly all year westerly climate, although here again, there is a weak summer dry period in eastern and northern districts.

24 | P a g e

Land-use, climate and limnology nexus of eastern Australia This overall pattern is modified by oscillatory climate phenomena embedded in the circulation systems. These range from short duration Madden–Julian cycles (~60 days) to long duration cycles such as the interdecadal Pacific oscillation (IPO). The main oscillatory systems that are effective at climate scales are El Niño-Southern Oscillation (ENSO) (18 months to 7 years), the Indian Ocean dipole, which operates on timescales similar to those of ENSO, the IPO (15–30 years and 50–70 years), and the southern annular mode (SAM), which is annual and inter-annual.

Altitude is the other major modifier of climate. The Snowy Mountains region in the southeast (Fig.3.1) is characterised by cool to cold weather all year around and snow falls in winter. The record lowest temperature in Australia of –23°C was recorded at Charlotte Pass (http://www.bom.gov.au/climate/extreme/records.shtml, accessed 18 August 2013). At similar latitudes at sea level, the climate is warm temperate.

Geology Australia is an old cratonic plate, with mainly granitoid rocks and sedimentary rocks derived from them. Locally hot-spot volcanism has generated basalts.

Granite. Granite batholiths are common along the high country in eastern Australia. Granites dominate the high country from Victoria to northern Queensland. Granite rocks are high in silica (>70%), and, consequently, lakes on these rocks tend to be acidic.

Basalt. Hotspot volcanism has occurred locally from northern Queensland to western Victoria over the past 30 million years. The youngest volcanic fields occur in western Victoria and the Atherton Tableland of northern Queensland and are Holocene in age. Basaltic rocks are mafic and – 2+ lakes on basalts tend to have high ionic values, including bicarbonate (HCO3 ) and calcium (Ca ). They tend to be alkaline.

Meta-sedimentary and metamorphic rocks. Limnologically, Tasmania is bisected by a line that runs northwest by southeast, along the 146th meridian, called Tyler’s Line. Western Tasmania consists of mainly Precambrian siliceous meta-sedimentary and metamorphic rocks, including quartzite, schist, phyllite and limestones. Lakes west of Tyler’s Line are typically acidic with pH ranges from 4.5 to 5, except for lakes on limestone. Central and eastern Tasmania is underlain by Permian mudstones and Triassic siltstones that were partly metamorphosed by Jurassic doleritic intrusions. These lakes are less acidic (pH of > 6). Sedimentary rocks occur on the mainland also (mainly sandstones) and these tend to be neutral to acidic.

25 | P a g e

Land-use, climate and limnology nexus of eastern Australia Vegetation In eastern Australia, there are four broad structural types of vegetation. These are rainforest (notophyllous vine thicket), wet and dry sclerophyll forest or woodland, shrubland and heathland and grassland. In the highest country, subalpine herbland occurs. There are many subcategories of these structural vegetation types.

Rainforest. Rainforest is confined to areas of high annual rainfall close to the eastern coast and along the dividing range. Rainforests prefer fire protected sites and fertile locations (basalt areas and river plains). The rainforests are dominated by hardwood taxa with emergent conifers.

Sclerophyll woodland. In areas with poorer soil and relatively lower rainfall, Eucalyptus woodland becomes dominant. The understorey contains small broadleaved trees, vines, ferns and shrubs. On the drier eastern side of Tasmania and inland from the main divide on the Mainland, dry sclerophyll eucalyptus forests and woodland dominate. They contain an understorey of small trees including Casuarina, and Acacia.

Shrubland and heathland. Shrublands, heaths and sandplain vegetation characterise the coastal regions, especially on sandy soils. The dominant species are mostly from the families Ericaceae, Proteaceae and Myrtaceae. Melaleuca spp. are dominant in swampy areas.

Grassland. The lowland country of Victoria (generally below 700 m) with 400–1000 mm annual rainfall is dominated by grassland ecosystems on relatively nutrient-rich loam, clay-loam and alluvial soils. Most of the grassland has been removed or altered for agriculture and urban development. The ground layer is dominated by perennial, tussock-forming grasses, with some rhizomatous or stoloniferous species. In parts of the high country of Victoria (subalpine areas) and on other montane areas, tussock grasslands occur. In western Tasmania, communities of the Cyperaceae, the button grass Gymnoschoenus sphaerocephalus, are extensive as a consequence of fire regimes (Harris and Kitchener 2005). In alpine areas, tall alpine-herbfield and heathland communities dominate (Costin et al. 2000).

Forty-five lakes and reservoirs were chosen to span as much as possible the variety of lake types in eastern Australia (Fig. 3.1, Table 3.1). These lakes span the latitude and elevation of eastern Australia and gradients of climate, water depth, trophic status and agricultural activity (Table 3.1 and Table 3.2 and 3.3).

26 | P a g e

Land-use, climate and limnology nexus of eastern Australia 3.4.2 Sample collection and laboratory analyses

Lakes and reservoirs were sampled in eastern Australia during the summers (January to February) of 2011–2012 and 2012–2013 (Fig. 3.1, Table 3.1, Table 3.2). Water for chemical analyses was sampled from the lake centre at ~30 cm below the water surface, into four pre-washed and labelled polyethylene bottles (two 1000 mL, one 250 mL and one 125 mL). Untreated water samples were + + 2+ 2+ – - 2– collected for the analysis of major ions (Na , K , Ca , Mg , HCO3 , Cl , SO4 ), total nitrogen and total phosphorus (TN and TP) analysis. These samples were kept frozen until analysis. A 1000-mL water sample was filtered for Chlorophyll a (Chl a) on a 4.7-cm-diameter GF/F filter (0.45-µm pore size). The Chl a filter was wrapped in foil and kept frozen for subsequent analysis. The 125-mL water sample was filtered using a syringe and a syringe-filter Whatman (0.45-µm pore size, Supor Membrane) and the filtered water was frozen in the bottle for later analysis for dissolved reactive P

(DRP) and reactive N (NOX). Total dissolve solids (TDS), pH, specific conductance (COND) and turbidity (TURB) were also obtained from water-chemistry analysis, which was carried out by the Forensic and Scientific Services in Brisbane, Queensland. In the field, water temperature, oxidation reduction potential (ORP), pH, dissolved oxygen (DO), COND, TDS, salinity (SAL) and TURB were recorded using an Aquaread multiparameter meter. The meter also recorded latitude, longitude and elevation. Lake depths at the sampling points were measured using a Speedtech portable depth sounder.

3.4.3 Spatial data

WorldClim for lake climatic data: Climate variables were obtained using the combination of the WorldClim program (available from http://www.worldclim.org/bioclim, accessed 20 August 2013) and ArcGIS 10.1. For the present study, mean annual temperature (MAT) was selected because it represents the regional annual average temperature. The aridity index (AI) is defined as AI = MAP/MAE, where MAP = mean annual precipitation and MAE = mean annual evaporation (Trabucco and Zomer 2009). The AI values were downloaded from the CGIAR–CSI GeoPortal at http://www.csi.cgiar.org (accessed 20 August 2013). Watershed boundaries and catchment land use ArcGIS 10.1 was used to define the lake catchment area and investigate the catchment land use of the sampled lakes. A digital elevation model (DEM) was extracted from the US Geological Survey (USGS) HydroSHEDS site (available at http://hydrosheds.cr.usgs.gov, accessed 22 August 2013), with a resolution of 15 arc-seconds (equivalent to 304 m). Catchment areas were delineated with the DEM, and percentage land cover within each catchment was determined using the Australian dynamic land-cover (for every 250 m

27 | P a g e

Land-use, climate and limnology nexus of eastern Australia by 250 m area) overlay from the Geoscience Australia website (available at http://www.ga.gov.au/earth-observation/landcover.html, accessed 22 August 2013). There were four broad categories of land cover used in the analysis, including bushland, grassland, agriculture and urban. The bushland category included open and closed forest and woodland; the grassland category included sparse and scattered forest, shrubland, sedgeland, tussock and alpine grassland; the agriculture category included all forms of agricultural land use.

The dominant rock type in the lake catchment was determined using an ArcGIS data layer with 1 : 100 000 surface-geology map from Geoscience Australia (2012 edition, available at http://www.ga.gov.au/meta/ANZCW0703016455.html, accessed 25 August 2013).

3.4.4 Statistical analyses

Prior to ordination, selected climatic and lake water-chemistry data were assessed for normality using the Shapiro–Wilk normality test (Shapiro and Wilk 1965), and through measurements of skewness and kurtosis (Zar 1999) in Minitab 16 (Ryan et al. 2004). All variables apart from MAT and pH were normalised using a log10 transformation.

Both indirect and direct ordinations were used to analyse the data. Principal components analysis (PCA) was performed using CANOCO version 4.5 (ter Braak and Smilauer 2002), with samples being constrained by 18 limnological and land-use variables. Geology was excluded, because it was represented by a nominal scale (categories of rock type). Climate (aridity index), lake-morphology (water depth) and land-use variables were initially selected as passive variables. Passive variables do not directly affect the analysis. They are visible in the results so that their pattern can be observed; however, they are kept passive so that the study can focus the analysis on in-lake (limnological) variables.

Redundancy analysis (RDA) was used to explore the relationship among lake depth, aridity, land use and lake trophic status. Partial RDAs were performed in CANOCO v 4.5 (ter Braak and Smilauer 2002).

To investigate the origins of major cations and anions in the lake water, tri-plots were generated in Excel following Graham and Midgley (2000). Lake distance from the closest coastal area was measured using the Google Earth ruler tool.

28 | P a g e

Land-use, climate and limnology nexus of eastern Australia

Table 3.2 Selected limnological and land-use/land cover variables for the forty-five water bodies

29 | P a g e

Land-use, climate and limnology nexus of eastern Australia

3.5 Results

The lakes covered a wide range of climatic and physical characteristics. The mean annual temperature ranged from 3.7°C in Blue Lake, Kosciuszko, to 24.1°C in Horseshoe Lagoon, northern Queensland (Table 3.3). Lakes were distributed from the semi-arid areas (aridity index <0.5) of north-western Victoria, to the very humid areas (aridity index = 3.7) of western Tasmania (Table 3.3). Lake depth ranged from 0.2 m (Eagle Tarn, Tasmania) to the 65-m-deep crater lake (Lake Barrine, Queensland; Table 3.3).

The lakes were also very diverse in terms of limnology. The pH of lake water ranged between 4.7 and 9.2 (Table 3.3). Most of the lakes of western central Tasmania were acidic, whereas those in western Victoria were strongly alkaline. Montane lakes had low specific conductance (5–6 µS cm–1 for Kosciuszko lakes), whereas some Victorian coastal lakes had very high specific conductance (>15 000 µS cm–1; Table 3.3). Algal biomass (Chl-a concentration) ranged from 1 µg L–1 in the alpine lakes to 90 µg L–1 in the Victorian lakes. TP concentrations were lowest in the alpine lakes and some Tasmanian lakes (<0.01 mg L–1) and highest (3.6 mg L–1) in the Victorian lakes. TN concentration ranged from 0.11 mg L–1 in the alpine lakes to 36 mg L–1 in the Victorian lakes (Table 3.3). The trophic state of the lakes ranged from oligotrophic to hypereutrophic (Table 3.1), following the same trends. Lowland natural lakes of mainland Australia were mostly eutrophic to hypereutrophic (Table 3.1), which contrasts strongly with the situation in New Zealand and Tasmania.

Limits for N and P limitation were derived from Guildford and Hecky (2000). TN : TP ratios ranged from as low as 1, which is indicative of severe N limitation (Coalstoun Lakes), to 139, which is indicative of severe P limitation (Table 3.1). Most (56%) of the studied lakes were co-limited (TN : TP between 20 and 50), whereas only four lakes (9% of our dataset) were N-limited (Coalstoun Lakes and UQ Lakes) (Abell et al. 2010). The remaining 35% (16 lakes) were P-limited.

The land-use patterns in the lake catchment were highly variable. Alpine and highland lakes that are within national parks or other reserves were, generally, dominated by native woodland and grasslands. Lakes on private land and around some large reservoirs in northern Queensland were dominated by agricultural land, with both extensive grazing and irrigated agriculture (cropping and intensive grazing) occurring (Table 3.3).

30 | P a g e

Land-use, climate and limnology nexus of eastern Australia

Table 3.3 List of the 19 variables used, with mean, minimum and maximum values cited and the appropriate units highlighted

3.5.1 Lake-water chemistry

A ternary plot of major cations (Na+ and K+, Ca2+ and Mg2+; Fig. 3.2) showed that the 45 lakes were distributed close to the world-average seawater ratio (WASW). Unsurprisingly, coastal lakes (lowland lakes that are <100 km away from the sea) were closer to the ratio than were lakes far inland or at high elevation. Fig. 3.3 bi-plot demonstrates that pH co-varied with bedrock type and salinity, but that salinity was a stronger control on alkalinity than was the pH of the bedrock.

31 | P a g e

Land-use, climate and limnology nexus of eastern Australia

Figure 3.2 Tri-plot of major cations, with lakes labelled by type, as follows: CL, coastal lakes (lowland lakes that are <100 km away from the sea); M, montane lakes (>1000 m asl); IL, inland lakes (lowland lakes that are >100 km away from the sea). The tri-plot shows cation concentration as a percentage. All 45 lakes are distributed closer to the world-average seawater ratio (WASW) than to the world-average freshwater ratio. On average, lowland coastal lakes are closer to the ratio than inland lakes or those at high elevation. This indicates that the primary source of the salts in eastern Australian lakes is seawater.

32 | P a g e

Land-use, climate and limnology nexus of eastern Australia

Figure 3.3 Biplot of pH against conductivity. This demonstrates that pH co-varies with bedrock type and salinity. All saline lakes are strongly alkaline, whereas there is more variability in the pH of freshwater lakes. This means that the freshwater lakes are dominantly controlled by bedrock types. Lakes above log conductivity of 3 are classified as brackish.

3.5.2 PCA results

An initial PCA plot of the 45 water bodies constrained to the 19 environmental variables, with land- use, climate and lake-depth variables entered as passive variables, is presented in Fig. 3.4. The first and second axes of this plot were predominantly specific conductance and pH gradients. Axis 1 and Axis 2 explained 75.1% and 9.6%, respectively, of the variation in the environmental data. This result is unusual in both the high explanatory power of the first axis and the low explanatory power of the second axis and is an artefact of the large number of correlated water conductivity proxies.

So as to address this, we removed all ions from the analysis and then re-ran the PCA. The revised primary y-axis was predominantly a nutrient gradient and the second axis was a pH gradient. Axes 1 and 2 explained 70.1% and 18.0% of the variance, respectively. This indicated that nutrient status and pH were the longest gradients in the lake-chemistry variables (see Fig. 3.5). In Fig. 3.6, we display the results of a final PCA, with all environmental data (excluding ion concentration data) entered as active variables. The primary y-axis explained 46% of the variance and was a trophic 33 | P a g e

Land-use, climate and limnology nexus of eastern Australia status (TN, TP, Chl a), conductivity and agriculture gradient. A further 17.2% of the variance was explained on the second axis, which was primarily a temperature gradient. This PCA plot also indicated that eutrophic saline lakes tend to be shallow lakes located in arid regions.

To summarise the information from the PCAs, the first axis is always a nutrient and conductivity related axis showing that shallow lakes, especially those located in agricultural catchments, are affected by eutrophication. Unsurprisingly, given that our data extended over a latitudinal range of more than 2000 km and altitudinal range of over 2000 m, temperature was also a significant gradient. To further examine the implications of these findings, we applied a RDA to investigate the interactions among water depth, aridity, agriculture and nutrient status.

34 | P a g e

Land-use, climate and limnology nexus of eastern Australia

TL 1.0 a. 1.0 b. LPO LEA FWL PL Chla WP LTP

WL UQ AI TN LRD SEL NGR TPSO42- GBL LCT LCN LH LEM Bushland Pasture Cl- LMO Na+ CTL LD K+COND

BL SWL Agricult Axis 2 Axis LS LLL 2 AXIS Urban Mg2+ LA GHL Grasslan GRL LE BR FHD CLU LMB Cropping LAW Depth ETCOD KID LSP MAT Ca2+ LBA CHD HCO3- LPA RRD HOL LTK LFY JUL

PH

-1.0 -1.0 -1.0 Axis 1 1.5 -1.0 1.5 AXIS 1

Figure 3.4 Initial PCA plot of a) the 45 water bodies constrained to b) the 19 environmental variables, with land-use, climate, and lake depth variables entered as passive variables to focus the analysis on in-lake (limnological variables). The first and second axes of this plot are predominantly specific conductance and pH gradients; it shows a large number of correlated water conductivity proxies (ion concentrations).

35 | P a g e

Land-use, climate and limnology nexus of eastern Australia 1.0 a. 1.0 b.

LPO TL PLLEA FWL

WP AI Chla

UQ SEL WL LMO NGR LRD Bushland GBL LTP TN BL LCN LD LLL CTL LS LEM LCT TP LA LH SWL Pasture AXIS2 Grasslan

LE FHD 1 AXIS COND BR LSP CLU Agricult LBA ET GHLLAW Depth LMB KID GRL CHD RRD MAT JUL COD HOL LFY LPA LTK

PH

-1.0 -1.0 -1.0 AXIS 1 1.0 -1.0 1.0 AXIS 1

Figure 3.5 Revised PCA plot of a) the 45 water bodies constrained to b) the 12 environmental variables (ionic variables removed), with land-use, climate, and lake depth variables entered as passive variables to focus the analysis on the in-lake (limnological) variables. The first axis is now a nutrient gradient and the second axis is a pH gradient.

36 | P a g e

Land-use, climate and limnology nexus of eastern Australia

1.0 b.

1.0 a. CTL

LA BL LCN ET LEA LTP Grassland PL LLL AI WP LD LS LRD SEL FWL LCT LEM NGR JUL TN Chla TP

TL GRL LFY SWL Agriculture UQ

CHD 2 AXIS

AXIS 2 AXIS LTK Urban GBL BR HOL COND WL LSP GHL CLU LPO LMB FHD pH LBA LH LAW LMO COD KID RRD Bushland Depth LE LPA

-1.0 -1.0 MAT -1.0 1.0 AXIS 1 AXIS 1 -1.0 1.0

Figure 3.6 Final principal components analysis (PCA) with all environmental data (excluding ion concentration data) entered as active variables. (a) The 45 water bodies, along with (b) morphology, environment, climate and land-use variation patterns. The primary y-axis is a trophic status (total nitrogen (TN), total phosphorus (TP), Chlorophyll a), conductivity and agriculture gradient; the second axis is aligned to temperature. The PCA also indicates that eutrophic saline lakes tend to be shallow lakes located in arid regions.

37 | P a g e

Land-use, climate and limnology nexus of eastern Australia We undertook additional PCA analyses to compare the changes in limnology for artificial lakes and natural lakes along a gradient that reflects lake depth and human impact (see Fig. 3.7a, b). The Axis 1 sample scores from the first PCA (Fig. 3.7a) indicated the amount of agricultural activity in the lake catchments. The Axis 1 sample scores for the second PCA (Fig. 3.7b) indicated the trophic status of the lakes. Plotting the Axis 1 scores for both PCAs as a bi-plot allowed us to visualise the relationship between stressors (agriculture) and response (trophic status). Figure 3.7c displays the regression of the PCA axis scores for all lakes, whereas Fig. 3.7d displays the artificial lakes only. The trend of the regression was similar on both graphs, indicating that artificial lakes were similar to natural lakes in response to land use in the catchment. The conclusion is perhaps unsurprising; artificial lakes in agricultural catchments also tend to be more eutrophic.

3.5.3 RDA results

RDA results showed that ~53.2% of the variation in trophic status was explained by the combined AI, depth and agriculture data (see Table 3.4). Individually, AI and lake depth (depth) each accounted for 29.8% of the variation in trophic status, whereas agriculture accounted for 22.6% (for all, P < 0.001). When combined, lake depth and aridity accounted for 50.1% of the total variation. When the effect of lake depth and aridity was partialled out, agriculture accounted only for 6.3% of the variation and this was not statistically significant.

Table 3.4 Partial redundancy analysis (RDA) results for trophic status as limnological variable. This table demonstrates that the effect of climate and water depth is more significant than the agricultural status of the basin. In short, shallow water bodies in semi-arid settings are more eutrophic than are other lakes. AI, aridity index; Chl a, Chlorophyll a; TN, total nitrogen; TP, total phosphorus

Limnological variable Environmental variable Co-variable Explanatory power P-value (%) Trophic (TN,TP,Chla) Agriculture None 22.6 0.001 Trophic (TN,TP,Chla) Depth None 29.8 0.001 Trophic (TN,TP,Chla) AI None 29.8 0.001 Trophic (TN,TP,Chla) Depth, AI None 50.1 0.001 Trophic (TN,TP,Chla) Agriculture, depth, AI None 53.2 0.001 Trophic (TN,TP,Chla) Agriculture Depth, AI 6.3 0.082 Trophic (TN,TP,Chla) Depth, AI Agriculture 42.0 0.001 Trophic (TN,TP,Chla) Depth AI, agriculture 26.6 0.001 Trophic (TN,TP,Chla) AI Depth, agriculture 22.0 0.001

38 | P a g e

Land-use, climate and limnology nexus of eastern Australia

AXIS 2 AXIS AXIS 2 AXIS

AXIS 1 AXIS 1 a) b)

39 | P a g e

Land-use, climate and limnology nexus of eastern Australia

c) d) Figure 3.7 Principal components analyses (PCAs) to compare the changes in limnology for natural lakes and artificial lakes along a gradient that reflects lake depth and human impact. The Axis 1 sample scores from plot (a) shows the amount of agricultural activity in the lake catchments. The Axis 1 sample scores from plot (b) shows the trophic status of the lakes. The Axis 1 sample scores for both PCAs (a and b) are used to visualise the relationship between land use, aridity and limnology. The regression for (c) all lakes and (d) artificial lakes only. The trend of the regression is similar between the two graphs, indicating that artificial lakes have traits similar to those of natural lakes. The lower slope on the artificial lakes may indicate lower residence times. Overall, these results demonstrate that lakes in agricultural catchments are more eutrophic; however, see the main text because this trend is overwritten by climatic effects and lake morphology.

40 | P a g e

Land-use, climate and limnology nexus of eastern Australia

3.6 Discussion

There are four major points to highlight from these results. These relate to (1) the contrast between natural and artificial water bodies, (2) salinity, pH and rock type, (3) N and P limitation and (4) the interaction among climate, lake morphology and human impacts.

3.6.1 The effect of including reservoirs and other artificial waterways

The trophic status and chemistry of both artificial waterways and natural lakes reflected land cover and land use in the catchment. However, exploring the data further revealed that the key difference between the regressions in Fig. 3.7c, d was the range of the data. The artificial lakes did not include sites that were strongly eutrophic or strongly affected by agriculture. This finding was entirely predictable. Typically, reservoirs and recreation lakes are managed so as not to be susceptible to algal blooms and, thus, nutrient fluxes from the catchment are normally designed or managed to be low.

The other factor that may affect artificial lakes is residence time (Søballe and Kimmel 1987), with an expectation that the residence time in artificial lakes is lower than in natural lakes. Søballe and Kimmel (1987) suggested that, for the same nutrient input, reservoirs tend to have lower productivity because turnover times are faster. Our data showed a weak indication that this might be the case, because the slope on our regression was lower for the artificial lakes, although the shorter range of the data limited our ability to make a strong statement. Nevertheless, it is highly likely that residence times in artificial lakes in Australia are shorter than in natural lakes and we predict that this effect will also be significant.

3.6.2 Salinity, pH and bedrock effects

The primary source of the salts in eastern Australian lakes is from the ocean and old marine bedrock and there were higher concentrations in drier regions (Figs 3.2, 3.3). Salt concentration is highest in lakes in arid zones with high evaporative flux, and in maar craters where there is no outflow and long residence times.

Lakes on basalt and limestone are on average relatively alkali. However, the alkalinity varies quite strongly with climate (aridity). The northern Queensland maars are much less alkali than the 41 | P a g e

Land-use, climate and limnology nexus of eastern Australia western Victorian ones. This is an effect of the higher precipitation and higher P : EP ratio in montane near coastal Queensland than in semi-arid western Victoria.

Overall, we conclude that salinity is a stronger control on pH than is bedrock type. This can be observed in the Fig. 3.3 bi-plot. This figure demonstrates the buffering effect of salinity in that all highly saline lakes are alkali, whereas there is more variability in the pH of freshwater lakes. These - freshwater lakes are controlled by bedrock type and, in particular, by the abundance of HCO3 . In more saline lakes, H+ is buffered by Na+ and other cations.

3.6.3 Nitrogen and phosphorus limitation

In the present paper, we used the molar ratio of TN : TP as a guide to the dominant nutrient limiting algal productivity in each lake. Assigning each lake to a TN, TP, or co-limited (TN and TP) category was based on the suggested ranges published in Guildford and Hecky (2000). On the basis of the criteria of Guildford and Hecky (2000), 56% of the lakes in our dataset are co-limited, 35% are P-limited, and only 9% are N-limited (Table 3.1). Inferences of nutrient limitation based solely on TN : TP ratios should be considered a preliminary assessment and should be followed up by macronutrient enrichment experiments (laboratory bioassays or in situ studies). This is especially important because the reliability of the TN : TP ratio as a proxy for nutrient limitation has been challenged (White et al. 1985). Although the TN : TP ratio has commonly been used as an indicator of nutrient limitation in lakes (e.g. Abell et al. 2010), some studies (such as Morris and Lewis 1988; Bergström 2010) have suggested that the dissolved inorganic N (DIN) : TP ratio is a better indicator for nutrient limitation.

3.6.4 The interaction of human impact and climatic effects on lake nutrient status

The nutrient status of most mainland Australian lakes ranges from eutrophic to hypereutrophic, with only the Kosciuszko lakes and some northern Queensland crater lakes classed as oligotrophic. McComb and Davis (1993) identified key factors that caused water bodies in southwestern Australia to become eutrophic, and we argue that these factors also apply widely in eastern Australia. First, Australian soils frequently have poor nutrient-retention capacities. When fertilisers are added (containing N and P), they are poorly absorbed by the soils and are washed into the waterways. Second, the coincidence of the wet season with minimum ground cover as a result of ploughing and/or harvesting, especially in the Mediterranean climate zones of southern Australia, 42 | P a g e

Land-use, climate and limnology nexus of eastern Australia enhances run-off and compounds problems of nutrient flows to the lakes. And finally, many Australian lakes are small and ephemeral. They tend to have few or no outflows (endorheic basins) and this results in the concentration of nutrients and ions by evaporative flux. McComb and Davis (1993) concluded that nutrients are derived mainly from fertiliser applications in the catchments and so inferred that the trophic status is primarily a human artefact.

In contrast, Beklioglu et al. (2011) investigated the eutrophication of shallow lakes on a global scale. For lakes in Mediterranean climates, they concluded that nutrient concentrations in these lakes could rise rapidly during drought periods, as a result of evaporative flux and internal processes. They also noted that changes in lake levels suppress macrophyte growth and ‘tip’ the lake toward algal production, thus enhancing eutrophication.

Beklioglu et al. (2011) concluded that these effects could control the concentration of lake nutrients and productivity, even where human impact was low. We used forward selection to identify the most important variables in explaining nutrient variability and then undertook partial RDAs on these. Interestingly, when we partialled out the elements that relate to high nutrient status in lakes during a series of RDAs (Table 3.4), we observed that whereas water depth and aridity are both significant as independent variables, agricultural land use is not significant after the effects of water depth and aridity are partialled out (see Table 3.4). Conversely, lake water depth and aridity (or combined effect) retained their significance level after agriculture land use was partialled out. What this suggests is that shallow Australian lakes in semi-arid climate are susceptible to eutrophication irrespective of the surrounding land use and that agricultural practice simply enhances a process that would naturally occur anyway. Paleoecological studies to examine pre-agricultural systems would be an obvious method to test this slightly unexpected observation.

On a world scale, eutrophication is recognised as a major impact on lakes and is most often attributed to anthropogenic activities (Moss 1998; Smith et al. 1999; Rovira and Pardo 2006). Clearly, in most settings, agriculture and industrial effluent are the major sources of eutrophication. However, in much of semi-arid and arid Australia, it appears that agriculture is at best a secondary effect and that the lakes are likely to be naturally moderately to strongly eutrophic. Beklioglu et al. (2011) made a point of implicating droughts and variable lake levels in the alteration of lake behaviour. Australia has the highest inter-annual variability in rainfall of any continent on earth and if fluctuating water levels and intermittent droughts are indeed the key to natural eutrophication, then perhaps the provisional results presented here are unsurprising. Clearly, suggesting that

43 | P a g e

Land-use, climate and limnology nexus of eastern Australia eutrophication in many Australian systems is ‘natural’ is a significant conclusion and we propose it as a testable hypothesis rather than a demonstrated occurrence at this point.

3.7 Conclusions

The main conclusions from this study are as follows: (1) Most mainland Australian lakes are naturally moderately to significantly eutrophic. Oligotrophic lakes are limited to high alpine and perennially wet settings. (2) Reservoirs and other artificial lakes behave similarly to natural lakes, except that they tend to be less eutrophic. This is most likely due to catchment management and a shorter residence time. (3) The primary source of the salts in east Australian lakes is seawater, either derived directly from aerosols or indirectly via marine bedrock. (4) Significant number of lakes in this dataset that are limited by N and P, suggesting that both of these nutrients should be considered in addressing eutrophication in Australian lakes. (5) Mainland lakes are susceptible to becoming highly alkaline as a result of long term increases in salinity that are related to the concentration of cations during drought periods. (6) We propose as a testable hypothesis that climatically related eutrophication is more significant than nutrient loading by agriculture in the eastern Australian lakes that we have examined. This is particularly likely for shallow endorheic basins in semi-arid regions.

44 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function

4 A CHIRONOMID BASED TRANSFER FUNCTION FOR RECONSTRUCTING SUMMER TEMPERATURES IN SOUTH EASTERN AUSTRALIA

Chapter 4 is published in Palaeogeography, Palaeoclimatology, Palaeoecology (Chang et al., 2015a)

4.1 Summary

This chapter reports the development of a chironomid based summer temperature transfer function from a training set spanning a latitudinal gradient from the subtropics to Tasmania in southeastern Australia.

Highlights:

 First subfossil chironomid based transfer function from mainland Australia  Reconstructs mean February temperatures and will give new tool for quantitative palaeoclimate estimates from Australia  Reservoirs were included in the development of the training set.  Integrating the existing chironomid transfer function from Tasmania with this new model is a real possibility.

45 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function 4.2 Abstract

We present a new chironomid based temperature transfer function which was developed from a training set of 33 natural and artificial lakes from southeast Australia from subtropical Queensland to cool temperate Tasmania. Multivariate statistical analyses (CCA, pCCA) were used to study the distribution of chironomids in relation to the environmental and climatic variables. Seven out of eighteen available variables were significantly (p < 0.05) related to chironomid species variation and these were mean February temperature (9.5%), pH (9.5%), specific conductance (8.2%), total phosphorous (8%), potential evapotranspiration (8%), chlorophyll a (6.9%) and water depth (6.2%). Further pCCA analyses show that mean February temperature (MFT) is the most robust and independent variable explaining chironomid species variation. The best MFT transfer function was 2 a partial least squares (PLS) model with a coefficient of determination (r jackknifed ) of 0.69, a root mean squared error of prediction (RMSEP) of 2.33 °C, and maximum bias of 2.15 °C. Chironomid assemblages from actively managed reservoirs appear to match assemblages from equivalent natural lakes in similar climates and therefore can be included in the development of the chironomid transfer function. Although we cannot completely rule out some degree of endemism in the Tasmanian chironomid fauna, our analyses show that the degree of endemism is greatly reduced. Therefore, integrating the existing chironomid transfer function for Tasmania (Rees et al., 2008) with this new model is a real possibility.

Key words: chironomids, southeast Australia, summer temperature, transfer function, palaeoclimate, lakes

46 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function 4.3 Introduction

In temperate Australia, high-resolution continuous palaeoenvironmental records are scarce due to the relative paucity of permanent water bodies (see Chapter 3, section 3.1). As a result, palaeoenvironmental research is focussed on a few regions where these records exist (Petherick et al., 2011). Continuous records that extend back to the last glaciation maximum (LGM: c. 21,000 years ago) are rare and geographical coverage is poor (Petherick et al., 2013). Pollen is the most widely used proxy for these palaeoenvironmental reconstructions, and most of the reconstructions of climate in Australia are qualitative. This has limited the climate inferences that can be made from these records, which is unfortunate, as there are few estimates of the absolute scale of temperature change between glacial and interglacial times. Bioclimatic modelling has been used with some success (D'Costa and Kershaw, 1997) but the technique does not allow biotic effects to be easily separated from climate change (Jeschke and Strayer, 2008). Statistical approaches using pollen transfer functions have been developed for mainland Australia (Cook and Van der Kaars, 2006) and for Tasmania (Fletcher and Thomas, 2010). The former has not been widely applied and the latter is appropriate only to Tasmania. Transfer functions have also been developed using other organisms, such as diatoms (Tibby, 2004; Tibby and Haberle, 2007), but diatoms are used primarily for salinity and other limnological variables rather than temperature estimates. Molluscs (Edney et al., 1990; D'Costa et al., 1993) and beetles (Porch et al., 2009; Sniderman et al., 2009) have also been applied but a critical gap remains in the tools that are available to determine past changes in temperature in Australia.

Chironomids (Diptera: Chironomidae) have been widely used as a proxy in palaeoclimate and palaeoenvironmental studies (Walker and Paterson, 1985; Hofmann, 1986; Walker, 1987). Since temperature is a dominant factor in every aspect of the chironomid life cycle (e.g. egg hatching, larval and pupal development, adult emergence) (Armitage, 1995), many studies have focused on the influence of temperature on the distribution and abundance of chironomids, in the context of past climate change (Walker, 2002; Porinchu and MacDonald, 2003; Walker and Cwynar, 2006).

Most chironomid-based reconstructions have been carried out in temperate and sub-polar regions of the Northern Hemisphere, including areas in central and northern Europe (Olander et al., 1999; Larocque et al., 2001; Brodersen and Anderson, 2002; Luoto, 2009; Heiri et al., 2011) and northern North America (Porinchu et al., 2002, 2009; Barley et al., 2006; Brodersen et al., 2008; Medeiros and Quinlan, 2011). There are only few applications in the Southern Hemisphere and these are from New Zealand (Boubee, 1983; Woodward and Shulmeister, 2006; Dieffenbacher-Krall et al., 2007),

47 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function Tasmania (Rees et al., 2008), east Africa (Eggermont et al., 2010) and South America (Massaferro and Larocque-Tobler, 2013). A summary of the training sets on a global scale is presented in Appendix 1.

Early palaeoecological and palaeoclimate investigations of chironomids from the Southern Hemisphere were restricted to semi-quantitative interpretations because the numerical techniques for creating transfer functions were still in an early stage of development. For example, Boubee (1983) and Schakau (1993) investigated the modern chironomid distribution from New Zealand lakes using cluster analysis and ordinations. Classification techniques were applied down-core to interpret the fluctuations in fossil chironomid abundances and species from a 6000 year record from Lake Grasmere in New Zealand (Schakau, 1991) and a glacial transition to Holocene record from Blue Lake in alpine Mount Kosciusko (Chapter 3, Fig. 3.1), Australia.

Dimitriadis and Cranston (2001) performed the first quantitative reconstruction based on chironomids from eastern Australian lakes. Instead of using chironomid head capsules in the training set, they used the presence and abundance of chironomid exuviae from 68 water bodies in eastern Australia. Dimitriadis and Cranston (2001) then used the mutual-climate-ranges (MCRs) of the pupal exuviae in the training set to create a Holocene climate reconstruction from Lake Barrine (Atherton Tableland, northeast Queensland, Chapter 3, Fig. 3.1) based on the down-core chironomid head capsule record. They inferred up to 6 °C temperature change in the Holocene.

The first Southern Hemisphere transfer function based on chironomid head capsules was developed by Woodward and Shulmeister (2006) for New Zealand. Both summer temperature and chlorophyll a (Chl a) transfer functions were presented. The summer temperature transfer function was successfully applied to a record spanning the marine oxygen isotope stage 3/2 transition (~26,600– 24,500 cal yr BP) from lake deposits in Lyndon Stream, New Zealand (Woodward and Shulmeister, 2007). A second independent chironomid model for New Zealand was produced by Dieffenbacher- Krall et al. (2007), which also yields satisfactory reconstructions (Vandergoes et al., 2008).

In Australia, a head capsule based chironomid transfer function was developed by Rees et al. (2008) for summer temperature from Tasmanian lakes. This transfer function was applied to produce lateglacial (~16,000 cal yr BP) to late Holocene summer temperature reconstructions from Eagle Tarn and Platypus Tarn in Mount Field National Park, Tasmania (Chapter 3, Fig. 3.1) (Rees and Cwynar, 2010). Although the Tasmanian chironomid transfer function appears to be robust,

48 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function biogeographical controls on the Tasmanian chironomid taxa may prevent the application of the Tasmanian transfer function to mainland sites.

Concerns over the influence of biogeography on Tasmanian and mainland Australian chironomid taxa stem from a chironomid exuviae survey of eastern Australian lakes by Wright and Burgin (2007). Wright and Burgin argue for the presence of 23 endemic taxa from Tasmania, including 5 genera, 4 species and 14 undescribed morpho-species. Despite this assertion, Wright and Burgin's study does not completely rule out the possibility of applying the Tasmanian transfer function to the mainland or producing a combined mainland and Tasmanian transfer function. Wright and Burgin (2007) did not include mainland alpine lakes in their study and the degree of Tasmanian endemism may be over-estimated. The level of taxonomic resolution provided by exuviae is typically higher than for chironomid head capsules. Even if there are endemic chironomid species in Tasmania, their presence might not dramatically affect the composition of sub-fossil chironomid head capsule assemblages.

Here we present a temperature transfer function based on sub-fossil assemblages of Chironomidae (non-biting midges), from 33 southeast Australian lakes. We included 7 lakes from Tasmania in this total as an initial test of the feasibility of producing a chironomid training set combining both mainland and Tasmanian lakes. We assessed Wright and Burgin's argument for Tasmanian chironomid endemism using our new training set and published information on the distribution of Wright and Burgin's endemic taxa. A combined training set would be desirable because it is difficult to find non-impacted, permanent freshwater lakes spread continuously along a long temperature gradient on mainland Australia. Due to the rarity of freshwater bodies on the mainland, we also investigated the possibility of including artificial water bodies in the training set.

49 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function Table 4.1 Selected climatic and environmental variables for the thirty-four water bodies sampled from southeast Australia used for transfer function development

50 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function

4.4 Materials and Methods

4.4.1 Study sites

This data set comprises 25 natural lakes and 8 artificial water bodies located in southeast Australia (Chapter 3, Fig. 3.1, lake numbers 1-34 and Table 4.1). The transect covers a distance of 2500 km along the east coast of Australia from Kureelpa, Queensland to Mount Field National Park, Tasmania (25.96°S to 42.67°S, 140.18°E to 153.26°E) (Chapter 3, Fig. 3.1, lake numbers 1-34 and Table 4.1). The climate ranges from subtropical in the north, to cool temperate in the south, and hence there are large temperature and precipitation gradients in the data set. Elevation of the sites ranges from sea level to ~2000 m above mean sea level (a.s.l) (Table 4.1), corresponding to estimated mean February air temperatures (MFT) of 10.7–24.7 °C (Table 4.1). A detailed description of the climate, vegetation and geology of the study area is provided in Chapter 3, section 3.

4.4.2 Chironomid collection and analysis

The lakes were sampled during the Southern Hemisphere summer (January or February) of 2012 and 2013 (Chapter 3, Fig. 3.1 and Table 4.1). Lakes were selected along an altitudinal and latitudinal range to ensure a long temperature gradient. Where possible we sampled shallow to medium depth (between 1 and 10 m) lakes to ensure a close relationship between bottom water temperature and air temperature. The sampling of very deep ( >30 m), stratified lakes was avoided to eliminate the effect of hypolimnetic anoxia on the chironomid species assemblages (Little and Smol, 2001). A minimum of three sediment cores were collected using a Glew Mini Corer (Glew, 1991) at the deepest point or lake centre where bathymetry was not available. The top 2 cm of each core were extruded on site and packaged at 0.5 cm intervals in Whirlpak® sample bags. Sediment samples were refrigerated prior to analysis.

Surface sediment core (top 1-2 cm) samples were processed for chironomid analysis following the method outlined in Hofmann (1986) with the following modification. Samples were deflocculated in warm 10% KOH for 20 min and washed on a 90 µm mesh with distilled water. Samples were transferred to a Bogorov counting tray and examined under a dissection microscope at 50 × magnification. Chironomid head capsules were hand-picked using fine forceps onto a glass slide, until a minimum of 100 head capsules were obtained (when possible). Chironomid head capsules were mounted on glass slides in a drop of Euparal® and covered with a glass coverslip. Head capsules were mounted ventral side up to assist identification. Chironomid species were identified 51 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function using a compound light microscope at 400 × magnification, following the published identification guides by Freeman (1961), Cranston (2000a, 2010), Brooks et al. (2007) and Dieffenbacher-Krall et al. (2008).

Several studies have examined the minimum number of chironomid head capsules that should be extracted to provide representative samples. Larocque (2001) found that 50 head capsules is the minimum required to provide an accurate temperature estimate, but counts of 90 head capsules or above will give much better representation of taxa in the assemblage. Quinlan and Smol (2001) concurred with this finding, whereas Heiri and Lotter (2001) argued that the minimum count size is model and location dependent. We therefore used rarefaction analysis (Birks and Line, 1992) in R (version 2.11.1, R Development Core Team, 2010) and the Vegan package (version 2.0-10, Oksanen et al., 2010) to test how representative different sample sizes were in our training set. A plot of observed species richness vs predicted species richness was derived using the ‘estimateR’ function which uses Chao's method (Chao, 1987) to estimate actual richness. We also created multiple rarefaction curves for each site based on multiple random subsamples from the full species pool. This allowed us to visualise the effect of simulated increased sampling intensity on the observed species richness.

4.4.3 Lake water chemistry and environmental variables

Water samples for chemical analyses were collected from 30 cm below the water surface at the location where the core samples were taken. Untreated water samples were collected for the + + 2+ 2+ 3- - 2- analysis of major ions (Na , K , Ca , Mg , HCO , Cl , SO4 (Chapter 3, Table 3.2) and total nitrogen/total phosphorus (TN/TP) analysis. These samples were kept frozen until analysis. A 1000 ml water sample was filtered for Chl a through a 4.7 cm diameter GF/F filter (0.45 µm pore size).The Chl a filter was wrapped in foil and kept frozen for subsequent analysis. The 125 ml water sample was filtered using a syringe and a Whatman® syringe filter (Supor® Membrane 0.45 µm pore size) and the filtered water was frozen in the bottle for later analysis for dissolved reactive phosphorus (DRP) and reactive nitrogen (NOX ). Total dissolved solids (TDS), pH, specific conductance (COND) and turbidity (TURB) were also obtained from water chemistry analyses which were carried out by the Forensic and Scientific Services, in Brisbane, Queensland. In the field, water temperature, oxidation reduction potential (ORP), pH, dissolved oxygen (DO), specific conductance (COND), total dissolved solids (TDS), salinity (SAL), and turbidity (TURB) were

52 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function recorded from 30 cm below the water surface using an Aquaread® multi-parameter meter. Lake depth at the sampling point was measured using a Speedtech® portable depth sounder.

Figure 4.1 (a) Plot of observed species richness vs predicted species richness, (b) rarefaction curves for individual sites indicating estimated species richness with respect to increasing sub-sample size. Rarefaction curves begin to flatten once true species richness is achieved. Together, these plots indicate that sample sizes are adequate for most samples to capture all but the most rare species. Only one site (CTL) with a low head capsule count may possibly underestimate the true species richness. Minimum sample size varied from site to site and counts as low as 50 may be sufficient to capture the actual species richness (e.g. LPO).

Climate variables were obtained using the combination of the WorldClim program (available from http://www.worldclim.org/ bioclim, accessed 20 August 2013) and ArcGIS 10.1. WorldClim data for Australia is based on climate surfaces derived from around 600 nation-wide weather stations that have climate records spanning the years 1950–2000 (http://www.bom.gov.au/climate/data/stations/, accessed 20 January, 2014). For this study, mean February temperature (MFT), mean annual temperature (MAT) and mean annual precipitation

53 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function (Precip) were considered. The potential evapotranspiration (PET) values were obtained from the Global Potential Evapo-Transpiration (Global-PET) and Global Aridity Index (Global-Aridity) data set (CGIAR-CSI, available from http://www.cgiar-csi.org/data/globalaridity-and-pet-database, accessed 20 August 2013).

4.4.4 Statistical analyses

Test for Tasmanian endemism and difference between natural and artificial lakes

A principal components analysis (PCA) of the chironomid taxa data was run for the 33 lakes. We tested to determine if the chironomid assemblages were significantly different in mainland lakes compared to Tasmanian lakes to look for potential biogeographical effects. We also tested for significant differences between natural and artificial lakes. The distribution of both sets, PCA axis scores was assessed for normality using the Shapiro–Wilk test (Shapiro and Wilk, 1965) respectively. The data were normally distributed around the mean so that a Student t-test was appropriate to test significance of the results. We first performed a t-test (α = 0.05, two sample assuming unequal variances) on sample axis scores from a principal components analysis (PCA) of the chironomid taxa based on head capsule assemblages for Tasmania vs. mainland lakes. This is not an exhaustive test for endemism as it is based on the taxonomic resolution possible with head capsules. In order to further test for endemism in the Tasmanian chironomid taxa we performed a literature search for records of Wright and Burgin's (2007) endemic Tasmanian taxa, and searched the Australian National Insect Collection records (Atlas of Living Australia database: http://bie.ala.org.au/species/CHIRONOMIDAE, accessed 20 July 2014). The same t-test was performed on natural vs. artificial lakes.

Selection of environmental variables and model development

Constrained ordinations were performed using CANOCO version 4.5 (ter Braak and Šmilauer, 2002) to determine which variable(s) explained a significant proportion of the variation in the chironomid species data. Prior to ordination, climate and lake water chemistry data were assessed for normality in Minitab 16® using the Shapiro–Wilk test (Shapiro and Wilk, 1965), and through measurements of skewness and kurtosis (Zar, 1999). All variables apart from MAT, MFT and pH were normalized using a log10 transformation. Chironomid species were used in the form of square root transformed percentage data.

54 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function A detrended correspondence analysis (DCA) (with rare taxa down-weighted) of the chironomid data was used to determine whether linear or unimodal methods were appropriate for selecting the best candidates for model construction. The gradient length for DCA axis 1 was 2.067 standard deviation units, which means that a unimodal technique (Canonical correspondence analysis, CCA) is appropriate (Birks, 1998). CCAs were run to determine which environmental variables explained the highest and most significant amount of the variation in the chironomid species data. The environmental variable with highest variance inflation factor (VIF) was removed after each CCA and the CCA was repeated until all VIFs were less than 20 (ter Braak and Šmilauer, 2002). The ability of the remaining environmental variables to explain a statistically significant amount of the variation in the chironomid species data was determined using a series of CCAs with manual forward selection and a Monte Carlo permutation test (999 unrestricted permutations) (ter Braak and Šmilauer, 2002). Variables that were significant (p ≤ 0.05) were retained for further analyses. To test the strength of the explanatory power of each of the significant variables for the chironomid distribution, a series of partial canonical correspondence analyses (pCCAs) were performed with the remaining significant variables included as co-variables. This step was used to distinguish between indirect and direct relationships between the environmental variables and the chironomid species data. Only environmental variables that retained their significance after this step were considered for transfer function development. CCA bi-plots of sample and species scores were generated using CanoDraw (ter Braak and Šmilauer, 2002). Chironomid taxon response curves for significant variables (p ≤ 0.05) were generated in CanoDraw using Generalized Linear Models (GLMs) with a Poisson distribution.

Transfer functions for the significant environmental variable(s) selected in the CCAs and pCCAs were developed in the computer program C2 (Juggins, 2005). A detrended canonical correspondence analysis (DCCA) was used to determine whether chironomids were responding in a linear or unimodal fashion along the gradient of the environmental variable selected for model development (Birks, 1998). Leave-one-out, cross-validation was used as this technique is more robust for data sets with fewer than 80 sites (Kim and Han, 1997).

Transfer functions were selected based on the performance of the jack-knifed coefficient of 2 determination (r jackknifed), average bias of jack-knifed predictions (AveBias jack), maximum bias of jack-knifed predictions (MaxBias jack), and root mean square error of prediction (RMSEP) (Birks, 1998). Additional components were only included in the model if the addition of an additional component reduced the RMSEP by at least 5% (Birks, 1998).

55 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function 4.5 Results

4.5.1 Chironomid taxa

The average sample size in the training set is 128 head capsules and only two samples (Chaffey Dam (n = 0) and Lake Cootapatamba (n = 59)) produced fewer than 100 head capsules. Counts of over 100 are generally regarded as reliable (Heiri and Lotter, 2001). Rarefaction analysis (Fig. 4.1) indicated that the sample size for most sites was adequate for including all of the common chironomid species. There is a significant correlation between observed and predicted species richness (Fig. 4.1a) and only one site (Lake Cootapatamba) produced a low head capsule count (59) which possibly under-represents the full chironomid species pool. All head capsules from the top 2 cm of the Cootapatamba sediment core were used. The plot of multiple rarefaction curves (Fig. 4.1b) indicates that the minimum number of counted head-capsules that is sufficient to capture the actual species richness varies from site to site.

One site (Chaffey Dam) had no head capsules and was removed from the dataset for further analyses. Forty-three (43) chironomid taxa were identified and counted from the training set. Five rare species were removed from further analyses as they have a maximum abundance of less than 2% and/or occurred in fewer than two lakes (Brooks and Birks, 2001).

4.5.2 Test for Tasmanian endemism and difference between natural and artificial lakes

Chironomid species assemblages for the seven Tasmanian lakes show no significant difference to mainland lakes based on t-test results (Fig. 4.2a, Table 4.2a). Fourteen (14) of Wright and Burgin's 23 endemic taxa are undescribed morpho-species (Table 4.3), so it is not possible to assess these taxa for endemism. The remaining 9 endemic taxa comprise 5 genera and 4 species. Head capsules from one of Wright and Burgin's (2007) endemic Tasmanian taxa (Thienemanniella sp.) were found on the Australian mainland in Blue Lake and Lake Albina (Mount Kosciuszko), from our training set. Head capsules from Wright and Burgin's (2007) other 4 “endemic” Tasmanian genera (Apsectrotanypus sp., Pentaneurini genus E, Nanocladius sp., “MO5” (Now = Echinocladius sp. Cranston)) were not found in our training set but have been previously recorded from mainland sites by Marchant et al. (1999), Cranston (2000a), Cranston (2000b), and Krosch (2011) respectively (Table 4.3). Wright and Burgin's (2007) four endemic Tasmanian species

56 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function (Botryocladius australoalpinus, B. grapeth, Riethia plumosa, Tanytarsus liepae) were not identified in the training set, but have been previously recorded from the Australian mainland by Cranston and Edward (1999), Cranston and Edward (1999), Cranston (Australian National Insect Collection, available from http://bie.ala.org.au/species/CHIRONOMIDAE, accessed 20 July 2014), and Cranston (2000a) respectively (Table 4.2).

Figure 4.2 PCA plots for exploring the difference between (a) mainland and Tasmanian lakes and (b) natural and artificial lakes. PCA axis 1 and axis 2 explains 16.9% and 10.8% of the variance in chironomid species data respectively. A t-test was performed on the sample score means for each (Tables 4.1a and 4.1b). The sample size for the Tasmania and artificial lakes is small (<10). There are no significant differences apparent between axis 1 scores for both tests. There are no differences in axis 2 scores either for Tasmania vs mainland lakes, but there is for natural vs artificial lakes. Warm taxa are over-represented in artificial lakes that are not reservoirs (Fig. 4.3).

The t-test of PCA axis scores shows that there is no significant difference in the chironomid assemblages from natural and artificial water bodies on PCA axis 1, but there is on PCA axis 2 (Fig. 4.2b and Table 4.2b). PCA axis 2 mainly separates warm stenotherms (low PCA axis 2 scores, e.g. Dicrotendipes, Kiefferulus, Cladopelma) from cold stenotherms (high PCA axis 2 scores, e.g. Parakiefferiella) (Fig. 4.3a). Warm stenotherms are more common in three high altitude, shallow artificial lakes (not reservoirs) (Highland Waters (LD), Lake Samuel (LS) and Lake Cantani (LCN)) than other high altitude, natural lakes (Fig. 4.3b).

57 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function

Figure 4.3 This shows a PCA of chironomid species used to compare natural and artificial lakes. In a) PCA axis 2 show separation of warm stenotherms (low PCA axis 2 scores, e.g. Dicrotendipes, Kiefferulus, Cladopelma, Procladius, Paratanytarsus etc.) from cold stenotherms (high PCA axis 2 scores, e.g. Parakiefferiella). b) shows surface sediment samples plotted as circles where the circle size is relative to mean February temperature (MFT). Large circles = high MFT, small circles = low MFT. Values for MFT for each site are also provided. Open circles = natural water bodies, closed circles = artificial water bodies. Note that some artificial water bodies (not reservoirs) with low MFTs have low PCA axis 2 scores, which indicates that warm stenotherms are more common in these sites compared to natural sites with similar MFTs.

Table 4.2 Statistical t-Tests for two samples assuming unequal variances performed on (a) natural against artificial lakes and (b) Tasmania against mainland lakes. PCA scores of axis 1 and axis 2 were used from both datasets from Fig. 4.2a and Fig. 4.2b respectively. (a) Tasmania VS Mainland Lakes AXIS 1 Variable 1 Variable 2 Mean 0.538314286 -0.144930769 Variance 1.105297491 0.951724631 Observations 7 26 Hypothesized Mean Difference 0 df 9 t Stat 1.549215177

58 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function Tasmania VS Mainland Lakes

AXIS 1 Variable 1 Variable 2

P(T<=t) one-tail 0.077869722 t Critical one-tail 1.833112933 P(T<=t) two-tail 0.155739445 t Critical two-tail 2.262157163 No Difference

AXIS 2 Variable 1 Variable 2 Mean 0.6701 -0.180415385 Variance 2.580589327 0.541095722 Observations 7 26 Hypothesized Mean Difference 0 df 7 t Stat 1.362846194 P(T<=t) one-tail 0.107574164 t Critical one-tail 1.894578605 P(T<=t) two-tail 0.215148328 t Critical two-tail 2.364624252 No difference

(b) Natural VS Artificial Lakes AXIS 1 Variable 1 Variable 2 Mean -0.05321 0.017028 Variance 0.86037 1.122792 Observations 8 25 Hypothesized Mean Difference 0 df 13 t Stat -0.17989 P(T<=t) one-tail 0.430005 t Critical one-tail 1.770933 P(T<=t) two-tail 0.860011 t Critical two-tail 2.160369 No difference AXIS 2 Variable 1 Variable 2 Mean -0.6306 0.201788 Variance 0.139766 1.159287 Observations 8 25 Hypothesized Mean Difference 0 df 31 t Stat -3.29437 P(T<=t) one-tail 0.001237 t Critical one-tail 1.695519 P(T<=t) two-tail 0.002473 t Critical two-tail 2.039513 Difference 59 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function

Table 4.3 Wright and Burgin (2007) identified 5 genera, 4 species and 14 morpho-species of endemic chironomids in Tasmania. These are arranged in alphabetic order, below. We performed a literature search and show that all the 5 genera and 4 species were previously recorded and described in mainland Australia, therefore endemism is not supported at these taxonomic levels. The 14 morpho-species are not described and we are unable to assess the reliability of the endemism claim for these taxa. (*) Indicates undescribed morpho-species. Name Genus Species Morphotype Comments species* Apsectrotanypus sp. X Paratypes from mainland Australia (originally Anatopynia maculosa Freeman 1961). Apsectrotanypus also collected from the mainland Marchant et al. (1999) Botryocladius X Described only from Pupae (Cranston and australoalpinus Edward 1999) Botryocladius X Botryocladius spp. head capsules collected grapeth from Blue Lake. Larvae collected from Blue Lake (Cranston and Edward 1999) sp. 1 1 Chironominae sp. 2 1 Cricotopus sp. 1 1 Cricotopus sp. 6 1 Nanocladius sp. X Not found in the training set in Tasmanian or mainland sites, however endemism to Tasmania can be elimated since this genus has been found in Western Australia (Cranston 2000a) Orthocladiinae sp. 1 1 Orthocladiinae sp. 2 1 Orthocladiinae sp. 3 1 Orthocladiinae X Echinocladius is distributed along the east "MO5" coast of Australia in streams from the tropics (now Echinocladius in the north to Tasmania (Krosch 2011) sp. Cranston 2000b) Parakeifferiella sp. 3 1 Pentaneurini genus X Actually states in Cranston (2000a) that this E (Cranston 2000a) genus has also been found in New South Wales Polypediulum sp. 3 1 Polypediulum sp. 8 1 Riethia sp. 1 1 Riethia sp. 2 1 Riethia plumosa X Riethia plumosa is also found on the mainland. Australian National Insect Collection e.g. catalogue number 29-002080-767, pupal exuviae, Mongarlowe River. Records available from The Atlas of Living Australia database Tanytarsus sp. 11 1 Tanytarsus sp. 12 1 Tanytarsus liepae X Collected from Victoria and Southern NSW 60 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function Name Genus Species Morphotype Comments species* Thienemanniella sp. X Head capsules found in Blue Lake and Lake Albina in the training set. SUM 5 4 14

4.5.3 Selection of environmental variables and model development

Individual ions (see Chapter 3, Table 3.2), precipitation (Precip), and mean annual temperature (MAT) were excluded prior to ordination. Cation and anion gradients are correlated in PCA space and are better represented by specific conductance (see Chapter 3, Fig 3.4). Rainfall is usually a secondary effect on chironomids species distribution where its effect on chironomids is through influencing or altering the lake water chemistry by dilution and in-lake macrophyte composition and structure through changes in water depth. However, potential evapotranspiration (PET) was included in this dataset since evaporative balance drives salinity and nutrient gradients in sub-humid and semiarid areas, especially for shallow lakes with endorheic basins (Chapter 3). In Chapter 3, we concluded that PET is a much stronger climate driver for lake water chemistry and nutrient status changes of the east coast Australian water-bodies than rainfall alone. The choice of summer temperatures as a control on chironomids is routine for the alpine lakes and those in more temperate settings. Only two of our sites (UQ lakes, Lake Poona) are not located in temperate settings as defined by the Köppen–Geiger climate classification (Kottek et al., 2006), so it is reasonable to assume that summer is the prime breeding period for chironomids in this study also.

Total nitrogen (TN) produced the highest VIF in a CCA with all selected environmental variables and chironomid species and it was therefore not considered for further analyses. The remaining seven variables individually accounted for a significant portion (p ≤ 0.05) of the variance (Fig. 4.4). In order of explanatory power, these were mean February temperature (9.5%), pH (9.5%), specific conductance (8.2%), total phosphorus (8%), potential evapotranspiration (8%), Chl a (6.9%) and water depth (6.2%).

Southeastern Australian lakes in this training set cover large productivity and temperature gradients. The first four axes of the CCA constrained by the seven significant environmental variables (Fig. 4.4) account for 28% of the variance in the chironomid species data (Table 4.4). Depth, MFT and pH are significantly correlated to the first axis and Depth, TP, pH and specific conductance are correlated with the second axis. Mean February temperature (MFT) shows the strongest correlation

61 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function with the first axis (Table 4.4). Total phosphorus (TP) shows the strongest correlation with the second axis (Table 4.4).

Partial CCAs (pCCAs) were then undertaken to determine the direct and indirect effects of each of the seven significant variables. The pCCA results show that lake water depth (DEPTH), nutrient variables (TP, Chl a) and specific conductance (COND) are confounded, while potential evapotranspiration (PET) is correlated with specific conductance. pH retained 7.7% of variance explained and remained significant (p ≤ 0.05) after the pCCAs with all other significant variables partialled out (Table 4.5). Although MFT and PET appear to be confounded, PET is the dependent variable because evaporation is partially a function of temperature and not vice-versa. In summary, pH and MFT are the two primary independent parameters that should be considered for model construction. They both explained the largest amount of variance (both 9.5%) (Table 4.5).

Table 4.4 CCA summary of the seven significant variables including canonical co-efficients and t- values of the environmental variables with the ordination axes including 33 lakes and 38 non-rare species Axis 1 Axis 2 Axis 3 Axis 4 Eigenvalue 0.199 0.109 0.078 0.042 Cum % var. spp. 13.0 20.2 25.3 28.0 Cum % var. spp. - env. relation 38.5 59.7 74.8 83.0

Regression/canonical co-efficient Variable Depth 0.384 -0.463 0.704 -0.813 MFT -0.745 0.523 0.869 -0.193 PET 0.262 0.230 -0.645 0.336 TP 0.277 -1.009 0.002 -1.034 Chl a -0.302 -0.402 0.097 0.432 pH -0.650 -0.654 0.506 0.804 COND 0.115 0.912 -0.545 -0.991 t-values for regression co-efficients

FR explained 0.385 0.212 0.151 0.082 Variable Depth 2.136 -2.432 3.453 -3.316 MFT -2.336 1.547 2.401 -0.442 PET 0.824 0.684 -1.787 0.775 TP 0.835 -2.871 0.005 -2.285 Chl a -1.032 -1.297 0.291 1.083 pH -3.207 -3.045 2.200 2.904 COND 0.443 3.316 -1.850 -2.797

62 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function Although pH is also a good candidate for a transfer function, the focus here is on developing a palaeoclimate proxy and MFT is a more useful parameter. The taxon response curves and a plot of chironomid species turnover with respect to temperature (Figs. 4.5 and 4.6) show that the major taxa (p ≤ 0.05 and N2 ≥ 5) which are dominant at warm sites are Polypedilum spp., Parachironomus spp., Coelopynia pruinosa type, Paratanytarsus spp., Tanytarsus lactescens type and Procladius spp. Taxa that are typical of cool sites include Paralimnophyes morphotype 1and Botryocladius. However, many taxa may respond to other environmental variables as well (e.g. Cladopelma, Dicrotendipes and Kiefferulus, see Fig. 4.6 and Table 4.6). A few taxa such as Chironomus, Procladius, Pentaneurini and undifferentiated Tanytarsini contain many species each of which is likely to have different environmental responses but cannot be separated from similar morphotypes.

Table 4.5 Partial CCAs of the seven significant (p ≤ 0.05) environmental variables alone and with the effects of other significant variables partialled out for 33 lakes with 38 non-rare species included.

Variable covariable λ λ / λ % Variance P-value 1 1 2 explained Depth none 0.095 0.302 6.2 0.016 MFT 0.100 0.412 7.2 0.002 PET 0.092 0.347 6.5 0.014 TP 0.062 0.207 4.4 0.163 Chl a 0.064 0.221 4.5 0.133 pH 0.077 0.304 5.6 0.031 COND 0.067 0.240 4.7 0.089 MFT, PET,TP, Chl a, pH, COND 0.071 0.353 6.5 0.034

MFT None 0.145 0.562 9.5 0.001 Depth 0.151 0.621 10.5 0.001 PET 0.062 0.240 4.4 0.150 TP 0.125 0.508 8.9 0.001 Chl a 0.112 0.446 7.9 0.001 pH 0.100 0.450 7.2 0.004 COND 0.095 0.380 6.8 0.007 Depth, TP, Chl a, pH, COND 0.069 0.332 6.2 0.029 Depth, PET,TP, Chl a, pH, COND 0.057 0.284 5.3 0.139

PET None 0.123 0.449 8.0 0.010 MFT 0.040 0.155 2.9 0.468 Depth 0.120 0.453 8.3 0.009 63 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function Variable covariable λ λ / λ % Variance P-value 1 1 2 explained PET TP 0.100 0.377 7.1 0.012 Chl a 0.082 0.306 5.8 0.035 pH 0.081 0.343 5.8 0.044 COND 0.069 0.261 4.9 0.075 MFT, COND 0.036 0.145 2.7 0.523 Depth, MFT, TP, Chl a, pH, 0.034 0.169 3.2 0.529 COND

TP None 0.122 0.408 8.0 0.003 MFT 0.102 0.439 7.4 0.004 PET 0.099 0.374 7.0 0.009 Depth 0.089 0.297 6.2 0.025 Chl a 0.058 0.201 4.1 0.173 pH 0.066 0.261 4.8 0.083 COND 0.086 0.308 6.1 0.012 Chl a, pH 0.063 0.278 4.8 0.107 Depth, MFT, PET, Chl a, pH, 0.046 0.229 4.4 0.282 COND

Chl a none 0.106 0.367 6.9 0.004 MFT 0.073 0.291 5.3 0.032 PET 0.066 0.246 4.7 0.080 Depth 0.075 0.260 5.2 0.050 TP 0.042 0.145 3.0 0.517 pH 0.065 0.261 4.7 0.068 COND 0.046 0.167 3.2 0.456 Depth, MFT, PET, TP, pH, COND 0.029 0.144 2.8 0.765

pH None 0.146 0.574 9.5 0.001 MFT 0.101 0.462 7.3 0.003 PET 0.105 0.445 7.4 0.004 Depth 0.129 0.514 9.0 0.002 TP 0.091 0.360 6.4 0.005 Chl a 0.106 0.426 7.4 0.002 COND 0.091 0.361 6.4 0.011 Depth, MFT, PET, TP, Chl a, 0.085 0.423 7.7 0.013 COND

COND None 0.125 0.448 8.2 0.002 MFT 0.075 0.300 5.4 0.044 PET 0.071 0.269 5.1 0.055 Depth 0.097 0.348 6.8 0.004 TP 0.089 0.319 6.3 0.015

64 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function Variable covariable λ λ / λ % Variance P-value 1 1 2 explained COND Chl a 0.065 0.236 4.5 0.127 pH 0.069 0.274 5.0 0.053 PET, Chl a, pH 0.059 0.251 4.7 0.114 Depth, MFT, PET, TP, Chl a, pH 0.058 0.289 5.4 0.125

4.5.4 The transfer function

The DCCA results (gradient length = 1.07 standard deviation units) suggest a linear response of chironomid taxa along the mean February temperature gradient (Birks, 1998), therefore, a partial least squares (PLS) model was appropriate for the transfer function construction (ter Braak and Juggins, 1993) for MFT (Table 4.7) in C2 (Juggins, 2005). The third component of the PLS model (with 3 components, jack-knifing, including 33 lakes and 38 non-rare species, Table 4.7) was 2 selected based on the criteria of Birks (1998). It produced a coefficient of determination (r jackknifed) of 0.69, RMSEP jack of 2.33 °C, maximum bias jack of 2.15 °C and an AveBias jack of 0.07 °C (Table 4.7, Fig. 4.7).

65 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function

a. a.

Figure 4.4 CCA biplots of (a) sample and (b) species scores constrained to seven environmental variables that individually explain a significant (p ≤ 0.05) proportion of the chironomid species data. Sites codes correspond to site names in Tables 3.1 and 3.2 (lake numbers 1-34). Taxon numbers correspond to taxa in Table 4.5. Sites and taxa in warmer environments tend to plot in the upper left quadrant. Taxa typical of warmer sites include Harnischia spp., Cryptochironomus spp., Polypedilum spp., Coelopynia pruinosa type, Cladopelma spp., Paratanytarsus spp., Procladius spp., and Riethia spp., while sites and taxa in colder environments tend to plot in the lower right hand quadrant. Taxa typical of cold sites include Paralimnophyes morphotype 3, Parakiefferiella morphotype 1, Orthoclad type 1, Orthoclad type 4, Pseudosmittia type 2, and Botrycladius. Eutrophic sites and taxa typical of these environments tend to plot in the lower left hand quadrant, while oligotrophic sites and taxa typical of these environments tend to plot in the upper right hand quadrant. 66 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function

Figure 4.5 Taxon response curves for taxa that show a significant response to temperature (p < 0.05) using a Generalized Linear Model with Poisson distribution (ter Braak and Šmilauer, 2002). (a) Taxa which are more common at lower temperatures, such as Paralimnophyes morphotype 1 and Parakiefferiella morphotype 2 respond strongly to cooling in temperature (b) Taxa which are more common at higher temperatures demonstrate weaker but still significant responses. Examples include Procladius, Polypedilum spp. and Tanytarsus lactescens type.

67 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function

Figure 4.6 Stratigraphy diagram of the 38 non-rare taxa included in the final model, where observed mean February temperature is on the y-axis and taxon abundance is in percentage. Taxa such as Cladopelma, Dicrotendipes and Kiefferulus (*) show high abundance in lowland warm lakes but are also present in highland artificial lakes (Grey bars).

68 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function

Table 4.6 List of chironomid taxa enumerated in this study along with data on distribution and environmental significance.

No. Taxa name N Hill's N2 Maximum Mean PLS β - Environmental variables that coefficent have a significant relationship to (Jack-knifed) each taxon’s response (at p ≤ 0.05) based on the GLM results 1 Chironomus Meigen 31 19.8 63.7 19.3 -0.2 - 2 Polypedilum nubifer Skuse 27 19.6 15.8 6.0 0.9 MFT 3 Cladopelma Kieffer 24 12.0 23.2 4.2 0.3 Depth 4 Dicrotendipes Kieffer 24 14.3 25.8 6.6 -0.1 Chla, Depth, TP, COND 5 Polypedilum spp. Kieffer 6 5.1 4.4 0.5 0.9 MFT 6 Cryptochironomus Kieffer 8 5.7 4.7 0.4 0.2 COND 7 Parachironomus Lenz 11 6.9 10.9 1.2 0.4 Chla, MFT, TP 8 Kiefferalus martini Freeman 16 11.7 18.5 3.5 0.3 Chla, Depth 9 Procladius Skuse 33 24.9 32.8 15.3 0.1 - 10 Coelopynia pruinosa Freeman 8 5.9 4.7 0.6 1.2 MFT 11 Pentaneurini undifferentiated 17 10.0 12.3 1.8 -0.7 COND 12 Riethia Kieffer 21 9.8 28.9 4.8 -0.3 COND 13 Tanytarsus lugens type 29 16.8 23.1 6.6 -0.3 - 14 Tanytarsus pallidicornis type 28 17.7 27.0 7.2 -0.3 TP 15 Tanytarsus glabrescens type 18 8.2 20.9 2.6 -0.4 Chla, TP, pH 16 Paratanytarsus Skuse 22 17.2 14.3 4.0 0.6 Chla, MFT, COND 17 Tanytarsus undifferentiated 22 17.3 6.4 1.9 0.1 MFT 18 Tanytarsus lactescens type 10 5.2 27.0 2.4 0.5 Depth, MFT, TP, pH 19 Tanytarsus nr chiyensis 5 3.8 3.6 0.3 -0.3 MFT, pH 20 Stempellina Thienemann & Bause 2 1.8 44.6 2.1 0 Chla, MFT, TP, pH 21 Harnischia Kieffer 4 2.9 3.1 0.2 1.2 MFT 22 Paralimnophytes type 1Unofficial morphotype 13 5.5 15.8 1.3 -0.5 Chla, MFT, COND, pH 23 Paralimnophytes type 2Unofficial morphotype 13 8.6 5.4 1.0 -0.3 Chla 24 Paralimnophytes type 3 Unofficial morphotype 9 6.9 3.6 0.5 -0.3 - 25 Parakiefferiella undifferentiated 7 6.0 4.3 0.6 -0.8 Depth 26 Parakiefferiella type1Unofficial morphotype 3 3.0 1.6 0.1 -0.3 -

69 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function No. Taxa name N Hill's N2 Maximum Mean PLS β - Environmental variables that coefficent have a significant relationship to (Jack-knifed) each taxon’s response (at p ≤ 0.05) based on the GLM results 27 Parakiefferiella type 2 Unofficial morphotype 4 3.1 13.7 0.9 0 Chla, MFT, TP, COND, pH 28 Parakiefferiella type 3 Unofficial morphotype 4 3.2 2.9 0.2 -1.7 Chla, MFT, COND, pH 29 Botrycladius Cranston & Edward 9 5.4 7.0 0.7 -0.6 MFT 30 Smittia Holmgren 3 1.9 7.8 0.3 0.3 - 31 Gymnometriocnemus type 1 Unofficial morphotype 7 3.2 8.8 0.6 0.1 - 32 Genus Australia 2 2.0 1.2 0.1 -1.1 - 33 Thienemanniella Kieffer 4 2.7 5.8 0.3 -0.3 Chla, Depth, MFT, TP, COND 34 Ortholcad type 1 Unoffical morphotype 3 2.5 1.8 0.1 -0.8 Chla 35 Orthoclad type 4 Unoffical morphotype 8 7.3 3.4 0.5 -0.3 - 36 Pseudosmittia type 2 Unofficial morphotype 2 1.6 2.3 0.1 0.2 MFT, TP, COND 37 Kosciuszko Orthoclad type 1 Unoffical morphotype 3 2.4 2.7 0.1 -0.4 TP 38 Cricotopus ‘Parbicintus’ 2 2.0 1.4 0.1 -0.1 Chla, Depth, MFT, TP, pH

70 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function

4.6 Discussion

4.6.1 Can we include artificial lakes in temperature training sets?

Recent study (Chapter 3) showed that reservoirs and other artificial water bodies respond to stressors in their catchments in a similar fashion to natural lakes. Despite the general preference for natural lakes for temperature training sets, this observation is not unexpected as human impacts that could change water quality in reservoir catchments are generally limited. We might expect chironomid species composition in reservoirs and natural lakes in environmentally equivalent settings to be similar.

However, we also note that for three high elevation artificial lakes (not reservoirs) (LD, LS and LCN), chironomid assemblages resemble lowland eutrophic lakes, with high values of Cladopelma, Dircrotendipes and Kiefferiulus (see Fig. 4.3, Fig. 4.6 and Table 4.6). These three artificial water bodies comprise two nutrient rich trout stocked lakes in Tasmania and a shallow, productive recreation lake on Mt Buffalo (Table 4.1, lake number 5, 9, 10). If the two Tasmanian lakes are excluded, the distinction between warm and cold lakes is markedly diminished (Fig. 4.6). On Mt Buffalo, the shallowness of Lake Cantani increases its mean temperature in summer, making it resemble a lower elevation lake due to the increased abundance of Cladopelma.

In summary, chironomid assemblages from true reservoirs appear to match assemblages from equivalent natural lakes in similar climates. We should be cautious about including other types of artificial water bodies, especially those where the food web or trophic status might be significantly altered.

4.6.2 Endemism and other considerations

In our dataset there was no significant difference between the chironomid assemblages from Tasmania and the Australian mainland (Fig. 4.2a, Table 4.2a) for the taxonomic resolution that is available for subfossil head capsules. This test can be considered a first order test of the possibility for combining Tasmanian and mainland training sets. As we will mention below, we did detect a genus in our dataset (Thienemanniella sp.) that was previously considered endemic to Tasmania (Wright and Burgin, 2007), but this does not fully rule out some degree of Tasmanian endemism.

71 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function To do this we need to have chironomid distribution records with the same taxonomic resolution that was provided in Wright and Burgin's (2007) study on Australian chironomid biogeography.

Five genera (including Thienemanniella sp.) and four species that Wright and Burgin (2007) identified as Tasmanian endemics have actually been reported from the Australian mainland (Table 4.3). This means that the basis for Tasmanian endemism now relies on the presence of 14 undescribed morpho-species in Wright and Burgin's (2007) dataset. It was not possible for us to further test for endemism in records of adult and larval chironomid distributions for morpho- species that are based on exuviae alone. Further collection of exuviae from mainland lakes and rearing of chironomid larvae to possibly associate morpho-species with other life stages of described species is required to rule out endemism. In the absence of these data, alternative methods for testing for endemism would include combining our dataset with Rees et al. (2008) and splitting the pooled dataset into Tasmanian and Mainland sites. Mainland sites can then be used to reconstruct temperatures from Tasmanian sites and vice-versa. This technique has been used to compare trans-Atlantic chironomid datasets (Lotter et al., 1999). At this stage we can only conclude that claims for endemism are greatly diminished, but we do not expect future efforts to combine datasets to be thwarted by endemism.

4.6.3 Reconciliation and integration with Tasmanian transfer function

We recognize that the temperature error in this transfer function is relatively high in comparison to other transfer functions. This reflects the long scalar length of the temperature gradient which extends from sub-tropical to sub-alpine locations, some 14 °C. The RMSEP is 2.33 °C which represents 16% of the scalar length. This is comparable with the recently developed western Irish chironomid-based calibration set (Potito et al., 2014). The Potito et al. (2014) dataset has a RMSEP of 0.57 °C, but the temperature range this dataset covers is only 3.8 °C; so the error represents 15% of the scalar length. Furthermore, it has been observed that data sets with large temperature gradients naturally have larger errors (Walker and Cwynar, 2006) but this in no way diminishes the value of the reconstruction.

The number of water bodies used in the study is small (33), but this is a function of the rarity of lakes on continental SE Australia. There are additional natural lakes to sample but they are exclusively in areas and elevations that we have already sampled. When we established this study we attempted to focus on natural water bodies. It is clear that in order to extend the training set,

72 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function more reservoirs will need to be included. Even allowing for this, there is a gap between summer temperatures of c. 13.3–14.8 °C for which there are no ideal candidate lakes or reservoirs. There are some reservoirs at these temperatures (e.g. Lake Jindabyne) but they are exceptionally large and deep lakes, and require both alternative sampling strategies and some analyses and consideration before including in the data set. The other alternative is to integrate this model and data set with the Tasmania model and data set of Rees et al. (2008). We have deliberately replicated some of the sites (e.g. Eagle Tarn) from Rees et al. (2008) so that the models can be compared and harmonised in due course and this is an obvious next step for this research.

a. b.

Figure 4.7 Performance of the three component PLS model where (a) shows the predicted versus observed mean February temperature and (b) displays residuals of the predicted versus observed mean February temperature. Note that the model has a potential to over predict temperatures from some very shallow high altitude lakes by up to ~6 °C. These lakes have increased mean water temperature in summer and chironomid assemblages may resemble lower elevation sites.

4.6.4 Value of the transfer function

This transfer function has relatively large errors for detecting change during periods of relative climate stability (i.e. in the Holocene). However, on longer time scales, the expected change from Last Glacial Maximum (LGM) to the Holocene in southeastern Australia is between 8 and 10 °C (Galloway, 1965; Miller et al., 1997). The precision of the current transfer function is ample to constrain climate change of this magnitude. Transfer functions with relatively large errors may still be able to provide a reliable indication of variation in temperature through time and this should be tested using the method developed by Telford and Birks (2011).

73 | P a g e

Development of the southeastern Australia subfossil chironomid transfer function

Table 4.7 Performance of partial least squares (PLS) model for reconstructing mean February temperature of southeast Australia using 33 lakes and 38 non-rare chironomid species. The bold indicates the model of choice. 2 2 Method r r Jack RMSEP MaxBias Av % (apparent) (Jack) (Jack) Bias reduced (Jack) PLS Component 1 0.67 0.45 3.15 4.39 0.13 - Component 2 0.87 0.57 2.80 3.28 0.18 10.85 Component 3 0.93 0.69 2.33 2.15 0.07 16.35 Component 4 0.94 0.67 2.45 2.07 -0.01 -4.30 Component 5 0.95 0.66 2.56 1.83 -0.05 -4.56

4.7 Conclusions

We constructed a February mean temperature transfer function based on the modern distribution of chironomid (Diptera: Chironomidae) species in southeast Australia. The training set comprises 33 natural and artificial lakes in locations that span the subtropics to the alpine zone. The February 2 mean temperature model is statistically robust with an r jackknifed of 0.69, a RMSEP of 2.33 °C and a maximum bias of 2.15 °C. The transfer function is suitable for the reconstruction of summer temperatures during the last glacial maximum (LGM) and the late glacial to Holocene transition in southeastern Australia. In a context where there are few reliable and no continuous estimates of palaeo-temperature available from mainland Australia, the transfer function represents a significant advance for palaeoclimatological studies.

We also conclude that chironomid assemblages in actively managed reservoirs (non-impacted) show no significant difference to natural lakes in the same climate and vegetation zones, and therefore can be included for transfer function development. Although we cannot completely rule out some degree of endemism in the Tasmanian chironomid fauna, our analyses show that the degree of endemism indicated by earlier studies is greatly reduced. This raises the real possibility of integrating the existing chironomid transfer function for Tasmania with this new model for the southeast Australian mainland.

74 | P a g e

Application of the transfer function - late-Quaternary temperature reconstruction

5 A CHIRONOMID-INFERRED SUMMER TEMPERATURE RECONSTRUCTION FROM SUBTROPICAL AUSTRALIA DURING THE LAST GLACIAL MAXIMUM (LGM) AND THE LAST DEGLACIATION

Chapter 5 is published in Quaternary Science Reviews (Chang et al., 2015b)

5.1 Summary

This chapter reports the first application of the newly developed chironomid based transfer function in Chapter 4, for temperature reconstructions from Welsby Lagoon, North Stradbroke Island, Australia, detailing the period of the last glacial maximum and the last deglaciation (c. ~ 23.2 and 15.5 ka cal yr BP).

Highlights:  A chironomid based summer temperature reconstruction from Australia.  Reconstruction covers the LGM and part of deglaciation.  Cooling is similar to that derived from nearby marine records.  Deglaciation is synchronous with Antarctica.  Suggests climate link from Australian subtropics to high latitudes.

75 | P a g e

Application of the transfer function - late-Quaternary temperature reconstruction 5.2 Abstract

A chironomid-based mean February temperature reconstruction from Welsby Lagoon, North Stradbroke Island, Australia covering the last glacial maximum (LGM) and deglaciation (between c. ~23.2 and 15.5 cal ka BP) is presented. Mean February temperature reconstructions show a maximum inferred cooling of c. ~6.5 °C at c. ~18.5 cal ka BP followed by rapid warming to near Holocene values immediately after the LGM. The inferred timing, magnitude and trend of maximum cooling and warming display strong similarities to marine records from areas affected by the East Australian current (EAC). The warming trend started at c. ~18.1 cal ka BP and is consistent with the start of deglaciation from Antarctic records. Near Holocene values are maintained through the deglaciation to 15.5 cal ka BP. These records suggest that changes in the Australian subtropics are linked to southern high latitudes.

76 | P a g e

Application of the transfer function - late-Quaternary temperature reconstruction 5.3 Introduction

Australia is the second most arid continent after Antarctica and, as a consequence, high quality terrestrial palaeoenvironmental records suitable to investigate late Quaternary climate changes are rare (Reeves et al., 2013; Petherick et al., 2013). A number of paleoecology records come from the Atherton Tablelands in North Queensland (Kershaw, 1986; Haberle, 2005; Turney et al., 2006), an area that is subject to tropical climate systems and is unlikely to provide good palaeoclimate information for non-tropical southeastern Australia. Similarly, significant research has occurred in western Victoria (reviewed in Gouramanis et al., 2013) but again this is in a climatically different environment to subtropical eastern Australia. One of the key time intervals for climate reconstruction is the last glaciation maximum (LGM; 21–17 ka yr, Barrows et al., 2002; Williams et al., 2009), as this interval represents the most significantly different global climate in the recent geological past and has been a focus for palaeoclimatic modelling (e.g. PMIP2: Braconnot et al., 2007). Australia has traditionally been poorly represented by these models, because there is a dearth of validation data. This remains true, despite concerted efforts through the Australasian INTIMATE project (Turney et al., 2007; Reeves et al., 2013) to improve the data sets. Traditional biological proxies such as pollen are widely applied (PalaeoWorks: Australian National University. Dept. of Archaeology and Natural History (2000)) but the records, at least away from the Atherton Tablelands and Victoria (e.g. Caledonia Fen and Tower Hill) (Kershaw et al., 1991, 2007; D'Costa et al., 1989), are mostly intermittent and have not yielded quantitative palaeoclimatic information (Kershaw et al., 1994; Kershaw and Bulman, 1996).

Existing quantitative reconstructions of climate at the LGM suggest the climate in southeastern mainland Australia was significantly cooler, by 8–12 °C (Galloway, 1965 from periglacial landforms, Miller et al., 1997 from emu shell amino acid racemisation) and with increased aridity and seasonality (Galloway, 1965) inferred but not quantified. Recently, an LGM mean annual temperature (MAT) reconstruction from Lake Mackenzie on Fraser Island in Southeast Queensland (Fig. 5.1; 25°26′51″S, 153°03′12″E, 90 m a.s.l) was developed using branched glycerol dialkyl glycerol tetraethers (GDGTs), which suggested a cooling of only 4.1 °C at the LGM (c. ∼18.8 ± 0.5 cal ka BP) (Woltering et al., 2014). This lower value is consistent with results from Tasmania (Fig. 5.1) where Fletcher and Thomas (2010) estimated only about 4.2 °C cooling at the LGM from a pollen transfer function. In summary, there is both a remarkable paucity of LGM paleothermometers and a clear disagreement between the available reconstructions, which may reflect either regional variability, or the unreliability of one or more of the techniques, or both.

77 | P a g e

Application of the transfer function - late-Quaternary temperature reconstruction Southeast Queensland lies at the northern limit of the Southern Hemisphere winter westerlies (Fig. 5.1) (McGowan et al., 2008). This is the interface between climate driven by the Inter-Tropical Convergence Zone derived from the Indonesian tropics and the southwest Pacific convergence zone, and the high latitudes driven by Antarctica and the Southern Ocean. Changes in this region are particularly valuable as they are likely to give insights into the re-organisation of the major circulation systems during the LGM.

Subfossil chironomids (Order: Diptera, Family: Chironomidae) have been proved as an applicable proxy for palaeo-temperature reconstructions since the late 1980s (Hofmann, 1986; Walker, 1987). Chironomid-based transfer functions have been successfully applied to Southern Hemisphere late Quaternary palaeoclimatic data (Woodward and Shulmeister, 2007; Rees and Cwynar, 2010; Massaferro et al., 2014) and have provided quantitative reconstructions. Here, we present the application of a newly developed subfossil chironomid-based transfer function for the reconstruction of mean February temperatures (MFTs) (in Chapter 4) to an LGM to Holocene record from subtropical Australia. We preferred MFT to Mean Summer Temperature because it provided the best model coefficient (r2) value and conceptually late January and February are the most stable and warm period in the summer at higher latitudes. The site is Welsby Lagoon in Southeast Queensland (27°26′12″ S, 153°26′ 56″ E) (Fig. 5.1). Pollen is applied to reconstruct the moisture balance from the site (Moss et al., 2013), whereas diatoms are absent from the core sediment of Welsby Lagoon. Therefore, chironomid based reconstructions from the site will provide valuable complementary information.

5.4 Regional setting

North Stradbroke Island is located on the east coast of Australia at 27.58° S, 153.47° E (Fig. 5.1). It is 32 km long and has a maximum width of 11 km at its northern end. It is 239 masl (metres above sea level) at its highest point (Mt Hardgrave). North Stradbroke Island is one of several large sand islands that form the distal end of the world's largest down drift sand system. This sedimentary system is fed by rivers on the mid-New South Wales coast, a thousand kilometres to the south, and the longshore drift transports roughly 500,000 m3 of sand per metre of beach front per year north (Roy and Thorn, 1981). The sand drift system has been operative for at least the last c. 750,000 ka yr (Tejan-Kella et al., 1990) and is likely much older. The island is composed of numerous sets of parabolic dunes that can be traced to beach terraces and are inferred to relate to cycles of sea-level rise (Ward, 2006). There is a large aquifer underlying the island and dune lakes and wetlands relate to either breaches of the aquifer (window lakes) or the impoundment of dune hollows by organics 78 | P a g e

Application of the transfer function - late-Quaternary temperature reconstruction (perched lakes/wetlands) (Laycock, 1975). Because of the antiquity of the dune fields, many of the lakes are partly infilled by organics and provide outstanding targets for paleoenvironmental and paleoclimatic work.

On North Stradbroke Island, research has focussed on aeolian records from Native Companion Lagoon (McGowan et al., 2008; Petherick et al., 2009) and a broader study of paleoecology from lagoon and swamp systems across the island including Welsby Lagoon, Tortoise Lagoon (Moss et al., 2013) and Blue Lake (Barr et al., 2013). The Welsby Lagoon record was selected as a target for this study as it contains thick gyttja covering the duration of the LGM and the last glacial – interglacial transition (Moss et al., 2013).

5.4.1 Climate

North Stradbroke Island lies in the east-coast subtropical zone under the Kӧppen – Geiger (Cfa: humid subtropical climate zone) classification but close to the border with the temperate zone to the south. The island has warm and wet summers with a mean summer temperature of ∼24 °C and mild winters with mean temperatures of ∼14 °C (BOM, 2015). The island has a mean annual rainfall of between 1500 (west coast) and 1700 mm (east coast) and precipitation is summer dominated (BOM, 2015). A number of modern climatic drivers influence precipitation in the region including the El Niño/Southern Oscillation (ENSO) and the Pacific Decadal Oscillation (PDO). In general, these oscillatory systems affect the climate by modifying sea surface temperatures (SST). During positive (negative) phases of ENSO and negative (positive) phases of the PDO, SSTs of the west Pacific are higher (lower) and there is a positive (negative) rainfall anomaly (BOM, n.d.). The coast is also more susceptible to cyclones and ex-tropical cyclones during the positive phases of ENSO and negative phases of the PDO. ENSO and PDO interact and ENSO either amplifies or damps the background PDO signal (BOM, n.d.).

5.4.2 Modern circulation of the Eastern Australian Current (EAC)

North Stradbroke Island is situated adjacent to the mid-section of the East Australian Current (EAC) which is formed between 10° and 15°S, and in between the two anti-cyclonic recirculation cells (Fig. 5.1). The EAC is known to be a highly variable and energetic system with large mesoscale eddies dominating the flow (Bowen et al., 2005; Mata et al., 2006). The EAC is both the western boundary of the South Pacific Gyre and the linking element between the Pacific and Indian Ocean 79 | P a g e

Application of the transfer function - late-Quaternary temperature reconstruction gyres (Speich et al., 2002). The current is accelerated southward along the coastal boundary, and then separates into north-eastward (Subtropical Counter Current), eastward (Tasman Front) and a residual southward (EAC Extension) component at around 31°S (Fig. 5.1, Ridgway and Dunn, 2003). The EAC has an important role in removing heat from the tropics and releasing it to the mid- latitude atmosphere (Roemmich et al., 2007). It has the strongest southward flow during the austral summer and the location of separation also migrates along the coast seasonally (Ridgway and Godfrey, 1997). Observations for the past 60 years show that the EAC has strengthened and extended further southward (Ridgway and Hill, 2009) in response to the modern warming of climate.

5.4.3 Physical environment of Welsby Lagoon

Welsby Lagoon is one of several perched lagoons on North Stradbroke Island and is situated near the northwest coast at 27°26′12″ S, 153°26′ 56″ E and 29 masl (Fig. 5.1). In 2012, it had a maximum water depth of c. ∼1.3 m and was dominated by sedges and reeds, with large areas of Melaleuca woodland fringing the lagoon. The surrounding vegetation consists of Eucalyptus forest and woodlands and Casuarinaceae woodlands, with a predominantly heath understorey (Moss et al., 2013).

80 | P a g e

Application of the transfer function - late-Quaternary temperature reconstruction Figure 5.1 Location of the study site and other sites discussed in this study. (a) A summary of the surface currents within the Tasman Sea. The denoted by A, B, C, and D shows the direction of flow of the relevant part of the EAC along with where the SST records are mentioned. Note that B is associated with the Tasman Front (Ridgway and Dunn, 2003). The locations of marine cores PC-27 (Dunbar and Dickens, 2003), GC-12 (Bostock et al., 2006), GC-25 (Troadson and Davies, 2001), PC-9 (Troadson and Davies, 2001) along the EAC, terrestrial temperature estimates from Lake Victoria (Miller et al., 1997), the Snowy Mountains (Galloway, 1965), and Lake Selina (Fletcher and Thomas, 2010) are displayed (b) Location of North Stradbroke Island (NSI) and Lake Mackenzie, Fraser Island (c) Location on North Stradbroke Island of Welsby Lagoon and other sites mentioned in the text, including Tortoise Lagoon (Petherick et al., 2008; Moss et al., 2013), Blue Lake (Barr et al., 2013), Eighteen Mile Swamp (sampling location) and Native Companion Lagoon

(McGowan et al., 2008; Moss et al., 2013).

5.5 Methodology

5.5.1 Chironomids sampling and analysis

One 450 cm core from Welsby Lagoon was collected by Moss et al. (2013) in October, 2009 using a Russian corer (Jowsey, 1966). Half of the core was used for pollen, micro-charcoal and radiocarbon dating analyses (Moss et al., 2013), while the remaining half core was sub-sampled for this study. Two-centimetre sections of sediment were collected for chironomid analyses with a sampling resolution of one sample every 30 cm from the top to 340 cm, and every 10–20 cm from 340 cm to the base of the core.

Chironomid analysis followed the standard procedure outlined by Walker and Paterson (1985). Samples were deflocculated in warm 10% KOH for 20 min and washed on a 90 µm mesh with distilled water. Samples were transferred to a Bogorov counting tray and examined under a dissecting microscope at 50× magnification. Chironomid head capsules were hand-picked using fine forceps and placed ventral-side up on a glass slide, then mounted in a drop of Euparal® with a cover slip. We classify this as a minimum of 50 head capsules that are fully identified (Heiri and Lotter, 2001; Quinlan and Smol, 2001) (Fig. 5.3b). Chironomid taxa were identified using a compound microscope at 400 × magnification following Cranston (2000, 2010) and Brooks et al. (2007).

81 | P a g e

Application of the transfer function - late-Quaternary temperature reconstruction Zonation of the chironomid record was achieved using the constrained incremental sums of squares (CONISS) function in Psimpoll (Bennett, 2002). This software also allows the determination of the optimal number of zones using the broken stick model. All chironomid taxa were included in the analysis and species data was square-root transformed.

5.5.2 Chronology

Chironomid samples for this study were taken from the same core as the pollen record described in Moss et al. (2013). We therefore utilise the chronology developed by Moss et al. (2013), which is based on 11 radiocarbon dates on a mix of pollen, charcoal and peat samples. However, we have updated the model by calibrating the radiocarbon ages against the more recent SHCal13 calibration curve (Hogg et al., 2013) (Table 5. 1). The resulting age-depth model is presented in Fig. 5.2.

Figure 5.2 Radiocarbon and calibrated ages for Welsby Lagoon with respect to depth. Open shapes represent calibrated ages calibrated using SHCal13 (Hogg et al., 2013).

82 | P a g e

Application of the transfer function - late-Quaternary temperature reconstruction Table 5.1 Radiocarbon Dates for Welsby Lagoon. Data shown includes sample depth, laboratory, materials used for dating, 14C age and calibrated age (SHCal13; Hogg et al., 2013) for each of the samples. (*) indicates samples excluded in the final age depth model. Depth Lab Code Sample Code Material AGE (14C yr) Calibrated (cm) Dated Age (Cal BP) 80 Wk-29036 Welsby 80cm charcoal 1803 +/- 30 BP 1710 125 Wk-29037 Welsby 125cm charcoal 4070 +/- 30 BP 4494 165 Wk-30668 Welsby 165-166cm charcoal 5556 +/- 26 BP 6311 251 Wk-29038 Welsby 251cm charcoal 8043 +/- 30 BP 8852 287 Wk-30669 Welsby 287cm charcoal 9903 +/- 38 BP 11295 304 Beta - 306711 WL p304 pollen 14250 +/- 50 BP 16970* 342 Wk-29039 Welsby 342cm Peat 18696 +/- 88 22245* 398 Beta - 306712 WL p398 pollen 16350 +/- 70 BP 19446 414 Wk-30670 Welsby 414cm charcoal 29096 +/- 250 33787* 435 Beta - 270553 Welsby-435cm peat 18320 +/- 90 21868 438 Beta - 306713 WL p438 pollen 18980 +/- 80 BP 22644

5.5.3 Statistical methods and reconstruction examination

The southeastern Australian chironomid based transfer function developed from a 33-lake training set for temperature (mean February temperature, MFT) (Chapter 4, Table 4.1) was applied to the chironomid data. The transfer function training set did not include any lakes from North Stradbroke Island so one adjacent site, Swallow Lagoon (c. ∼7 km south of Welsby Lagoon) was added to the original training set to provide a local analogue site. MFT of each site in the modern calibration data set was obtained using the combination of the WorldClim program (available from http://www.worldclim.org/bioclim, accessed 20 August 2013) and ArcGIS 10.1. WorldClim data for Australia is based on climate surfaces derived from around 600 nation-wide weather stations that have climate records spanning the years 1950–2000 (http://www.bom.gov.au/climate/data/stations/, accessed 20 January, 2014) (see Chapter 4). The model was created using partial least squares (ter Braak and Juggins, 1993) in C2, v1.5 (Juggins, 2005) where percent values of fossil taxa were square-root transformed. The addition of Swallow Lagoon did not alter the parameters of the previously established transfer function. The model has a 2 jack-knifed coefficient of determination (r jack ) of 0.69 and a root mean square error of prediction (RMSEP) of 2.2 °C.

To evaluate the reliability of the chironomid-inferred MFT reconstruction, we used the reconstruction diagnostics outlined in Birks (1995). Goodness-of-fit (Birks et al., 1990) was assessed by passively fitting fossil samples to the canonical correlation analysis (CCA) ordination 83 | P a g e

Application of the transfer function - late-Quaternary temperature reconstruction axis derived from the modern training set constrained to MFT in CANOCO version 4.5 (ter Braak and Šmilauer, 2002). The squared residual length (SqRL) was used to assess the quality of fit by comparing the distance between modern samples and the constrained axis to the distance between fossil samples and that same axis. Any fossil sample whose residual distance was equal to or larger than the residual distance of the extreme 5% of the training set is considered to have a ‘very poor’ fit to temperature, and those with values equal to or larger than the extreme 10% were deemed to have a ‘poor’ fit (Birks et al., 1990). The chi-square distance (dissimilarity coefficient) to the closest modern analogue for each fossil sample was calculated in C2 (Juggins, 2005) using the modern analogue technique (MAT) (Birks et al., 1990). Fossil samples with a chi-square distance to the closest sample in the modern calibration data set larger than the 5th percentile of all chi-square distances in the modern assemblage data were identified as samples with ‘no good analogues’ (Anderson et al., 1989; Larocque-Tobler et al., 2010). The percentage of rare taxa in the fossil samples was calculated, where a rare taxon has a Hill's N2 ≤ 5 in the modern data set (Hill, 1973) in C2 (Juggins, 2005). Fossil samples that contain high percentage (>10%) of these rare taxa were likely to be poorly estimated (Brooks and Birks, 2001). Finally, the fossil assemblages were plotted passively to a CCA of the training set with all significant variables (p < 0.05) in CANOCO version 4.5 (ter Braak and Šmilauer, 2002) to determine the direction of change through time and the influence of the environmental variables on the fossil chironomid assemblages.

5.6 Results

5.6.1 Chironomid analysis

A total of 42 sub-samples were taken throughout the 450 cm core where the first sample was undertaken at the depth of 55 cm. Chironomids were absent from the depth between 277 cm and 340 cm of the core which corresponds to the period between c. ∼10.5 cal ka BP and c. ∼15 cal ka BP. Discontinuous chironomid counts were obtained between 55 cm–277 cm, which covers roughly the last ∼10.5 ka calendar years. Altogether, 17 samples yielded enough chironomids for reliable quantitative reconstructions. These include four from the top 277 cm (0 – c. ∼10.5 cal ka BP) (Fig. 5.3), eight from the depths between 340 cm and 377 cm (c. ∼15.5–17.8 cal ka BP) and five from the depths between 385 cm and 442 cm (c. ∼18.5–23.2 cal ka BP) (Fig. 5.3).

Goodness-of-fit analyses of the 17 samples from depths of 55 cm to the bottom of the core (442 cm) show that only one sample (at the depth of 275–277 cm) does not have a good fit to temperature

84 | P a g e

Application of the transfer function - late-Quaternary temperature reconstruction (Fig. 5.4a and b). Rare taxon analysis of these samples indicate 4 of the samples between 375 and 442 cm (c. ∼18.1–23.2 cal ka BP) have abundances of taxa that are considered rare (Hill's N2 ≤ 5) totalling to greater than 10% (Fig. 5.4c). Rare taxa Tanytarsus chinyensis type, Parakiefferella morphotype 3 and Stempellina (with Hill's N2 = 4.62, 3.74 and 1.95 respectively) which occurred in these samples are currently restricted to a small number of lakes from Tasmania and one lake in western Victoria in the modern training set. This is nearly 2000 km south of North Stradbroke Island. The analysis of dissimilarity between fossil and modern assemblages using the modern analogue technique (MAT) shows that ten out of the 17 samples have ‘no good’ analogue in the modern calibration set (Fig. 5.4d). These include three samples from the top 277 cm and seven of the depths between 340 cm and 442 cm.

A total of 21 taxa were identified from the 17 samples (Fig. 5.3). Each of the chironomid taxa that have marked temperature preferences in south eastern Australia were highlighted and discussed in Chapter 4. The Holocene component of the record is intermittent and more samples in the c. ∼15.5– 23.2 cal ka BP range contained sufficient head capsules (>50) for quantitative reconstructions. The record is split into three zones based on the constrained incremental sums of squares (CONISS) analysis of the fossil chironomid assemblages (Fig. 5.5) and these zones correspond to the LGM (from the beginning of this record c. ∼23.2 to ∼18.1 cal ka BP) (Barrows et al., 2002; Clark et al., 2009), deglaciation (c. ∼18.1 to ∼13 cal ka BP) and the Holocene (c. ∼13 cal ka BP to near present). The break for the Holocene based on CONISS analysis is ∼13 cal ka BP for this record which is different from the well recognized 11.5 cal ka BP is likely due to the low sample resolution in this part of the record. These zones (Fig. 5.5) are consistent with the observed changes in the chironomid stratigraphy (Fig. 5.3).

5.6.2 The LGM (~23.2 cal ka BP – 18.1 cal ka BP)

The LGM (as defined in Barrows et al., 2002 and Clark et al., 2009) from Welsby Lagoon is characterised by the presence and increased abundance of cold indicators such as Parakiefferiella morphotype 3, Paralimnophyes morphotype 1, Tanytarsus glabrescens type and Paralimnophyes morphotype 3. The cooling also corresponds to an apparent decrease in the abundance of warm adapted taxa (Chapter 4, Fig. 4.5, Table 4.6) including Cladopelma, Parachironomus, Paratanytarsus, Tanytarsus lactescens type and Procladius (Fig. 5.3).

85 | P a g e

Application of the transfer function - late-Quaternary temperature reconstruction The reconstructed MFT at the start of the record (at c. ∼23.2–21.7 cal ka BP) is between 19.9 and 21.7 °C. This represents a cooling between 2.3 and 4.1 °C relative to the present (Fig. 5.4a). Decreasing inferred temperature is observed immediately after c. ∼21 cal ka BP. The mean reconstructed MFT at 20.6 cal ka BP is 18.4 °C which is 5.6 °C cooler than the present (Fig. 5.4a). Temperature continued to fall and by c ∼18.5 cal ka BP, an MFT minimum of 17.4 °C was obtained from this site which is a ∼6.5 °C cooling relative to the present day. A rapid warming trend is then observed, as MFT started rising from c. ∼18.1 cal ka BP and by 17.3 cal ka BP, it rose 3 °C to 20.4 °C (Fig. 5.4a).

5.6.3 The deglacial period (~18.1 cal ka BP – 13 cal ka BP)

During the deglacial, the reconstructed MFTs are warm and highly stable (Fig. 5.4a). This period is characterised by the high abundance of warm indicating taxa Cladopelma, Procladius, Paratanytarsus, Tanytarsus lactescens type, Pentaneurini, Tanytarsus lugens type while cold stenotherms, such as Paralimnophyes morphotype 1 and 3, Parakiefferiella morphotype 3 are absent during this time period (Fig. 5.3). The average temperature is between 20. 4 and 21.4 °C from ∼17.3 to ∼15.5 cal ka BP, where fluctuations between samples do not exceed ∼1 °C. This indicates that local MFT at this time is ∼2.6–3.6 °C cooler than the present.

One sample from the depth of 350–352 cm (c. ∼15.9 cal ka BP) shows a reconstructed MFT as low as 17 °C at the site (Fig. 5.4a). This sample is characterized by exceedingly high abundances of Dicrotendipes (31%) and Procladius (25%) (Fig. 5.3). No other warm or cold indicators (apart from warm stenotherms Pentaneurini and Tanytarsus lugens) are present in this sample (Fig. 5.3). The composition of this sample is very similar to the Holocene samples that are discussed below.

5.6.4 The Holocene (~13 cal ka BP – present)

The four Holocene samples at the depths of 55–57 cm, 85–87 cm, 257–259 cm and 275–577 cm, which represent c. ∼0.2, 2, 9.3 and 10.5 cal ka BP respectively, have assemblages dominated by one taxon – Dicrotendipes (at abundances of 65%, 37%, 62% and 66% in each of the four samples respectively). Dicrotendipes was widespread in mainland Australian training set and occurred in lakes at MFTs ranging from 12 °C to 24 °C (Chapter 4, Fig. 4.6). Chironomus sp. is the other dominant taxon in the Holocene samples and is also present in lakes spanning a long temperature gradient. Only a small percentage (<5%) of warm stenotherms composed the assemblages of these 86 | P a g e

Application of the transfer function - late-Quaternary temperature reconstruction samples. These include Cladopelma, Polypedilum nubifer, Paratanytarsus, Tanytarsus lactescens type, Pentaneurini, Tanytarsus lugens type (Fig. 5.3).

Quantitative MFT reconstructions of these samples provide generally lower values than expected (e.g Dimitriadis and Cranston, 2001) from this site for the Holocene (19.2 °C, 18.5 °C, 15.8 °C and 16.3 °C at the age of c. ∼0.2, 2, 9.3 and 10.5 cal ka BP respectively) (Figs. 5.3 and 5.4a). These show a cooling between 4.8 and 8.2 °C compared to the present in the Holocene at the site. However, previous studies show that Dicrotendipes at least in the Holarctic occurs in the littoral of lentic environments (Pinder and Reiss, 1983) where it is often associated with macrophytes (Brodersen et al., 2001) and mesotrophic to eutrophic waters (Brodin, 1986). The temperature reconstructions of these samples are highly dependent on the abundance of one taxon, Dicrotendipes, which is discussed below and this factor may cause inaccurate MFT reconstructions in these four samples.

87 | P a g e

Application of the transfer function - late-Quaternary temperature reconstruction

Figure 5.3 Chironomid stratigraphy from Welsby Lagoon plotted with respect to the radiocarbon chronology. Chironomid inferred mean February temperatures (MFTs) are shown in (a) and the number of head capsules counted for each sample are shown in (b). The dashed line in (a) indicates the current mean February temperature at Welsby Lagoon and the dashed line in (b) indicates the minimum number of head capsules for a quantitative reconstruction from a sample. All fossil taxa present in the Welsby Lagoon record are included in the diagram. Age in calibrated years is on the y-axis and taxa were grouped according to their optima in the modern training sets based on Chapter 4. Grey areas represent sediment layers where chironomids were absent. Note that Dicrotendipes, a benthic taxon that does not have a significant correlation (p > 0.05) to MFT in the modern calibration data set, is dominant in all four Holocene samples. Each of the taxon was photographed and included in Appendix 3.

88 | P a g e

Application of the transfer function - late-Quaternary temperature reconstruction

5.6.5 Relationships between changes in chironomid assemblages and temperature and other variables

Passive plots of the fossil samples (Fig. 5.6) in a canonical correspondence analysis (CCA) of the modern chironomid assemblages against the significant (p < 0.05) environmental variables (MFT, depth, Total Phosphorus (TP), Chlorophyll a (Chl a), Conductivity, pH) indicate the directional change of trend of fossil assemblages through time. In Fig. 5.6, sample numbers 1–33 represent modern sites from Chapter 4, Table 4.1 and sample 34 is Swallow Lagoon. Fossil samples are labelled by their respective ages in Fig. 5.6a and b respectively. Changes in trend of fossil assemblages through most of the deglacial and LGM samples (Fig. 5.6a and c. ∼23.2–16.6 cal ka BP) follow and are parallel to the mean February temperature (MFT) gradient, except for two samples (Fig. 5.6a, at c. ∼17.8 and c. ∼23.2 cal ka BP) that are possibly influenced by the secondary gradient. The trend of changes of the fossil assemblages in the deglacial through to the Holocene samples (Fig. 5.6b and c. ∼16.3–0.2 cal ka BP) are parallel to the secondary gradient that is driven by changes of lake depth, nutrient variables (TP, Chl a), conductivity and pH. It was impossible to distinguish among these four driving variables, since these parameters are highly correlated in the south eastern Australian modern lake training set (see Chapter 3 and Chapter 4). This implies that the MFT reconstructions from the LGM and the deglacial samples are reliable as the trend of changes in the fossil assemblages is less influenced by a secondary gradient. While only one sample from c. ∼16.3–0.2 cal ka BP does not have a good fit to temperature, MFT reconstructions may still be inaccurate through this period. This is due to the fact that the trend of changes in fossil sample assemblages is overwhelmed by the changes in variables that are driven by the secondary gradient.

89 | P a g e

Application of the transfer function - late-Quaternary temperature reconstruction

Figure 5.4 Reconstruction diagnostics for Welsby Lagoon samples. (a) Reconstructed mean February temperatures (MFTs) with the RMSEP (2.2 °C) from the modern training set (Chapter 4) plotted in dashed horizontal lines for each sample. (b) Goodness-of-fit statistics where the fossil samples are passively fitted to the CCA ordination axis derived from the modern training set constrained to MFT. The vertical dashed line represents the 95th percentiles of modern squared residual lengths beyond which fossil samples are considered to have poor fits to MFT. (c) Rare taxa plot where closed circles represent fossil samples with well-represented taxa in the modern training set and open circles represent fossil samples with rare taxa (Hill's N2 ≤ 5) summing to greater than 10% abundance. (d) Non-analogue plot where closed circles indicate fossil samples with either close or good modern analogues, whereas open circles indicate fossil samples with no good modern analogues (at the 5th percentile cut-off level). The vertical lines represent the 2nd, 5th and 10th percentiles respectively.

90 | P a g e

Application of the transfer function - late-Quaternary temperature reconstruction

Figure 5.5 The Welsby Lagoon chironomid record was split into three zones using the constrained incremental sums of squares (CONISS) function in Psimpoll (Bennett, 2002). W-1 represents the last glacial maximum (LGM), W-2 represents the deglacial and W-3 represents the Holocene. This zonation is consistent with the observed changes in chironomid assemblages shown in Fig. 5.3.

91 | P a g e

Application of the transfer function - late-Quaternary temperature reconstruction

a. b.

Figure 5.6 Trajectory changes of trend through time of fossil samples (black circles) in the Welsby Lagoon record, passively plotted in a CCA of the training set lakes (open circles, where site 1–33 are waterbodies from (Chapter 4, Table 4.1) and site 34 is Swallow Lagoon, a waterbody near Welsby Lagoon with similar physical characteristics) against significant variables (p < 0.05). (a) Samples covering the last glacial maximum (LGM) and deglacial period, where changes in trend are parallel to the mean February temperature (MFT) gradient, indicating that MFT is a primary controlling variable and drives the changes of chironomid assemblages in the samples from c. ∼23.2 to 16.6 cal ka BP. (b) samples covering the period of deglacial and Holocene from c. ∼16.3 to 0.2 cal ka BP. The trend of changes of chironomid assemblages in these fossil samples are parallel to the secondary gradient that is represented by nutrients, conductivity, pH and lake depth, indicating that non-climatic factors are likely dominant in Holocene changes. 92 | P a g e

Application of the transfer function - late-Quaternary temperature reconstruction

5.7 Discussion

5.7.1 Precision and reliability of the reconstruction

There are three factors that may complicate the temperature reconstruction and may have influenced the reliability of the record. First, in the Holocene samples and one sample at 15.9 cal ka BP, the dominance of one taxon, Dicrotendipes, which has a lentic habit and is highly correlated to macrophyte abundance in the lake (Cranston and Dimitriadis, 2004), led to a lower than expected reconstruction of temperature for the Holocene (4.8 to −8.2 °C compared to present). Dicrotendipes is neither a warm nor a cold stenotherm and it does not show a significant correlation (p > 0.05) to temperature in the modern training set of Australian lakes based on the Generalized Linear Model (GLM) results (Table 4.6 in Chapter 4). Instead it has a significant correlation to nutrient variables, conductivity and lake depth. In a study of 21 Danish lakes Brodersen et al. (2001) found that Dicrotendipes has a high co-occurrence with macrophytes and is found predominantly in shallow and eutrophic lakes. Luoto (2009) also points out that Dicrotendipes is not distributed along a temperature gradient in the 77 lakes of a modern training set from Finland. Consequently, the temperature inferences based on Dicrotendipes as a warm stenotherm in some cases (e.g. Chase et al., 2008 in cold temperate Canadian lakes) is probably an artefact, although, some studies (e.g. Dimitriadis and Cranston, 2001; using mutual-climate-ranges technique from eastern Australian lakes) recognized that Dicrotendipes can be a warm temperature indicator. In this data set, the presence of Dicrotendipes in eutrophic lakes in southern areas eliminates any association with warmth.

Secondly, changes in nutrients, conductivity, pH and lake depth may all have had an impact on chironomid assemblages through time. However, most of the deglacial and LGM samples followed the MFT gradient closely (Fig. 5.6a), suggesting these samples are not strongly influenced by variables other than MFT changes. Fig. 5.6b shows the lagoon had experienced large fluctuations in lake water chemistry from the deglacial throughout the Holocene (c. ∼16.3–0.2 cal ka BP) and these changes are characterised by significant shifts in pH and the concentration of nutrients, salts and fluctuations of lake level. Most of these samples are also coincident with the high occurrence of Dicrotendipes (Fig. 5.3). In contrast, more stable lake conditions were observed during the early deglaciation and the LGM (c. ∼23.1–16.6 cal ka BP) (Fig. 5.6a). Welsby Lagoon had a relatively more consistent water level and was not significantly influenced by the secondary gradient during

93 | P a g e

Application of the transfer function - late-Quaternary temperature reconstruction this period of time compared to the Holocene. MFT reconstructions from these time intervals are consequently regarded as reliable.

Finally, the modern calibration training set contains a relatively small number of lakes (33) and this may be the cause of the ‘no good’ modern analogues situation for a number of samples in the deglacial and LGM as larger training sets generally allow better estimates of species' optima and tolerances in relation to important environmental variables (Lotter et al., 1999). A similar situation was observed in the temperature reconstructions of Eagle Tarn and Platypus Tarn from Tasmania (Rees and Cwynar, 2010). However, this is less problematic than in some reconstructions (Heiri and Millet, 2005; Larocque-Tobler et al., 2010), where an abundant taxon (Corynocera ambigua) was present in a large number of fossil samples and was fully absent from the modern training set. In many cases, samples with ‘no good’ modern analogues still provide reliable and robust trends in temperature reconstructions (e.g. Lotter et al., 1999; Rees and Cwynar, 2010) because the (WA) PLS model performs relatively well in no analogue situations (Birks, 1998; Lotter et al., 1999) and, consequently, the temperature reconstructions are considered reliable from the deglacial and LGM samples. Merging a regional training set with Rees et al. (2008) may increase the likelihood of finding good modern analogues for fossil assemblages from Welsby Lagoon.

Overall, the transfer function model has the tendency to underestimate temperatures at warm sites and overestimate from colder sites (see Chapter 4), which is not unusual for a model of this type (Barley et al., 2006). Welsby Lagoon (MFT = ∼24 °C at the present) is located among the warmest sites of the modern calibration set (Fig. 4.7 in Chapter 4), and consequently the reconstructed mean February temperatures are likely to be under-predicted.

Finally, one of the more serious limitations of the transfer function is the relatively large RMSEP of 2.2 °C. Nevertheless, the trends between the samples are likely to be very robust, possibly due to the consistency of local setting (Lotter et al., 1999). In addition, we note that more than two thirds of the modern calibration sites fall with MFTs between 18 °C and 24 °C and since most reconstructed temperature estimates fall within this envelope (Chapter 4, Table 4.1), this reinforces the suitability of the model for application to Welsby Lagoon data.

In summary, the maximum cooling observed in our data set occurred at c. ∼18.5 cal ka BP with a reconstructed MFT of between 15.2 and 19.6 °C. This gives a range of cooling between 4.4 and 8.8 °C but with cooling considering the errors likely to be close to the midpoint estimate at about 6.5 °C. This cooling times well with Antarctica (Stenni et al., 2011) and is also compatible with the late 94 | P a g e

Application of the transfer function - late-Quaternary temperature reconstruction termination of glaciation in Australia (e.g. Barrows et al., 2002; Hughes et al., 2013). It is clear that most of the LGM and all of the deglaciation period were much warmer than this maximal cooling. Pollen work from Welsby Lagoon by Moss et al. (2013) shows rainforest taxa (mainly Araucaria and palms) present through the LGM. A cooling of c. ∼6.5 °C would change the composition of the local rainforest but would not of itself eliminate rainforest from the site. Consequently, these findings are compatible with the persistence of rainforest on North Stradbroke noted in Moss et al. (2013).

95 | P a g e

Application of the transfer function - late-Quaternary temperature reconstruction

Figure 5.7 Chironomid mean February temperature (MFT) transfer function based reconstruction 18 from Welsby Lagoon (WEL) (in plot d) compared to (a) the δ O and CO2 records from Antarctica

96 | P a g e

Application of the transfer function - late-Quaternary temperature reconstruction (Stenni et al., 2011; Pedro et al., 2012); (b) insolation curve for seasonality (Berger and Loutre, 1991); (c) δ18O measured from G. ruber from marine cores along the East Australian Current (EAC) and (d) temperature estimated based on branched glycerol dialkyl glycerol tetraether (GDGT) from Fraser Island (Woltering et al., 2014). The chironomid results from WEL (d) indicate a maximum cooling relative to the modern MFT along the Southeast Queensland coast of c. ∼6.5 °C with much lower cooling at other times during the late LGM and a rapid recovery at about 18.1 cal ka BP. The cooling is significantly less than the 9–12 °C for the adjacent uplands from periglacial landforms from Galloway (1965), but is in line with estimates from nearby marine cores shown in (c) and lacustrine record from Lake Mackenzie (LM), Fraser Island shown in (d) (note the reconstructed temperature from Fraser Island is lower than Welsby Lagoon because the GDGT is based on mean annual temperatures calibration whereas chironomid transfer is based on MFT). It is also noteworthy that the chironomids provide an estimate of summer temperatures whereas the periglacial landforms are indicators of winter temperature. This suggests a larger seasonal temperature contrast at the LGM which is consistent with the insolation curves (b). Finally the 18 observed timing and pattern is very similar to both the δ O curve and CO2 record from Antarctica (a) (Stenni et al., 2011; Pedro et al., 2012).

5.7.2 The deglacial and LGM temperatures of North Stradbroke Island and regional climate

To further examine this new seasonal temperature reconstruction at the LGM, we compared our chironomid-based mean February temperature record to Antarctica ice cores (TALDICE: Stenni et 18 al., 2011 for δ O isotope and Byrd: Pedro et al., 2012 for CO2 concentrations) (Fig. 5.7a), summer/winter insolation curves (Fig. 5.7b , Berger and Loutre, 1991), marine reconstructions from near the EAC (Troedson and Davies, 2001; Dunbar and Dickens, 2003; Page et al., 2003; Bostock et al., 2006) (Fig. 5.7c) and a lacustrine temperature record from Lake Mackenzie, Fraser Island (Woltering et al., 2014) (Fig. 5.7d) respectively.

Previous terrestrial proxies (block stream, AAR and most recently branched glycerol dialkyl glycerol tetraether (GDGT)) used in mainland Australia to estimate temperatures have not been widely applied. Pollen based transfer functions for temperature reconstructions in mainland Australia were tested (Cook and Van der Kaars, 2006) but have not been successfully applied to reconstruct LGM temperatures, in contrast to the Tasmanian pollen transfer function (Fletcher and Thomas, 2010). For our study, the recent GDGT record from Lake Mackenzie (LM) (25°26′51″S, 153°03′12″E, 90 m asl), Fraser Island (Woltering et al., 2014) provides the best comparison

97 | P a g e

Application of the transfer function - late-Quaternary temperature reconstruction although the record has a hiatus between 18.3 and 14 cal ka BP, which coincides with the time of best detail in our study. At Lake Mackenzie, the lowest mean annual temperature occurred at 18.8 ± 0.5 cal ka BP, with MAT estimated to be ∼4.1 °C lower than modern day temperature (Fig. 5.7d). The timing of this maximum cooling is in good agreement with our Welsby Lagoon record (Fig. 5.7d), and is also consistent with the offshore SST estimates based on G. ruber δ18O values from the north (GC-12, 23°34.37′S, 153°46.94′E; Bostock et al., 2004, 2006), and south (GC-25. 26°35′S, 153°51′E ; Troedson, 1997; Troedson and Davies, 2001) of Fraser Island, and further south (GC-9, 33°57′S; 151°55′E; Troedson, 1997; Troedson and Davies, 2001) along the East Australian Current (EAC) (Fig. 5.7c). Our record shows good agreement with the long term temperature trends with marine records GC-25 and GC-9 (Fig. 5.7c) and is particular similar to the reconstructed SST from GC-25 (at 26°35′S) (Bostock et al., 2006) (Figs. 5.1 and 5.7c), which suggested a ∼6 °C of mean annual SST cooling. All these records show a maximum cooling at c. ∼18.5 cal ka BP, followed by rapid warming to (near) Holocene values by c. ∼17.3 cal ka BP (Fig. 5.7c and d). Lastly, the timing of this cooling from all these records show a coincidence with the AIM2 event, which marked the start of deglaciation at 18.2 ± 0.7 cal ka BP (Stenni et al., 2011) represented in the δ18O TALDICE ice core record (72°49′S, 159°11′E) (Fig. 5.7a) and the increase of CO2 concentration estimated from Byrd ice core record (80°S, 119°49′W), Antarctica. The TALDICE and Byrd records also indicate a rapid warming trend observed immediately after the maximum cooling, similar from Welsby lagoon and the southern marine sites (GC-25 and GC-9) near the EAC.

The scale of cooling in the Fraser record (4.1 °C, Woltering et al., 2014) is less than we observe at Welsby Lagoon (6.5 °C) but the results are within the errors of the respective techniques and it is possible that GDGT data also reflect summer temperatures, assuming that the microbes producing branched GDGTs are most active during the period of peak photosynthetic activity in summer (Pearson et al., 2011). However, though Welsby Lagoon is only 2° of latitude south of Fraser Island, it is possible that the difference reflects changes in the EAC offshore as the degree of summer cooling from Welsby Lagoon and both of these are similar to a southern site GC-9 (at 33°57′) (Figs. 5.1 and 5.7c). In contrast, the Fraser island record is closer to the reconstruction from the northern site GC-12 (23°34′S) (Figs. 5.1 and 5.7c). Bostock et al. (2006) interpreted the difference of SST amplitude (∼3–4 °C temperature change across ∼3° of latitude) of the LGM between cores GC-12 (23°34′S) and GC-25 (26°35′S) as reflect a northwards shift in the separation of the Tasman Front (Fig. 5.1) from the EAC during the LGM, resulting in warm waters from the EAC not reaching as far south as GC-25 and Welsby Lagoon. More marine and terrestrial proxy data from the LGM are required from subtropical Australia to allow this hypothesis to be further tested. 98 | P a g e

Application of the transfer function - late-Quaternary temperature reconstruction

The reconstructed cooling is substantially less than earlier terrestrial reconstructions from periglacial landforms (8–12 °C) (Galloway, 1965) and amino acid racemization (AAR) values from Emu egg shells (Miller et al., 1997). This does not necessarily imply a direct disagreement between the techniques as the chrionomids and the GDGT data may both reflect temperatures from summer months, whereas the periglacial landforms are interpreted as winter temperature estimates, and the AAR reconstructions come from semi-desert environment several thousand kilometres west and south of this region. These differences may be reflecting enhanced seasonality during the LGM, in which large temperature differences between summers and winters would be expected (Fig. 5.7b). In addition, it is unsurprising that coastal records resemble the thermal history of adjacent marine sites.

5.7.3 Wider paleoclimatic considerations

One of the key observations from the Welsby Lagoon chironomid record and other temperature records from subtropical Australia (including marine records), is that whatever the scale of cooling, the timing of the deglaciation is synchronous with changes in the Antarctic (Fig. 5.7). This is noteworthy on two counts. Firstly, Welsby lagoon sits on the transition between the tropics and the temperate zone. The record clearly responds directly to changes at high latitudes and implies that high latitude forcing controls climate change at least as far north as 27°S. This is unexpected, because this is an austral summer month reconstruction and at the present day, while winter climate is affected by the mid-latitude westerlies, modern summer climate is dominated by tropical air masses (Ledru and Stevenson, 2012). The possibility that tropical air masses did not penetrate this subtropical area in the summer during the glacial period needs to be considered. Secondly, there is a disconnection between the temperature trend we observe and insolation in the Southern Hemisphere. The last glacial maximum (LGM) coincides with a summer insolation maximum while the deglaciation corresponds to a period of reduced summer insolation (Fig. 5.7 b–d). This suggests that regional insolation trends are overwhelmed by hemispheric changes driven from high latitudes. Complex feedback mechanisms through either greenhouse gases or the thermohaline circulation are implied.

99 | P a g e

Application of the transfer function - late-Quaternary temperature reconstruction 5.8 Conclusion

A summer temperature reconstruction based on the recently developed chironomid transfer function from southeastern Australia (developed in Chapter 4) is presented from Welsby Lagoon of North Stradbroke Island. This is the first detailed seasonal reconstruction of LGM and the last deglacial temperatures from terrestrial Australia. Factors that may affect the reliability of the inferred temperatures include changes in lake chemistry and lake level, which had an impact on chironomid assemblages, particularly in the Holocene. The relatively small modern calibration training set also limits the ability to fully understand the environmental controls on some of the fossil samples. Despite these caveats, the temperature reconstruction from the LGM to the deglacial period of this site is sensible and a stable lake condition (in relation to lake level and lake chemistry changes) is observed. The overall reconstruction displays strong similarities to marine records from areas affected by the East Australian current (EAC), both in terms of timing and magnitudes of cooling during and immediately after the LGM.

The results show a maximum mean February temperature cooling of c. ∼6.5 °C at the late LGM at c. ∼18.5 cal ka BP, followed by rapid warming to (near) Holocene values by 17.3 cal ka BP. This warming is consistent with the AIM2 event, which marked the start of deglaciation at 18.2 ± 0.7 cal ka BP (Stenni et al., 2011). It is also compatible with the increase of CO2 concentration estimated from the Byrd ice core record in West Antarctica (Pedro et al., 2012). It suggests that Antarctic climate influence reached the Australian subtropics during the LGM.

100 | P a g e

Stable isotope analyses of the chironomid head capsules

6 CAN STABLE OXYGEN AND HYDROGEN ISOTOPES FROM AUSTRALIAN SUBFOSSIL CHIRONOMID HEAD CAPSULES BE USED AS PROXIES FOR PAST ENVIRONMENTAL CHANGE?

Chapter 6 is submitted to Limnology and Oceanography

6.1 Summary

This chapter reports the investigation of the potential of using stable oxygen (δ18O) and hydrogen (δ2H) isotopes of subfossil chironomid head capsules as independent proxies for past climate and environmental reconstructions from southeastern Australia.

Highlights:

 This is the first of such study in the Australasian region and among one of the very few similar investigations worldwide.  Previous studies have focussed on the relationship of temperature and δ18O of chironomid head capsules. The effects of in-lake variables were overlooked.  This work suggests that nutrients and salinity are important to consider for both δ18O and δ2H analyses of chironomid head capsules, at least for the genus of Chrionomus, when applied as proxies for past changes.  This study concludes and re-affirms that both δ18O and δ2H are potentially valuable tools for reconstructing temperature in low nutrient and dilute lakes in humid areas.  This work is complementary to the previous work included in Chapter 3 and 4 but also completely independent.

101 | P a g e

Stable isotope analyses of the chironomid head capsules 6.2 Abstract

Results of stable isotopes δ18O and δ2H analysed on head capsules of chironomids (Chironomus spp.) from sixteen southeastern Australian lakes are presented. Lakes that are located in semi-arid and sub-humid areas have high evaporation, prolonged water residence time and experience modified local hydrological conditions including increased salinity and nutrient status and evaporative enrichment of δ18O values in lake water. These lakes do not appear suitable to be used for temperature calibration in paleoclimate reconstructions. High nutrient and salt concentrations create an ecologically challenging aquatic environment for chironomids, however, haemoglobin helps regulate the oxygen level in Chironomus spp. and allows this taxon to thrive in these environments. For these lakes, Chironomus spp. was not in equilibrium with δ18O lake water due to strong vital effects and we recommend against its use to infer climatic variables in these settings. For oligotrophic to mesotrophic lakes, the relationship between δ18O of Chironomus spp. and temperature appears robust (r2 = 0.78) and very similar to results from European lakes (e.g. Verbruggen et al., 2011). Similar results were also obtained for δ2H and temperature, with an r2 of 0.72. The overall findings of this study are that both δ18O and δ2H are potentially valuable tools for reconstructing temperature in cooler, low nutrient and low salinity regions of Australia. Chironomus spp. is an obvious target for stable isotope work because it is both large and ubiquitous but its ability to adapt to low oxygen conditions makes it a poor climate indicator in non-humid settings.

102 | P a g e

Stable isotope analyses of the chironomid head capsules 6.3 Introduction

Development of proxies that have the potential to provide long-term and reliable palaeoclimate records is a key to reconstructing and understanding past climate systems. Ideally these proxies should be widely distributed and common in the environment. Chironomids (Diptera: Chironomidae) are non-biting midges whose larvae occur in virtually all permanent and semi- permanent terrestrial water bodies (Cranston, 1995). The fossilized head capsules of chironomid larvae are preserved in lake sediment and the structure of these chitinous exoskeleton fragments is usually well preserved (Walker, 1987). The growth of chironomid larvae is strongly controlled by water temperature (Armitage, 1995). Based on these characteristics, transfer functions based on chironomid species assemblages have been developed from various parts of the world for paleoclimate and environment inferences (e.g. Lotter et al., 1999; Larocque et al., 2001; Woodward and Shulmeister, 2006; Walker and Cywnar, 2006; Rees et al., 2008; Massaferro et al., 2014 and Chang et al., 2015a – Chapter 4). These have been successfully applied to Late Quaternary palaeoclimatic data and have provided quality quantitative seasonal temperature reconstructions (e.g. Woodward and Shulmeister, 2007; Rees and Cwynar, 2010; Samartin et al., 2012; Chang et al., 2015b – Chapter 5).

In addition to the traditionally applied chironomid-based transfer functions, other studies (e.g. Wooller et al., 2004, 2008; Gröcke et al., 2006) have shown that the δ18O value of chironomid and aquatic beetle chitin is largely a reflection of lake water δ18O which correlates with the δ18O of local precipitation and therefore, has the potential to be used for regional paleoclimate reconstructions. More recently, δ18O variations in subfossil chironomid head capsules were applied to quantitatively characterize the late glacial lake water δ18O changes and, indirectly, past air temperature changes near Rotsee, Switzerland (Verbruggen et al., 2010b).

There are two key assumptions that need to be considered before stable isotope measurements from chironomid head capsules in lake sediment can be used as proxies for climate. Firstly, it is often assumed that the modern and past isotopic composition of lake water reflects mean annual precipitation (Rozanski et al. 1993). While this is generally true in humid regions, evaporation affects the stable isotope composition of water via evaporative enrichment of the heavy isotope (Leng et al., 2006) and this assumption needs to be tested for sub-humid and semi-arid areas. In addition, the lake examined needs to be volumetrically large enough but un-stratified for its isotope composition to reflect the average isotope values in mean annual precipitation (Leng et al., 2006). But, stratification may protect a large portion of the water from evaporative enrichment 103 | P a g e

Stable isotope analyses of the chironomid head capsules (Verbruggen et al., 2011). Lake water δ18O values may approximate those of precipitation in stratified lakes if there is at least one mixing event per year.

Secondly, it is sometimes assumed that the isotopic composition of the head capsules is in equilibrium with the lake water and that there are no vital and micro-environmental effects that might modify head capsule isotope values. Vital effects could be induced by chironomid metabolic processes and diet. Such effects have been examined in foraminifera (Erez, 1978), (von Grafenstein et al. 1999), and cocoliths (Ziveri et al., 2003) but there has been very little work on chironomid vital effects (Grey et al, 2004a, b; van Hardenbroek et al., 2014).

In an attempt to address these issues in chironomids, Wang et al (2009) quantified how the δ18O and δ2H of water and diet influenced the δ18O and δ2H of chironomid larvae. Their results revealed that both water and diet affect the δ18O and δ2H isotope composition of chironomid larvae, whereby ~70% of the oxygen in the total organic composition is derived from the water of the larval habitat. In contrast, diet dominates the hydrogen isotope ratios of chironomid larvae (~70%), and only 30% of hydrogen in the chironomid larvae is derived from the ambient water (Wang et al. 2009). However, this laboratory experiment is based on feeding the chironomid larvae powdered Spirulina algae as controlled diet. In natural aquatic environments, the total proportion of hydrogen in the chironomid larvae that are derived from ambient water could be higher if indirect uptake of δ2H in the natural diet is considered. Their results also suggested that the majority of the hydrogen in lipids, protein and keratin are derived from diet rather than water, and these indicated vital effects may have a large influence on the δ2H values and to some degree on the δ18O values of chironomids. Differences in metabolic pathways also likely influenced δ18O and δ2H fractionation in different taxa.

Here we present the first study of stable isotope (δ18O and δ2H) analyses on chironomid head capsules from southeastern Australian lakes. Based on previous research (Wang et al., 2009 ; Grey et al., 2004a, b; van Hardenbroek et al., 2014), it is likely that different chironomid taxa exhibit differences in stable isotopic composition (via vital effects) and will thus display different relationships to climatic or environmental variables. We therefore performed the δ18O and δ2H analyses on head capsules from a single genus (Chironomus spp.) because it is the most abundant genus in southeastern Australian lakes (Chapter 4). We investigate the relationship between the stable isotope composition (δ18O, δ2H) of Chironomus spp. head capsules and how this relates to the composition of lake water and local precipitation. The main goal of this study was to examine the

104 | P a g e

Stable isotope analyses of the chironomid head capsules feasibility of the application of stable isotopes of Chironomus spp. head capsules as a proxy to reconstruct past changes in climate in southeastern Australia.

6.4 Materials and Methods

6.4.1 Study sites and sample collection

Fifteen natural lakes and one reservoir located in southeast Australia (Table 6.1, Fig. 6.1) were included in this dataset. The dataset covers the latitudes between 30˚S and 41.6˚S (Table 6.1, Fig. 6.1) and elevation of the sites ranged from sea level to c. ~2000 m above mean sea level (a.s.l) (Table 6.1). The sixteen waterbodies included in this dataset had a large spatial distribution across eastern Australia (Table 6.1). One lake (LLL) was located on top of New England Tablelands (Fig 6.1), which lies in the subtropics but maintains a cool temperate climate due to high elevation (Table 6.1). Summer is the dominant season for the source of precipitation for this site. Three lakes were from Mt Kosciusko in the Australian Alps (Fig. 6.1), which is the coolest and highest (Table 6.1) area of Australia and the precipitation (with significant snowfall) is winter and spring dominant. Ten lakes were from western Victoria, where a true Mediterranean climate exists with warm dry summers and cool wet winters. The region consists of semi-arid to sub-humid regions and the lakes are highly susceptible to both natural and agricultural related eutrophication and salinization (as concluded in Chapter 3). Two lakes were from northwestern Tasmania (Fig. 6.1) and this is the region that is humid and precipitation is winter and spring dominant (similar to Australian Alps). Detailed descriptions of the climate, vegetation and geology of the study area were presented in Chapter 3.

All lakes were sampled during the summer (January and February) of 2012 and 2013 (Table 6.1). A minimum of three sediment cores were taken using a Glew Mini Corer (Glew 1991) at the geographical centre of each lake. The top 2 cm of each core were extruded on site and packaged at 0.5 cm intervals into Whirlpaks®. Lake water for stable isotope analyses was sampled from the location where the core samples were taken, at approximately 30 cm below the water surface. They were collected into pre-washed and labelled polyethylene bottles. All bottles were sealed to prevent evaporation. Sediment and water samples were refrigerated until analysis.

Lake water samples were collected at the same time for the analysis of the concentrations of major ions, total nitrogen/total phosphorus (TN/TP), Chlorophyll a, pH and specific conductance

105 | P a g e

Stable isotope analyses of the chironomid head capsules (COND). In the field, water temperature, oxidation reduction potential (ORP), dissolved oxygen (DO), total dissolved solids, salinity and turbidity were also recorded using an Aquaread multi- parameter meter (AQUAREAD, Kent, UK). Detailed methodology for water chemistry analyses was outlined in Chapter 3 and analytical results are presented in Table 6.1.

6.4.2 Climatic and stable isotopes in precipitation data

Climate variables were obtained using the combination of the WorldClim program (available from http://www.worldclim.org/bioclim, accessed 20 January, 2014) and ArcGIS 10.1. WorldClim data for Australia is based on climate surfaces derived from around 600 nation-wide weather stations that have climate records spanning the years 1950–2000 (http://www.bom.gov.au/climate/data/stations/, accessed 20 January, 2014). For this study, only mean annual temperature (MAT) and precipitation (Precip) were considered (Table 6.1), because the spatial variation in δ18O in precipitation is usually strongly related to MAT along latitudinal gradients (Rozanski et al. 1993). The development of chironomid larvae in lowland Australia can span seasons of spring, summer and autumn, therefore the incorporated chironomid chitin may reflect annual average isotopic variations. Potential evapotranspiration (PET) and Aridity Index (AI) values were obtained from the Global Potential Evapo-Transpiration (Global-PET) and Global Aridity Index (Global-Aridity) dataset (CGIAR-CSI, available from http://www.cgiar- csi.org/data/globalaridity-and-pet-database, accessed 20 March 2015) (Table 6.1). Stable isotopes in precipitation data were obtained from the Global Network of Isotopes in Precipitation (GNIP) data set (IAEA/WMO, 2015) and interpolated using ArcGIS 10.1. for each site (Table 6.2).

6.4.3 Stable isotope sample preparation

Sediment treatment and stable isotope sample preparation were performed in the School of Geography, Planning and Environmental Management, The University of Queensland laboratories. Both sample preparation protocols developed and used in Wang et al. (2008) and Verbruggen et al. (2011) were considered and employed with some modifications. Sediment samples were deflocculated in cold 10% solution of potassium hydroxide (KOH) for two hours at room temperature and subsequently sieved with a 180 μm mesh-size sieve (van Hardenbroek et al. 2010). The top 2 cm was processed for chironomid head capsules from three surface core samples. Sieved residues were stored in vials with distilled water. Head capsules of the genus of Chironomus (Chironomus spp.) (Fig. 6.2), were picked from the sieved residue under a dissection microscope at 106 | P a g e

Stable isotope analyses of the chironomid head capsules 50 × magnification using fine forceps and placed into pre-labelled vials. Verbruggen et al (2010a) suggested that chemical pre-treatment of chironomid head capsules could affect the resulting δ18O values. Therefore, a physical treatment was used instead where after a minimum of 150 head capsules (average count is 185) was picked into the vials, they were placed in an ultrasonic bath for 20-30 seconds to separate contaminants. The head capsules were then examined for contaminants under the microscope and later transferred to a pre-weighed silver cup of 3.5 × 5 mm (Costech Analytical Technologies, INC., code: 041066). This was based on a modified two-step transfer protocol (Wang et al. 2008), which had an advantage of allowing any contaminants to be detached and manually separated during the process. Silver cups containing the head capsules were allowed to dry for several days in a covered Petri dish at room temperature. The silver cups were again weighed and, if a minimum of 100 µg was present, folded and shape-trimmed (Verbruggen et al. 2011). Samples were subsequently measured for stable oxygen (δ18O) and deuterium (δ2H) isotopes in Purdue Stable Isotope laboratory (PSI) facility, West Lafayette, USA.

Figure 6.1 Sample lakes are distributed across southeast Australia but there is a strong concentration (ten out of 16 lakes) in western Victoria. This reflects Chironomus spp. abundance in waterways with the tolerant Chironomus spp. strongly represented in the brackish and saline lakes of western Victoria.

107 | P a g e

Stable isotope analyses of the chironomid head capsules 6.4.4 Stable isotope analyses

Stable oxygen isotope analyses of the lake water samples were prepared by vacuum pump filtration and samples were placed in 10 ml Gas Chromatography (GC) vials. A volume of 0.3 µl aliquots were auto-injected on the PSI lab high Temperature Conversion Elemental Analyzer (TC/EA, Thermo Fisher Scientific) by a GC-PAL auto-sampler. The injected water sample was pyrolyzed under reducing conditions at 1400 °C to produce H2 and CO gases. These were separated chromatographically in a helium carrier gas stream and introduced sequentially into the ion source of an isotope ratio mass spectrometer (IRMS, Thermo Fisher Scientific) (Delta V Plus, TheremoFinnigan) for isotope ratio determination (Gehre et al., 2004). Six sequential injections were made for each sample and the reported values represent the average of three final injections (Nielson and Bowen, 2010). All lake water stable isotope data were calibrated using repeated analysis of three internal standards (PT, UT and PZ) relative to Vienna-Standard Mean Ocean Water and Standard Light Antarctic Precipitation (VSMOW/SLAP) (Coplen, 1995). All data were reported as per mil (‰) relative to the V-SMOW standard. Average uncertainties for lake water were ± 0.44‰ for δ18O and ± 3.6‰ for δ2H.

The Chironomus spp. head capsule samples were stored in the PSI facility at room temperature for seven days prior to analysis for the samples to equilibrate with the analytical environment. The TC/EA coupled to IRMS were used to determine ratios of stable isotopes for Chironomus spp. head capsules (sample weight > 100 µg). Samples were transferred to a Zero Blank Auto-sampler (Costech Analytical) interfaced with the TC/EA. They were pyrolyzed at 1400 °C in an oxygen-free environment to produce H2 and CO gases, which were chromatographically separated and introduced sequentially to the source of the IRMS (Gehre et al., 2004). Two blanks were measured at the start of every run. Three commercial keratin laboratory reference materials (powdered KHS, FH and UH) were used to normalize the results. Oxygen isotope δ18O data and δ2H values were calibrated against primary reference materials (Nielson and Bowen, 2010) and all data were reported as per mil (‰) relative to the VSMOW standard (Coplen, 1995). Average precisions of ± 0.8‰ and ± 4.0‰ for δ18O and δ2H of Chironomus spp. head capsules were achieved respectively.

108 | P a g e

Stable isotope analyses of the chironomid head capsules

Figure 6.2 Chironomus spp. head capsules from (a) Freshwater Lake, Victoria, 20 ×. (b) Lake Albina, Mount Kosciuszko, New South Wales, 10 ×. Photos of head capsules were taken by J. Chang at the Physical Geography Laboratory of School of Geography, Planning and Environmental Management, University of Queensland

6.4.5 Statistical analyses

Regression analyses and partial redundancy analysis (RDA) in CANOCO version 4.5 (ter Braak and Šmilauer, 2002) were used to test the correlation significance, regression coefficient and the independence of the variable correlations between the stable oxygen (δ18O) and hydrogen (δ2H) isotope data of Chironomus spp. head capsules against climatic and environmental variables. The purpose of these analyses was also to determine if fractionation or vital effects were constant or varied with changes in the environment.

Changes between isotope compositions of two substances are given as fractionation factor (α). The α for Chironomus spp. and lake water of oxygen (δ18O) and hydrogen (δ2H) isotopes is defined as:

Eq. 6.1

Eq. 6.2 respectively.

For testing correlations between fractionation factors (α) and environmental and climatic variables, the natural logarithm of α (lnα) was used by convention (Clark and Fritz, 1997) and was multiplied

109 | P a g e

Stable isotope analyses of the chironomid head capsules by 103 to adapt to the ‰ notation for δ values. Both regression and RDA (CANOCO version 4.5) were used to test the correlation significance, regression coefficient and the independence of the variable correlations between α[δ18O(Chironomus spp. - water)] and α[δ2H(Chironomus spp. - water)] against climatic and environmental variables, respectively.

6.5 Results

Stable oxygen (δ18O) and hydrogen (δ2H) isotope analytical results on head capsules of Chironomus spp. and the respective host lake water were obtained from sixteen southeastern Australian waterbodies (Table 6.2). Fractionation factors (α) between Chironomus spp. and host lake water were calculated applying Eq. 1 and Eq. 2 for δ18O and δ2H, respectively and results are presented in Table 6.2.

Figure 6.3 (a)Plot of δ18O of precipitation against δ18O of lake water. The δ18O of precipitation and lake water is correlated but significant enrichment of lake water is observed at δ18O of precipitation of around -5 ‰ V-SMOW, which the offset is possibly due to high regional evaporation rate and prolonged residence time or local hydrological conditions (b) the plot of δ18O of lake water against Chironomus spp. showed no signifcant correlation (P = 0.13) between the two data sets, suggesting that not all oxygen stable isotopic composition of Chironomus spp. head capsules were reflecting changes of their ambient water

110 | P a g e

Stable isotope analyses of the chironomid head capsules

Figure 6.4 (a) Plot of aridity index (AI) against δ18O of Chironomus spp.. AI has the best correlation and is the independent variable that explains the largest percent variance (Table 1) of Chironomus spp. δ18O data. At semi-arid and sub-humid areas (e.g. when AI <1), the variation in δ18O of Chironomus spp. data is large. (b) Total Nitrogen (TN) against fractionation factor

(α[δ18O(chiro-water)]). TN explains an independent and large percent of variance in the fractionation factor between Chironomus spp. and host lake water (Table 6.2), suggesting lake eutrophication is an important consideration for the interpretation of stable isotope data for Chironomus spp.

6.5.1 Stable oxygen isotope (δ18O) analyses results

The annual averaged of δ18O of precipitation varied between -4.65 and -7.20‰ and was positively correlated (r2 = 0.7) with the lake water δ18O (Fig. 6.3a), which varied between -6.17 and 12.97‰. Lake water δ18O in a number of western Victorian lakes was significantly enriched relative to precipitation (Table 6.2, Fig. 6.3a). The measured δ18O values from the chitinous head capsules of Chironomus spp. from the surface sediments of all sixteen southeastern Australian lakes showed no significant correlation (p > 0.05) with lake water isotopic composition (Fig. 6.3b).

18 Regression analyses of δ O of Chironomus spp. values against climatic and environmental variables showed that five variables were significantly correlated (p < 0.05). These were mean annual temperature (MAT), Aridity index (AI), potential evapotranspiration (PET), pH and specific conductance (COND). Redundancy analyses (RDAs) were then performed using the δ18O of Chironomus spp. values against these five variables (Table 6.3). Results showed that AI was the variable that explained the largest percent (53.7%) of the total variance in the δ18O Chironomus spp. 111 | P a g e

Stable isotope analyses of the chironomid head capsules data, although it did not retain its significance when MAT was partialled out (Table 6.3). However, the RDA on MAT showed that when AI was partialled out, it only explained 0.9% of the total variance (Table 6.3). AI and MAT are correlated and this is due to the effect of temperature on evaporation of the lake waters. AI includes a potential evapotranspiration (PET) component which is dependent on MAT, as PET is primarily controlled by temperature (MAT). However, AI was selected because it has the most powerful explanatory of variance in the δ18O of Chironomus spp. data. The regression plot of δ18O of Chironomus spp. against AI showed a relatively strong positive correlation (r2 = 0.54, p < 0.05) (Fig. 6.4a). The variation in the δ18O of Chironomus spp. values was large for sites with an AI value of less than one (i.e. semi-arid and sub-humid sites) (Fig. 6.4a).

Figure 6.5 (a)Plot of δ2H of precipitation against δ2H of lake water. The δ2H of precipitation and lake water is closely correlated, with r2 = 0.92 (b) Plot of δ2H of lake water aginst δ2H of Chironomus spp.. δ2H of lake water and Chironomus spp. showed a close correlation (r2= 0.69) suggesting that a large proportion of δ2H stable isotopic composition of Chironomus spp. head capsules that were directly derived from their ambient water or indirectly from feeding on planktonic algae, where algae also contain δ2H stable isotopes that are derived from lake water.

112 | P a g e

Stable isotope analyses of the chironomid head capsules Table 6.1 Summary of information available for the 16 lakes included in this study. The complete data-set is available in Chapter 3.

Nutrient No Lake Name Lake Coordinates ALT Depth PET Precip AI MAT TP TN Chl a pH COND Trophic status Limitation m m mm/Year mm/Year ˚C mg/L mg/L µg/L - µs/cm

Little S30.09 1 Llangothlin LLL 1361 3 1142 944 0.83 11.6 0.16 1.5 31 8.11 212 N,P Eutrophic E151.78 Lagoon S36.41 2 Blue Lake BL 1901 28 724 1708 2.36 3.9 0.007 0.11 2 6.44 5 N,P Oligotrophic E148.32 S36.42 3 Lake Albina LA 1919 9 728 1693 2.33 4.4 0.007 0.12 1 6.6 6 N,P Oligotrophic E148.27 Lake S36.46 4 CTL 2048 3 673 1691 2.51 3.7 0.011 0.15 1 6.47 5 N,P Oligotrophic Cootaptamba E148.26 Nuggety S37.00 5 Gully NGR 221 1.2 801 537 0.67 14.1 0.041 1.6 11 7.02 400 P Eutrophic E143.73 Resrvoir S37.14 6 Lake Fyans LFY 219 2.7 1367 595 0.44 13.8 0.019 0.64 3 8.79 206 P Eutrophic E142.62 Freshwater S37.59 7 FWL 227 2.1 802 660 0.82 13.2 0.56 5.1 60 6.74 564 N,P Hypereutrophic Lake E142.32 Lake S37.98 8 LTK 167 3 1539 591 0.38 13.4 0.12 2.2 1 9.17 2470 P Hypereutrophic Tooliorook E143.28 Lake S38.06 9 LSP 107 6 1208 763 0.63 13.3 0.031 0.88 9 8.08 698 P Eutrophic Surprise E141.92 Lake S38.13 10 LMB 21 4.8 1096 793 0.72 13.8 0.021 0.79 5 8.03 1140 P Eutrophic Mombeong E140.18 Lake S38.14 11 LTP 133 0.5 1239 649 0.52 13.5 3.6 36 90 8.68 15700 N,P Hypereutrophic Terangpom E143.33 S38.21 12 Swan Lake SWL 23 1.1 1058 797 0.75 13.7 0.044 1.3 10 7.61 744 P Eutrophic E141.31 Lake S38.24 13 LCT 90 1.1 1241 835 0.67 13.4 1.1 11 24 8.33 4700 N,P Hypereutrophic Cartcarrong E142.45

113 | P a g e

Stable isotope analyses of the chironomid head capsules

Nutrient No Lake Name Lake Coordinates ALT Depth PET Precip AI MAT TP TN Chl a pH COND Trophic status Limitation m m mm/Year mm/Year ˚C mg/L mg/L µg/L - µs/cm Lake S38.35 14 LEM 147 1.2 1241 891 0.72 13.3 0.11 4.4 22 7.98 6420 P Eutrophic Elingamite E143.00 S41.65 15 Lake Lila WP 957 13 720 2207 3.07 6.7 0.019 0.29 1 4.91 23 N,P Mesotrophic E145.96 Lake Lea S41.53 16 LEA 837 0.75 1484 2095 1.41 7.4 0.028 0.44 5 4.83 33 N,P Mesotrophic Pond E145.91

114 | P a g e

Stable isotope analyses of the chironomid head capsules Table 6.2 Stable isotope analyses (δ18O and δ2H) results of southeastern Australian modelled precipitation, lake water, Chironomus spp. and the fractionation factor between chironomus head capsules and lake water (α). δ18O of δ18O of δ18O of α δD of δD of δD of α No Lake Name Lake modelled lake Chironomus [δ18O(Chironomus modelled lake Chironomus [δ2H(Chironomus precipitation water spp. spp - water)] precipitation water spp. sp - water)] VSMO VSMO VSMOW VSMOW VSMOW VSMOW W W Little Llangothlin 1 LLL -6.70 -1.25 18.70 19.78 -39.13 -34.04 -91.70 -61.55 Lagoon 2 Blue Lake BL -7.20 -7.76 12.27 19.99 -44.27 -50.69 -84.82 -36.61 3 Lake Albina LA -7.20 -7.02 13.27 20.23 -44.27 -54.33 -101.17 -50.80 4 Lake Cootaptamba CTL -7.20 -6.53 13.10 19.56 -44.27 -44.49 -94.87 -54.17 5 Nuggety Gully Resrvoir NGR -4.80 11.27 15.77 4.44 -26.61 57.81 -63.29 -121.58 6 Lake Fyans LFY -4.91 3.57 15.28 11.61 -27.47 28.15 -48.18 -77.13 7 Freshwater Lake FWL -5.01 5.57 15.94 10.26 -28.30 40.80 -54.70 -96.25 8 Lake Tooliorook LTK -4.91 1.05 16.88 15.69 -27.61 27.87 -73.17 -103.47 9 Lake Surprise LSP -4.91 1.18 16.80 15.47 -27.70 27.48 -75.40 -105.50 10 Lake Mombeong LMB -4.65 3.34 16.28 12.82 -26.71 28.63 -72.82 -103.83 11 Lake Terangpom LTP -4.81 10.54 13.91 3.33 -26.94 56.67 -55.32 -112.04 12 Swan Lake SWL -4.72 4.02 13.63 9.52 -26.46 43.26 -69.18 -114.04 13 Lake Cartcarrong LCT -4.74 12.97 15.09 2.09 -26.52 61.63 -70.00 -132.38 14 Lake Elingamite LEM -4.85 5.52 16.09 10.45 -27.27 51.33 -50.93 102.33 15 Lake Lila WP -6.21 -6.17 12.43 18.54 -38.17 -34.55 -79.53 -47.71 16 Lake Lea Pond LEA -5.73 2.43 15.22 12.67 -34.60 17.26 -75.81 -95.95

115 | P a g e

Stable isotope analyses of the chironomid head capsules

Regression analyses of the δ18O fractionation factor between the chironomid head capsules and lake water (α[δ18O(Chironomus spp. - water)] ) with respect to climatic and environmental variables showed that seven variables: MAT, Total Nitrogen (TN), Total Phosphorus (TP), Chlorophyll a (Chl a), depth, specific conductance (COND) and AI were significantly correlated (p < 0.05). Redundancy analyses (RDAs) were performed using the α[δ18O(Chironomus spp. - water)] values against all seven variables (Table 6.4). TN was the variable that explained the largest percent (64.4%) of the total variance in the α[δ18O(Chironomus spp - water)] data, although it did not retain its significance level when COND was partialled out (Table 6.4), the RDA on COND showed that when TN was partialled out, it only explained 0.1% of the total variance (Table 6.4). Thus, there is a strong interaction between COND and TN. But clearly TN was the more independent variable. The regression plot of the 18 values of ɑChironomus spp. – water for δ O against TN showed a close negative correlation (with regression coefficient r2 = 0.64) (Fig. 6.4b).

Figure 6.6 (a) Plot of Mean annual temperature (MAT) against δ2H of Chironomus spp.. MAT had the best correlation and is the independent variable that explains the largest percent variance (Table 6.5) of Chironomus spp. δ2H data. The correlation could possibly be complicated by the 2 2 fractionation between δ H of Chironomus spp. and δ H of lake water (α[δ2H(chiro-water)]), which was influenced by other environmental variables. (b) Plot of ln(COND) against fractionation factor

α[δ2H(chiro-water). Specific conductance (COND) explained an independent and large percent of variance in the fractionation factor between δ2H of Chironomus spp. and host lake water (Table 6.6), lake water COND had a close correlation (r2 = 0.695) with the fractionation factor (α) between δ2H of Chironomus spp. and lake water.

116 | P a g e

Stable isotope analyses of the chironomid head capsules Table 6.3 Redundancy Analysis (RDA) of δ18O of Chironomus spp. head capsules and environmental variables shows that Aridity Index (AI) has the strongest explanatory power (53.7%) among the five variables that have a significant correlation (P < 0.05). It is not independent of MAT but it is impossible that it would be independent as temperature plays a key role in potential evapotranspiration (PET). AI explains significantly more variance than MAT and MAT is reduced to very low explanatory power when AI is removed.

Variable Co-variable % of Variance λ1 Correlation p-value explained MAT none 42.5 0.425 0.652 0.005 Cond 24.5 0.179 0.495 0.06 pH 23.9 0.179 0.489 0.07 PET 20.5 0.134 0.453 0.091 AI 0.9 0.004 0.093 0.737 ALL 1.4 0.006 0.119 0.707

AI none 53.7 0.537 0.733 0.002 MAT 20.1 0.115 0.448 0.118 Cond 39 0.285 0.625 0.015 pH 38.3 0.287 0.619 0.014 PET 31.4 0.204 0.56 0.029 ALL 9.5 0.044 0.309 0.326

PET none 34.9 0.349 0.591 0.016 AI 3.6 0.016 0.188 0.494 MAT 10 0.057 0.315 0.246 Cond 15.2 0.111 0.389 0.159 pH 20.7 0.155 0.455 0.093 ALL 7.2 0.032 0.268 0.403

pH none 25.1 0.251 0.501 0.055 AI 0.3 0.001 0.056 0.83 MAT 0.9 0.005 0.093 0.743 Cond 5.2 0.038 0.228 0.411 PET 8.8 0.058 0.297 0.28 ALL 0.1 0 0.028 0.93

COND none 26.9 0.269 0.519 0.035 AI 3.8 0.017 0.194 0.497 MAT 4 0.023 0.199 0.468 pH 7.4 0.056 0.272 0.329 PET 4.8 0.031 0.218 0.424 ALL 6.6 0.029 0.257 0.421

117 | P a g e

Stable isotope analyses of the chironomid head capsules Table 6.4 Redundancy Analysis (RDA) of the fractionation factor (α) (δ18O between Chironomus spp. and lake water) and environmental variables shows that Total Nitrogen (TN) is the most independent variable that explains the largest variance (64.4%) among all seven variables that are significantly correlated (P < 0.05) for the stable oxygen isotope (δ18O) fractionation factor

(α[δ18O(Chironomus spp - water)]) (Table 6.2) between Chironomus spp. and host lake water

Variable Co-variable % of λ1 Correlation p-value Variance explained MAT none 35.7 0.357 0.597 0.006 AI 1.1 0.007 0.107 0.666 TP 10 0.054 0.316 0.244 TN 1.9 0.007 0.136 0.581 Chla 8.8 0.054 0.296 0.29 DEPTH 7.7 0.031 0.278 0.286 COND 0.3 0.002 0.058 0.837 ALL 20.8 0.045 0.456 0.142

TN none 64.4 0.644 0.802 0.001 AI 40.2 0.232 0.634 0.008 MAT 45.7 0.294 0.676 0.006 Depth 29 0.115 0.538 0.03 TP 42.9 0.233 0.655 0.004 Chla 42.4 0.263 0.651 0.004 COND 20.3 0.09 0.45 0.079 ALL 41.9 0.123 0.647 0.036

TP none 45.7 0.457 0.676 0.003 AI 22.9 0.132 0.478 0.061 MAT 23.9 0.154 0.489 0.068 Depth 9.3 0.037 0.304 0.25 TN 12.8 0.046 0.358 0.181 Chla 14.4 0.089 0.379 0.158 COND 6.1 0.027 0.248 0.348 ALL 37.7 0.104 0.614 0.039

Chla none 38.1 0.381 0.617 0.007 AI 11.4 0.066 0.337 0.195 MAT 12.3 0.079 0.35 0.186 Depth 4.8 0.019 0.22 0.433 TP 2.5 0.014 0.159 0.537 TN 0 0 0 1 COND 7.9 0.035 0.281 0.302 ALL 2.3 0.004 0.153 0.681

COND none 55.4 0.554 0.744 0.003 AI 23.9 0.138 0.489 0.06 MAT 30.8 0.198 0.555 0.02 Depth 25 0.099 0.5 0.051 118 | P a g e

Stable isotope analyses of the chironomid head capsules Variable Co-variable % of λ1 Correlation p-value Variance explained TP 22.9 0.124 0.479 0.068 TN 0.1 0 0.356 0.916 Chla 33.6 0.208 0.579 0.022 ALL 24.3 0.055 0.493 0.128

DEPTH none 60.4 0.604 0.777 0.001 AI 37.5 0.216 0.612 0.01 MAT 43.1 0.278 0.657 0.005 TP 33.8 0.184 0.582 0.022 TN 21 0.075 0.458 0.084 Chla 39.1 0.242 0.625 0.013 COND 33.4 0.149 0.578 0.024 ALL 15.9 0.032 0.399 0.224

AI none 42.3 0.423 0.65 0.004 MAT 11.3 0.073 0.337 0.208 TP 18.1 0.099 0.426 0.103 TN 3.1 0.011 0.176 0.518 Chla 17.4 0.108 0.417 0.12 COND 1.7 0.008 0.13 0.635 Depth 9 0.036 0.3 0.245 ALL 6.8 0.012 0.26 0.466

6.5.2 Stable hydrogen isotope (δ2H) results

For deuterium (δ2H), the annual average δ2H of precipitation varied between -44.27 and -26.46 ‰ VSMOW and was closely correlated with the lake water δ2H which varied between -6.17 and 12.97‰ VSMOW (r2 = 0.92) (Fig. 6.5a). The measured δ2H values from the chitinous head capsules of Chironomus spp. from the surface sediments of the sixteen southeastern Australian lakes showed a significant positive correlation (p < 0.05) with host water isotopic composition, with a regression coefficient (r2) of 0.69 (Fig. 6.5b).

2 Regression analyses of δ H of Chironomus spp. values against climatic and environmental variables showed that seven variables (MAT, TN, TP, Chl a, Depth, COND and AI) were significantly correlated (p < 0.05). Redundancy analyses (RDAs) were then performed using the δ2H of Chironomus spp. values against these seven variables (Table 6.5). Results showed that MAT was the variable that explained the largest percent (55.2%) of the total variance in the δ2H Chironomus spp. data. The explanatory power of MAT was reduced by partialling out COND, AI and TN (Table 6.5). The reduction in explanatory power was greatest after partialling out COND and AI (Table 6.5). However, MAT appears to be the most independent variable since the explanatory power of AI 119 | P a g e

Stable isotope analyses of the chironomid head capsules and COND is almost eliminated by partialling out MAT. MAT is clearly more independent than TN since the explanatory power of TN is much more severely affected by the partialling out of MAT than MAT by the partialling out of TN (Table 6.5). COND, TN, AI and MAT are correlated and this is because of the effect of temperature on evaporation and the concentration of ions and nutrients in the lake waters. The regression plot of δ2H of Chironomus spp. against MAT showed a positive correlation with a regression coefficient of r2 = 0.55 (Fig. 6.6a).

2 Regression analyses of the fractionation (ɑChironomus spp. – water) values for δ H against climatic and environmental variables showed that TN, TP, Chl a, COND, Depth, MAT and AI were significantly correlated (p < 0.05). The partial RDAs showed that COND was the sole variable that retained significance (p < 0.05) when all other variables were partialled out and explained 68.4% of the total 2 2 variance in the ɑChironomus spp. – water values for δ H (Table 6.6). The regression plot of δ H fractionation values against COND has a correlation coefficient of r2 = 0.695 (Fig. 6.6b).

120 | P a g e

Stable isotope analyses of the chironomid head capsules Table 6.5 Redundancy Analysis (RDA) of δ2H of Chironomus spp. head capsules and environmental variables shows that for δ2H there are no independent variables. MAT explains the largest variance (55.2%) in stable hydrogen isotope (δ2H) data of Chironomus spp. but interacts with aridity index (AI), specific conductance (COND) and total nitrogen (TN).

Variable Co-variable % of λ1 Correlation p-value Variance explained MAT none 55.2 0.552 0.743 0.001 AI 18 0.098 0.424 0.136 TP 40.2 0.297 0.634 0.016 TN 22.4 0.122 0.473 0.085 Chla 39 0.284 0.624 0.015 DEPTH 35.2 0.209 0.593 0.028 COND 11.6 0.057 0.34 0.24 ALL 5.7 0.014 0.238 0.516

TN none 45.6 0.456 0.675 0.004 AI 13.5 0.073 0.367 0.19 MAT 5.8 0.026 0.242 0.343 Depth 15 0.089 0.387 0.18 TP 55.4 0.41 0.745 0.005 Chla 26 0.189 0.509 0.051 COND 0.7 0.003 0.083 0.777 ALL 38.4 0.14 0.619 0.054

TP none 26.1 0.261 0.511 0.045 AI 4 0.022 0.2 0.5 MAT 1.5 0.007 0.124 0.643 Depth 1.9 0.011 0.138 0.636 TN 39.5 0.215 0.628 0.008 Chla 3.2 0.023 0.178 0.511 COND 1.2 0.006 0.108 0.678 ALL 37.5 0.136 0.613 0.058

Chla none 27.2 0.272 0.521 0.052 AI 1.9 0.011 0.139 0.614 MAT 0.9 0.004 0.092 0.759 Depth 3.3 0.02 0.182 0.526 TP 4.5 0.033 0.212 0.438 TN 0.8 0.004 0.091 0.739 COND 0.4 0.002 0.06 0.817 ALL 9.8 0.024 0.313 0.37

COND none 50.4 0.504 0.71 0.002 AI 14.2 0.077 0.376 0.145 MAT 2.2 0.01 0.147 0.61 Depth 24.8 0.148 0.498 0.06

121 | P a g e

Stable isotope analyses of the chironomid head capsules Variable Co-variable % of λ1 Correlation p-value Variance explained TN 9.4 0.051 0.307 0.303 Chla 32.2 0.234 0.567 0.03 ALL 28.2 0.089 0.531 0.112

Depth none 40.5 0.405 0.636 0.007 AI 12.6 0.069 0.355 0.196 MAT 14 0.063 0.374 0.192 TP 21 0.155 0.458 0.074 TN 7 0.038 0.265 0.337 Chla 21.1 0.153 0.459 0.089 COND 9.8 0.049 0.313 0.258 ALL 0 0 0.018 0.959

AI none 45.5 0.455 0.674 0.004 MAT 0.2 0.001 0.045 0.879 TP 29.1 0.215 0.54 0.04 TN 13.2 0.072 0.364 0.194 Chla 26.6 0.194 0.516 0.046 COND 5.6 0.028 0.237 0.368 Depth 19.9 0.118 0.446 0.101 ALL 0.1 0 0.033 0.934

122 | P a g e

Stable isotope analyses of the chironomid head capsules Table 6.6 Redundancy Analysis (RDA) of the fractionation factor (α) (δ2H between Chironomus spp. and lake water) and environmental variables shows that specific conductance (COND) is the most independent variable that explains the largest amount of variance (68.4%) among all seven variables that are significantly correlated (P < 0.05) in the hydrogen isotope (δ2H) fractionation factor (α) between Chironomus spp. and host lake water

Variable Co-variable % of λ1 Correlation p-value Variance explained MAT none 58 0.58 0.762 0.001 AI 6.4 0.023 0.253 0.357 TP 38.2 0.236 0.618 0.01 TN 27.7 0.119 0.526 0.036 Chl a 36.4 0.235 0.603 0.013 DEPTH 36.3 0.168 0.603 0.011 COND 12.3 0.039 0.35 0.181 ALL 3.1 0.006 0.175 0.608

TN none 57.1 0.571 0.755 0.001 AI 18.6 0.067 0.431 0.089 MAT 26 0.109 0.51 0.036 Depth 21.7 0.1 0.466 0.062 TP 41.3 0.256 0.643 0.01 Chla 33.7 0.218 0.581 0.021 COND 0.02 0 0 0.945 ALL 0.3 0.001 0.052 0.888

TP none 38 0.38 0.617 0.007 AI 9.2 0.033 0.304 0.262 MAT 8.7 0.036 0.295 0.299 Depth 5 0.023 0.223 0.422 TN 15.3 0.066 0.391 0.119 Chla 8.2 0.053 0.286 0.28 COND 0.2 0.001 0.043 0.857 ALL 1.7 0.003 0.129 0.698

Chla none 35.3 0.353 0.594 0.01 AI 2.6 0.01 0.163 0.544 MAT 1.9 0.008 0.136 0.633 Depth 4.6 0.021 0.215 0.399 TP 4.1 0.025 0.202 0.467 TN 0.1 0 0.031 0.927 COND 2.5 0.008 0.159 0.579 ALL 0.1 0 0.037 0.93

123 | P a g e

Stable isotope analyses of the chironomid head capsules Variable Co-variable % of λ1 Correlation p-value Variance explained COND none 68.4 0.684 0.827 0.001 AI 26.4 0.095 0.514 0.024 MAT 33.9 0.142 0.582 0.015 Depth 46 0.213 0.678 0.003 TP 49.1 0.304 0.7 0.003 TN 26.4 0.113 0.514 0.049 Chla 52.4 0.339 0.724 0.002 ALL 2.4 0.005 0.154 0.638

Depth none 53.8 0.538 0.733 0.002 AI 21.5 0.078 0.464 0.089 MAT 29.8 0.125 0.546 0.028 TP 29.1 0.18 0.54 0.034 TN 15.7 0.068 0.397 0.128 Chla 31.8 0.206 0.564 0.022 COND 21 0.066 0.459 0.071 ALL 17.5 0.044 0.419 0.209

AI none 64 0.64 0.8 0.001 MAT 19.7 0.083 0.444 0.089 TP 47.3 0.293 0.688 0.001 TN 31.7 0.136 0.563 0.029 Chla 45.8 0.297 0.677 0.004 COND 16.2 0.051 0.403 0.142 Depth 38.9 0.18 0.624 0.006 ALL 0.9 0.002 0.094 0.806

6.6 Discussion

6.6.1 Analytical considerations

The application of stable isotope analyses on subfossil chironomids is still at an early stage of development and a standard chitin sample purification protocol for chironomid head capsules is not yet established. There have been only two studies that reported the measurement of δ2H on chironomid head capsules (Deines et al., 2009 and Wang et al., 2009) but studies on keratin material (e.g. Wassenaar and Hobson, 2003) suggested that the use of hydrogen isotope ratios needs to be considered carefully because of the rapid, partial exchange of loosely bound hydroxyl and/or amine hydrogen atoms with the environment. Here, hydrogen isotope ratios were not corrected for the contribution of exchangeable hydrogen. This was because at the time of analysis, a set of exchange- calibrated chitin standards was not yet established at PSI (Nielson and Bowen, 2010). The amount of exchangeable hydrogen is usually expressed as percentage of the total amount of hydrogen and is 15.3 ± 2.9% in chitin (Schimmelmann et al., 1993). Hydrogen atoms in exchangeable hydroxyl and 124 | P a g e

Stable isotope analyses of the chironomid head capsules amine groups should have the same isotopic composition across all chitin samples analysed at the same time under the same conditions. This concern was discussed in details in Nielson and Bowen (2010) for the analysis of shrimp chitin from PSI. As all samples reported in this study were processed, pre-treated and analysed using internally consistent methods, the offsets should be constant. Consequently while the absolute δ2H values of Chironomus spp. head capsules obtained from this study will not be suitable to compare directly with other studies, the trends should still yield useful information and are worthy of discussion.

6.6.2 Relationship between δ18O of lake water and regional climate

Due to evaporative enrichment in arid areas and wide range spatial distribution of the sites, there is a disparity in between the δ18O of source water (i.e. precipitation) and the δ18O of lake water. Therefore, it is not surprising that although the δ18O of precipitation was correlated with lake water (Fig. 6.3a, r2 = 0.71) and mean annual temperature (MAT) (Fig. 6.7, r2 = 0.84), the sites are unevenly distributed around the respective regression lines.

We observe that δ18O of lake water is closely correlated with δ18O of precipitation for lakes in perenially wet settings. However, there is a wide scatter of the δ18O of lake water values in lakes with similar precipitation δ18O values, specifically when δ18O precipitation is > 5 ‰ VSMOW (Fig. 6.3a). Sites with precipitation δ18O values > 5 ‰ are located in sub-humid and arid areas (Fig. 6.3a, Table 6.1). Evaporative enrichment of δ18O in lake water is obviously important at these sites and variations in enrichment probably reflect differences in local evaporation rates and residence time as the lighter isotope (16O) is preferentially lost due to evaporation (Leng and Marshall, 2004). Lakes with enriched δ18O with respect to source water are closed maar lakes and all are located in western Victoria. In fact, in two of the western Victorian lakes (LTP and LCT) and one reservoir (NGR), the δ18O values exceed 10‰ VSMOW which is considered an excessively large value (Lamb et al., 1999) and difficult to achieve by natural evaporation processes (Gibson et al., 2002). In this case, very long residence times are likely the primary cause (Barton et al., 2007; Chivas et al., 1993). In summary, Australia has a high inter-annual variability in rainfall and in seasonally arid regions, such as western Victoria, where agricultural activities (e.g. water abstraction) are also predominant, lakes were especially sensitive to water-balance (Herczeg et al., 1992) and changes in local hydrology induced isotopic variations. This finding indicated that lakes that are highly enriched in δ18O may not be suitable to be included in a calibration set for paleoclimate reconstructions.

125 | P a g e

Stable isotope analyses of the chironomid head capsules

Figure 6.7 Plot of mean annual temperature (MAT) against δ18O of precipitation for the 16 study lakes. There is a positive correlation between MAT and δ18O of precipitation, but since the differences within the spatial distribution for both study sites and precipitation sources are large, the data points are scattered.

6.6.3 Lake trophic status related δ18O fractionation process and the metabolism and respiration of Chironomus spp.

A few previous studies (Wang et al., 2009; Verbruggen et al., 2011; Mayr et al., in press) suggested that oxygen stable isotopic composition of chironomid head capsules can serve as a useful tool for lake water isotope reconstructions for δ18O. However, these results were not replicated in this study (Fig. 6.3b). The implication is that insect metabolism and/or respiration in Chironomus spp. may play a vital part in the fractionation between δ18O of lake water and Chironomus spp.. Figure 6.8 shows the experimental chironomid δ18O values against δ18O of host water in Wang et al (2009) compared to this study. Sites where there is a strong relationship between δ18O of lake water and δ18O Chironomus spp. are mostly oligotrophic to mesotrophic lakes in humid areas. In most of the

126 | P a g e

Stable isotope analyses of the chironomid head capsules eutrophic to hypereutrophic lakes in the semi-arid to sub-humid areas, the relationship does not apply.

Figure 6.8 Plot of the experimental chironomid δ18O values against δ18O of host water in Wang et al (2009) compared to values derived from this study. Eight of the lakes from this study (CTL, BL, LA, WP, LEA, LLL, LSP, LTK), fall along the regression line of Wang et al (2009) these are oligotrophic and mesotrophic lakes from Kosciuszko and Tasmania, and two western Victorian lakes that are less enriched in δ18O, i.e. less affected by prolonged residence time. The other eight lakes (FWL, LEM, LCT, LTP, SWL, LFY, LMB, NGR) produced similar Chironomus spp. δ18O values.

There was a reasonably strong, positive (r2 = 0.64) correlation between total nitrogen (TN) and the 18 fractionation factor of Chironomus spp. and lake water (α[δ18O(Chironomus spp - water)]) for δ O (Fig. 6.4b), showing that lake trophic status is an important consideration for the interpretation of δ18O

127 | P a g e

Stable isotope analyses of the chironomid head capsules data for Chironomus spp.. Since seven out of sixteen of our lakes are phosphorous (P) limited (nine are N and P co-limited), TN was always an excess nutrient available in these lakes (Table 6.1). High levels of nutrients (such as TN) fuel algae blooms, which can initially boost dissolved oxygen (DO) level. However, more algae means more plant respiration, drawing down DO, and when the algae die, bacterial decomposition spikes, using up much of the dissolved oxygen available (Granéli and Solander, 1988). The process of the exchange of nutrients across the sediment-water interface in shallow water also enhances low oxygen available environments (Boström et al., 1998). Shallow lakes located in semi-arid regions, such as western Victoria, with high residence times are highly susceptible to this effect (Chapter 3). These lakes are ecologically challenging but Chironomus spp. are dominant in most lakes of this type. It reached a maximum abundance of 63.7% (e.g. Lake Toolirook (LTK), Victoria (Table 6.1)) in lowland eutrophic to hypereutrophic lakes and is present in 94% of the lakes sampled in Chapter 4 in southeastern Australia. This is similar to New Zealand (e.g. Woodward and Shulmeister, 2006), where in some warm lowland eutrophic lakes, Chironomus spp. reached a maximum abundance of 80%.

One of the key features that allows Chironomus spp. to thrive in eutrophic and even anoxic conditions is that they have the ability to metabolise and respire when oxygen availability is low (Walker, 1987; Brodersen et al., 2004). This is because Chironomus larvae have haemoglobin which regulates the respiration and oxygen level in the Chironomus spp. body (Osmulski et al., 1986) and enables them to survive in low oxygen waters.

In summary, stable oxygen isotope values (δ18O) of Chironomus spp. from eight out of sixteen sites within the dataset appear to reflect lake water δ18O and those are mostly lakes from humid and/or cooler regions. A strong local evaporation signal and prolonged residence time caused the lake water δ18O to be enriched by > ~8 ‰ VSMOW with respect to precipitation δ18O in eight of the ten western Victorian sites. High lake nutrient status is associated with the dominance of Chironomus spp. in the species composition of these lakes. We infer that this is an indication of ecological adaptation of Chironomus spp. to extreme lake conditions. It is likely that although Chironomus spp. can survive in these extreme environments due to the oxy-regulatory effect of haemoglobin. Fractionation effects are altered in these extreme (e.g. oxygen limited) environments. Therefore it is unsurprising that δ18O of Chironomus spp. head capsules do not reflect the lake water δ18O for these eight sites (due to vital effects) and consequently, they are not directly correlated to climatic factors. We re-affirm the findings of Mayr et al. (in press) that for semi-arid regions where δ18O of precipitation is not a priori correlated with mean annual air temperature, δ18O values in chironomids cannot be used for temperature reconstructions. 128 | P a g e

Stable isotope analyses of the chironomid head capsules

6.6.4 Relationship between δ18O of Chironomus spp. and regional climate

The previous discussion highlighted the large vital effects on Chironomus spp. δ18O in western Victorian lakes, or lakes subject to eutrophication and evaporative enrichment. When all lakes were considered (Fig. 6.9a) there was a reasonable correlation between δ18O and MAT. However, the correlation was greatly improved when only lakes that have a δ18O enrichment less than ~ 8‰ VSMOW compared to precipitation δ18O were considered, δ18O of Chironomus spp. and mean annual temperature (MAT) (r2 = 0.78) (Fig. 6.9b). The slope (0.51) and intercept (10.7 ± 0.8‰) values of the regression obtained in these eight southeastern Australian lakes were very similar to Verbruggen et al.’s (2011) European dataset, which had a slope of 0.57 and an intercept value of 12 ± 0.32‰ (Fig. 6.9b). This is pleasing, but surprising, as the δ 18O analyses of chironomid head capsules from Verbruggen et al. (2011) were based on a mixed taxon assemblage, including Heterotrissocladius subpilosus-type, Cladotanytarsus, and Cricotopus (Verbruggen et al., 2011). At first inspection, interspecific vital effects may not be as significant as predicted, but this observation needs further investigation.

Our results confirm that lakes that are not affected by significant δ18O enrichment caused by high local evaporation and residence time can be used for temperature reconstructions. Consequently temperature reconstructions from Chironomus spp. δ18O in Australia are applicable only to areas such as Tasmania and the higher altitude areas of the southeastern mainland.

6.6.5 The δ2H of precipitation, lake water, Chironomus spp. and temperature

Our analyses of lake water δ2H revealed a close correlation with δ2H of precipitation (Fig. 6.5a) and the enrichment of δ2H in lake water was not as strongly affected as δ18O by local evaporation and residence time (Fig. 6.3a). The δ2H values of Chironomus spp. and lake water showed a moderately strong and significant (p < 0.05) correlation (r2 = 0.69) (Fig. 6.5b), which indicated a large proportion of Chironomus spp. δ2H was derived from or exchanged with lake water. At first glance, this observation does not appear to agree with the cultured experiment by Wang et al (2009), who suggested that only ~30% of the δ2H of chironomids is derived from ambient water and ~70% is from diet. However, this is expected because in nature, the genus Chironomus is mainly feeding on planktonic algae, where algae also contain δ2H stable isotopes that are derived from lake water (Zhang and Sachs, 2007). In addition, the inter-specifc dietary variability has been reduced in our study as the chironomids samples were composed of a single genus of Chironomus morpho-species. 129 | P a g e

Stable isotope analyses of the chironomid head capsules

Figure 6.9 (a) Plot of mean annual temperature (MAT) against δ18O of Chironomus spp. with all sixteen lakes included (b) Plot of mean annual temperature (MAT) against δ18O of Chironomus spp. with only the eight lakes that have the δ18O enrichment of lake water less than ~8‰ VSMOW relative to precipitation δ18O are included. There is a close correlation (r2 = 0.78) between MAT and Chironomus spp. δ18O for this subset of study lakes. The slope and intercept values were very similar to those observed by Verbruggen et al. (2011).

2 The fractionation (α[δ2H(Chironomus spp - water)]) between δ H of Chironomus spp. and lake water however, follows a specific conductance (COND) gradient (Fig. 6.6b) with a r2 = 0.69 indicating that it is an important factor in the Chironomus spp. δ2H and lake water δ2H relationship. However, without further laboratory experiments and more testing, we cannot provide an insight of how specific conductance affects the fractionation process between δ2H in Chironomus spp. and lake water. In freshwater lakes (specific conductance < ~500 μs/cm, Behar 1997), that support diverse aquatic life, mean annual temperature (MAT) and COND co-varied (Fig. 6.10) and there are potentially three reasons for this observation. First, lake order effect may be driving the salinity gradient where high altitude (cooler temperature) low order lakes are often dilute (e.g. Mt Kosciusko lakes, Fig. 6.1) and lowland (warmer temperature) high order lakes have higher salinity. Secondly, when specific conductance is > 500 μs/cm (e.g. Victorian lakes, Fig. 6.1), the changes in specific conductance are driven by catchment hydrology, e.g. residence time, which affects evaporative enrichment of salts. Finally, the overall pattern is perhaps due to regional evaporation which is a function of temperature.

130 | P a g e

Stable isotope analyses of the chironomid head capsules

Figure 6.10 Plot of ln(COND) against mean annual temperature (MAT). Eight out of ten lakes from Western Victoria are considered saline. Lake Fyans (LFY) and Nuggetty Gully Reservoir (NGR) were the only lakes that had a specific conductance value < 500 μs/cm from this region. Lake order effect and residence time are potentially affecting the observed relationship between temperature and specific conductance and the overall pattern is presumably due to that regional evaporation is a function of temperature

In summary, high specific conductance obscures any relationship between δ2H of Chironomus spp. and MAT but for freshwater lakes there is a strong correlation obtained between δ2H of Chironomus spp. and MAT (r2 = 0.72 (Fig. 6.11)). Since specific conductance is closely related to lake eutrophication and salinization (Chapter 3), temperature values can only be derived from oligotrophic and mesotrophic lakes. These lakes are often found in the cold temperate and high altitude areas in southeastern Australia and are the same lakes as are suitable for δ18O temperature estimates.

131 | P a g e

Stable isotope analyses of the chironomid head capsules

Figure 6.11 Eight lakes that had a ln(COND) < ~ 6 were plotted for mean annual temperature against δ2H of Chironomus spp.. A close correlation was obtained with an r2 of 0.72. These include two lakes (Lake Fyans and Nuggetty Gully Reservoir) from western Victoria and six lakes from non-arid environments. These two western Victorian lakes are freshwater bodies presumably due to short residence times.

6.7 Conclusions and future work

Stable isotopes δ18O and δ2H analysed on head capsules of the chironomid genus Chironomus spp. from sixteen southeastern Australian lakes suggested that the following aspects need to be considered when applying them as temperature proxies. For δ18O, regional aridity and its consequential issues are critical. For lakes that are located in semi-arid and sub-humid areas, residence time adds to the evaporative enrichment effect and this results in a high concentration of nutrients that create low oxygen environments. Haemoglobin helps regulate the oxygen level in Chironomus and allows this genus to dominate in these low oxygen water bodies but their chitin δ18O cannot be used to reconstruct temperatures from these lakes. In contrast, for lakes in humid settings and where nutrient concentrations are low, the relationship between δ18O of Chironomus and MAT appears robust.

132 | P a g e

Stable isotope analyses of the chironomid head capsules

Although, the δ2H values obtained on Chironomus spp. head capsules in this study were not calibrated for the contribution of exchangeable hydrogen and the absolute values are not reliable, the trends in the data are robust and yield a similar story to δ18O. For fresh water lakes, the δ2H values of Chironomus spp. head capsules showed that it is a promising proxy for MAT reconstructions. Stable hydrogen isotope (δ2H) values of Chironomus spp. have the apparent advantage that they are less influenced by insect biological controls (i.e. haemoglobin effects). It also appears that the δ2H of Chironomus spp. values are less strongly affected by evaporative enrichment. Since salinity and eutrophication are closely related in these Australian lakes (Chapter 3), the overall findings of this paper are that both δ18O and δ2H are potentially valuable tools for reconstructing temperature in low nutrient lakes in humid parts of Australia (e.g. Tasmania and the southeastern highlands).

133 | P a g e

Conclusions and future work

7 CONCLUSIONS AND FUTURE WORK

7.1 Summary of research and major findings

This thesis aimed to develop and apply two methods that use subfossil remains of non-biting midge larvae of chironomids (Diptera: Chironomidae) preserved in lake sediments to reconstruct past changes in the Australian climate system during and since the Last Glacial Maximum (LGM).

For the first method, a model (transfer function) for reconstructing past summer temperatures based on the temperature tolerance of chironomid species living in modern eastern Australian lakes was created. This involved the physio-chemical investigation of 45 water bodies from tropical Queensland to Tasmania. The transfer function was developed based on 33 water bodies from southeastern Australia (the eleven tropical lakes are not included). It was applied to data derived from chironomid remains extracted from lake sediment deposits from Welsby Lagoon, a subtropical site in North Stradbroke Island, Australia spanning the last LGM and the last deglaciation. A quantitative reconstruction was produced and this has provided the first quasi-continuous for the period between 23.2 – 15.5 cal ka BP, detailed seasonal temperature record from mainland terrestrial Australia covering the critical periods in the late Quaternary.

After the first method was established and applied, a second method which was based on the stable oxygen and hydrogen isotope composition (δ18O, δ2H) of the chitinous heads from chironomids from the same lakes was first explored for the Australasian region. The potential of applying stable isotopes δ18O and δ2H analysed on a single genus of chironomid head capsules (Chironomus spp.) from southeastern Australian lakes as proxies to reconstruct paleo-temperature was investigated. The applicability and the complexities of using this method as independent temperature proxies in Australia were discussed.

7.1.1 Connections of modern limnology to climate and land-use

This thesis has first provided an overview of the current status of 45 water bodies and further investigated the interactions and connections of limnology, climate and land-use of these natural and artificial water bodies extending across the eastern coast of Australia from the tropics of Queensland to Tasmania. This research has been presented in Chapter 3 (Objective 1) in the form a published paper (Chang et al., 2014). The paper comprises a literature review of the past limnological studies on Australian lakes. A detailed description of the climate, vegetation and 134 | P a g e

Conclusions and future work geology of the study area has also been provided. Measurements of a broad variety of physio- chemical, land-use and climatic parameters were obtained using field, laboratory and spatial techniques. Multivariate statistical analyses, notably Principle Component Analyses (PCAs) and Redundancy Analyses (RDAs) were applied to explore the relationships and inter-correlations of the measured variables.

The major findings from this chapter were 1) that most mainland Australian lakes are naturally moderately to significantly eutrophic. Oligotrophic lakes are limited to high alpine and perennially wet settings. 2) Reservoirs and other artificial lakes behave similarly to natural lakes, except that they tend to be less eutrophic. This is most likely due to catchment management and a shorter residence time. 3) The primary source of the salts in east Australian lakes is seawater, either derived directly from aerosols or indirectly via marine bedrock. 4) Significant number of lakes in this dataset are limited by nitrogen (N) and phosphorus (P), suggesting that both of these nutrients should be considered in addressing eutrophication in Australian lakes. 5) Mainland lakes are susceptible to becoming highly alkaline as a result of long-term increases in salinity that are related to the concentration of cations during drought periods. Finally, a testable hypothesis was proposed stating that climatically related eutrophication is more significant than nutrient loading by agriculture in the eastern Australian lakes that were examined in this study. This is particularly likely for shallow endorheic basins in semi-arid regions.

The results and findings obtained from this chapter were critical for the development of a chironomid (Diptera: Chironomidae) based transfer function (Chapter 4, Objective 2) and the interpretation of a paleo-record (Chapter 5, Objective 3). This is because understanding the characteristics and the nexus of the geology, climate, water chemistry and the major environmental controls on modern lakes is essential to identify the primary driving variables of the changes in chironomid species assemblages and the species distribution in a region for both temporal and spatial applications. This chapter has also provided valuable information to assist the understanding of the stable isotope analyses results obtained on chironomid head capsules (Chapter 6, Objective 4).

7.1.2 The transfer function

A chironomid based temperature transfer function developed from a training set of 33 natural and artificial lakes from southeast Australia spanning the subtropics to the alpine zone has been presented in Chapter 4 (Objective 2 – Chang et al., 2015a). Eleven lakes from the tropics were not

135 | P a g e

Conclusions and future work included because this thesis is focussed on climate reconstructions from southeastern Australia. Previously published transfer functions were reviewed from other parts of the globe and a summary table has been provided in Appendix 1. Chironomid taxonomic analyses were performed for each of the sampled lakes and a list of taxa identified with corresponding pictures was discussed and shown in Chapter 4 and Appendix 2 respectively. An ecological note is accompanied with the pictures. Multivariate statistical analyses, canonical correspondence analysis (CCA) and partial CCA were used to study the distribution of chironomids in relation to the environmental and climatic variables. The results showed seven out of eighteen available variables were significant (p < 0.05) with respect to chironomid species variation. These were mean February temperature (9.5%), pH (9.5%), specific conductance (8.2%), total phosphorus (8%), potential evapotranspiration (8%), chlorophyll a (6.9%) and water depth (6.2%). Though the total explained variance was low, Mean February temperature (MFT) was found to be the most robust and independent variable explaining chironomid species variation after further pCCA analyses.

A MFT transfer function was then constructed based on the modern distribution of chironomids species in southeast Australia. The best MFT transfer function was a partial least squares (PLS) 2 model with a coefficient of determination (r jackknifed) of 0.69, a root mean squared error of prediction (RMSEP) of 2.33 °C, and maximum bias of 2.15 °C. Compared with transfer function models developed from other parts of the world, this transfer function has comparable r2 (Jackknifed) but a relatively higher RMSEP. This reflects the long scalar length of the temperature gradient which extends from sub-tropical to sub-alpine locations, some 14 °C. The RMSEP represents 16% of the scalar length and this is comparable with most other transfer functions (e.g. Potito et al., 2014). It has been observed that data sets with large temperature gradients naturally have larger errors (Walker and Cwynar, 2006) but this in no way diminishes the value of the reconstruction.

The main findings from this chapter were that 1) chironomids assemblages in actively managed reservoirs (non-impacted) show no significant difference to natural lakes in the same climate and vegetation zones, and therefore can be included for transfer function development. 2) Although we cannot completely rule out some degree of endemism in the Tasmanian chironomid fauna, our analyses show that the degree of endemism indicated by earlier studies is greatly reduced. This raises the real possibility of integrating the existing chironomid transfer function for Tasmania (Rees et al., 2008) with this new model for the southeastern Australian mainland. 3) The transfer function is appropriate (e.g. the RMSEP is relatively too to be used for the late Holocene period) for the reconstruction of summer temperatures during the last glacial maximum (LGM) and the late

136 | P a g e

Conclusions and future work glacial to Holocene transition in eastern Australia. The transfer model was then applied to an LGM and deglacial record.

7.1.3 The application of the transfer function: temperature reconstruction covering the Last Glacial Maximum and the last deglaciation from subtropical Australia

A summer temperature reconstruction based on the newly developed chironomid transfer function from southeastern Australia in Chapter 4 (Objective 2) has been presented from Welsby Lagoon of North Stradbroke Island in Chapter 5 (Objective 3 – Chang et al., 2015b). This is the first quantitative and detailed seasonal reconstruction of the LGM and the last deglacial temperatures from terrestrial Australia. A review of the past inferences of the LGM temperatures in southeastern Australia was provided in the chapter. Fossil chironomid taxonomic analyses from the LGM to the present were performed throughout the 450 cm sediment core, with 13 samples concentrated between the periods of 23.2 – 15.5 cal ka BP. Each of the identified fossil taxa was listed and photographed and the pictures have been provided in Chapter 4 and Appendix 3 respectively. Sediment samples that generated sufficient chironomid head capsules for quantitative temperature reconstructions detailed the period between 23.2 and 15.5 cal ka BP. Several reconstruction diagnostics, including goodness-of-fit, modern analogue technique (MAT) and rare taxa analysis were applied to evaluate the reliability of the reconstruction. Trajectory diagrams were also constructed to show if the changes of chironomid assemblages in the fossil samples were driven by variables other than temperature through time.

The major findings from this chapter were 1) the temperature reconstruction from the LGM to the deglacial period of this site is sensible. 2) Factors that may affect the reliability of the inferred temperatures include changes in lake chemistry and lake level. These had an impact on chironomid assemblages, particularly in the Holocene. These Holocene samples are dominated by a temperature insensitive taxon, Dicrotendipes. 3) The relatively small modern calibration training set also limits the ability to fully understand the environmental controls on some of the fossil samples. 4) The overall reconstruction displays similarities to marine records from areas affected by the East Australian current (EAC), both in terms of timing and magnitudes of cooling during and immediately after the LGM. 5) A maximum summer temperature cooling of c. ~6.5°C at the late LGM at c. ~18.5 cal ka BP is shown, followed by rapid warming to (near) Holocene values by 17.3 cal ka BP. 6) the warming is consistent with the start of deglaciation at 18.2 ± 0.7 cal ka BP derived from ice core records in Antarctica. 7) Results suggest that this coastal record resembles the thermal

137 | P a g e

Conclusions and future work history of adjacent marine sites and the Antarctic climate influence reached the Australian subtropics during the LGM and the last deglaciation.

This long term temperature record has filled a critical gap in our current knowledge of temperature changes during the glacial times in the subtropics and the connections to changes of the ocean surface current and the higher latitudes. The transfer function method developed in Chapter 4 (Objective 2) has been confirmed as a valuable tool to quantify terrestrial temperatures from eastern Australia during glacial times. However, since there have been very limited independent proxies that can be used for quantitative temperature reconstructions extending back to the glacial times in Australia, a second method based on the stable isotope measurements of chironomid head capsules was explored in Chapter 6 (Objective 4).

7.1.4 The potential of applying stable oxygen and hydrogen isotopes from Australian subfossil chironomid head capsules as independent proxies for past change

Stable isotopes (δ18O and δ2H) analysed from head capsules of the chironomid genus Chironomus spp. from sixteen southeastern Australian lakes were measured and this work has been included in Chapter 6 (Objective 4 – manuscript submitted). This is the first use worldwide of both δ18O and δ2H instantaneously from the same samples using a single genus of chironomid head capsules. It is also the first exploration of the potential of their application to reconstruct paleo-temperature in the Australasian region. The relationship between the stable isotope composition (δ18O, δ2H) of Chironomus spp. head capsules and how this relates to the composition of lake water and local precipitation was investigated. The study also inspected whether fractionation or vital effects were constant or varied with changes in the environment. Regression and redundancy analyses were applied to examine the relationships of the measured variables.

Results suggested that the following aspects need to be considered when applying them as temperature proxies. 1) For δ18O, regional aridity and its consequential issues are critical. For lakes that are located in semi-arid and sub-humid areas, residence time adds to the evaporative enrichment effect and this results in a high concentration of nutrients that create low oxygen environments. 2) Haemoglobin helps regulate the oxygen level in Chironomus spp. and allows this taxon to dominate in low oxygen water bodies but their chitin δ18O cannot be used to reconstruct temperatures from these lakes. 3) For lakes in humid settings and where nutrient concentrations are low, the relationship between δ18O of Chironomus spp. and mean annual temperature (MAT) appears robust. 4) The δ2H values obtained on Chironomus spp. head capsules in this study were

138 | P a g e

Conclusions and future work not calibrated for the contribution of exchangeable hydrogen and the absolute values are not reliable, but the trends in the data are robust and yield a similar story to δ18O. 5) For fresh water lakes (specific conductance measurements of < 500 μs/cm), the δ2H values of Chironomus spp. head capsules showed that it is a promising proxy for MAT reconstructions. 6) Stable hydrogen isotope (δ2H) values of Chironomus spp. have the apparent advantage that they are less influenced by insect biological controls (i.e. haemoglobin effects) and it also appears that the δ2H of Chironomus spp. values are less strongly affected by evaporative enrichment. 7) The overall findings of this study are that both δ18O and δ2H are potentially valuable tools for reconstructing temperature in low nutrient lakes in humid parts of Australia (e.g. Tasmania and the southeastern highlands) and this is because, as it has been concluded in Chapter 3 of the thesis, salinity and eutrophication are closely related in the Australian lakes dataset.

7.2 Research significance

In this thesis, valuable tools to quantify past climate in Australia were explored and established. It demonstrated that the transfer function can be applied to the late Quaternary data to generate quantitative summer temperature reconstructions in eastern Australia. The chironomid-based temperature record from Welsby Lagoon showed cooling and warming trends similar to those derived from nearby marine records and that deglaciation is synchronous with the records from Antarctica. This suggests a climate link from terrestrial Australian subtropics to the nearby ocean and hence to the high latitudes. It has demonstrated that critical gaps in our knowledge on climate change during the glacial period from Australia can be addressed by applying the chironomid-based transfer function developed in this thesis. Stable oxygen and hydrogen isotopes were measured on chironomid head capsules and showed the potential of using chironomid δ18O and δ2H as proxies for temperature reconstructions. The application will be likely limited to non-arid and non-impacted areas of Australia.

The Welsby Lagoon record has also shown that chironomid-based summer temperature reconstructions from a coastal site can contribute to establishing links among oceanic, terrestrial, atmospheric, polar and insolation driven changes in temperature. This has further advanced our knowledge on the hemispheric and global climate teleconnections.

7.3 Critical reflections

139 | P a g e

Conclusions and future work 7.3.1 The debate and controversy about transfer function and its application

Recently, the debate about whether chironomid-based transfer functions are reliable tools to quantify past temperatures has been re-opened. One of the key arguments is that the transfer function method simplifies the complexity of the real world by maximizing single ecological gradients, by not taking into account co-varying variables (Velle et al., 2010). This is typical of empirical or mathematical models ranging from the field of engineering, through ecology to the social sciences. The basic principle in developing a quantatitive method to model real world complexity is the need to minimize the degrees of freedom and this is the key issue regarding the appropriate use of the technique. Velle et al. (2010) raises the problem related to confounding factor (such as temperature may interact with trophic status, water depth etc.). In reply, to Velle et al. (2010), Brooks et al. (2012) argue in many cases, temperature is an independent variable. In Chapter 3 of this thesis, a partial Canonical Correspondence Analysis (pCCA) was applied to the modern chironomid calibration data and results show that mean Febuary temperature (MFT) is not confounded by other variables.

Velle et al. (2010) also argues that one of the main reasons for the strong performance of the chironomid-based transfer function models is because the temperature gradient is maximised in the calibration data set and the gradients of other environmental variables are minimised. This is true in many studies (e.g. Diffenbacher-Kraal et al., 2007). This is not true for the data set (Chapter 3) developed in this thesis. Due to the limited number of lakes in Australia, manipulation of a temperature gradient is unachievable. Temperature, trophic status, salinity and pH all have a large range of values. This calibration data set suggests, in line with evidence as argued in Brooks et al. (2012), that temperature is a particularly significant variable in driving the composition of chironomid assemblages.

In addition to Velle et al. (2010; 2012), many other studies and reviews (e.g. Hann et al., 1992; Walker and Matthewes, 1989; Warwick, 1989; Brooks and Birks, 2001) in the literature critically analysed the application of the chironomid-based transfer function method to the reconstruction of past temperatures. There are several limitations: first, the environmental preferences of one or more taxa might change on geological timescales. Second, different taxa might bloom during different times of the year (although the prime season for chironomids blooming is summer in temperate and polar climate zones), and this may also change due to evolutionary pressures. Third, taxa assemblages could be subjected to some confounding effects such as that water depth, temperature and oxygen concentration can be correlated and the structure of the correlation changes on

140 | P a g e

Conclusions and future work geological timescales. These can result in reconstructions from adjacent sites showing inconsistent patterns regarding inferred temperature changes, such as the Holocene records from Scandinavia discussed in Velle et al. (2010).

As recognized in both Velle et al. (2010; 2012) and Brooks et al. (2012), the possible solutions are as follows: sites for downcore records should be carefully selected, chironomid-based temperature reconstruction should be undertaken within a multiproxy framework, and fossil assemblages should be carefully evaluated for modern analogues and tested against random data (Telford and Birks, 2011). Reconstruction diagnostics were adequately applied for the application to Welsby Lagoon data in Chapter 5. Multiproxy approaches are preferable but not easily available. Diatoms are absent from my sites so I compared and cross-validated our reconstruction with the pollen record from the same site and temperature records based on other independent proxies from the region. This is less desirable as pollen is not strongly correlated with temperature in subtropical Australia (Moss et al., 2013) and with the exception of aquatics represents regional terrestrial variation rather than changes within the lake basin.

Despite these caveats, the transfer function technique has been a cornerstone of biologically based paleoclimatological reconstruction for about thirty to forty years. Palaeoecologists and paleolimnologists should not dismiss this useful method, but rather should treat the reconstructions with due caution and continue to refine the understanding of the proxies (such as chironomids) (e.g. Brooks et al., 2012; Velle et al., 2012).

7.3.2 The limitation of the stable isotope method

A number of limitations of the stable isotope method were discussed in Chapter 6. These include first, the method is only useful for reconstructing temperatures in low nutrient lakes in humid parts of Australia (e.g. Tasmania and the southeastern highlands). Second, the analyses of stable isotopes currently requires a minimum dry weight of 100 μg (equivalent to 150 head capsules), and this has limited the application to some parts of the record from palaeo-sites as chironomids are simply not common enough to collect this mass of a single taxon. Third, similar to the application of the transfer function method, it is challenging to recognize and validate what parts of a chironomid- based stable isotope record are likely to be reliable.

Stable isotope reconstruction should be undertaken within a multiproxy framework. Chironomid taxa should be identified from the same part of the record and the transfer function method can be

141 | P a g e

Conclusions and future work applied as a cross-validation for the stable isotope record. Alternatively, other independent proxies such as pollen which could provide information about positive/negative moisture balance, diatoms which could provide information about salinity changes and other groups such as zooplankton which could provide information about trophic status will help to validate the temperature reconstructions based on chironomid derived stable isotope records.

7.4 Recommendations and future work

7.4.1 Extension and integration of the modern calibration dataset in Australia

In Chapter 4 of this thesis, previously claimed endemism for Tasmanian chironomid taxa was assessed and results showed that the degree of endemism indicated by earlier studies is much less than expected, although some degree of endemism in the Tasmanian chironomid fauna cannot be completely ruled out. Very recent studies using molecular markers (Cranston and Krosch 2015; Krosch et al., 2015) have demonstrated that the previously suspected endemic Tasmanian chironomid fauna are not endemic. On the mainland, the previously assumed endemic Tasmanian taxa are present at several high elevation sites including lakes on Mount Kosciuszko. Therefore, we can be more confident about integrating the existing chironomid transfer function for Tasmania with this new model for the southeastern Australian mainland.

In the future, the integration of the lakes incorporated in this thesis with the Rees et al. (2008) Tasmania dataset is planned. This will increase the modern chironomid calibration set from eastern Australia to ~100 lakes. The major advantages of including a large number of modern calibration sites are that it can be expected that most chironomid taxa encountered in the late Pleistocene and Holocene lake sediment records from both lowland and mountainous sites are well-represented. Consequently, it will be more widely applicable, both temporally and geographically. This also has the potential to help with improving the statistical performance of the robustness (model coefficient) and accuracy (root mean squared error in prediction, RMSEP) of the transfer function model, which will then further constrain the absolute scale of past temperature variations. Well- constrained quantitative temperature values reconstructed from the glacial times will be highly valuable as validation data for palaeoclimate modelling.

To encourage the wider application of the chironomid-based techniques in Australia, a photographic subfossil chironomid identification guide will be developed for Australia. Pictures of the identified subfossil chironomid taxa used in this thesis have been provided in Appendix 2 and Appendix 3. 142 | P a g e

Conclusions and future work These materials will be merged with the chironomid taxa identified and photographed from Tasmanian lakes in Dr Andrew Rees’s PhD thesis (Rees 2013). The eleven tropical lakes that were sampled and included in Chapter 3 will also be encompassed in the future subfossil chironomid work.

7.4.2 Application of the transfer function to quantify long-term (late-Pleistocene to Holocene) temperature changes from various sites

The 450 cm sediment core taken from Welsby Lagoon analysed in this thesis dated back to c. ~ 24 cal ka BP. During recent field work, much longer sediment cores have been recovered from this site and the record of this site has the potential of extending through the last glacial cycle. If there are sufficient chironomids throughout the older sections of the record, the current transfer function will be applied to reconstruct temperature changes of the earlier glacial times.

Early palaeoecological and palaeoclimate investigations of chironomids from the Southern Hemisphere were restricted to qualitative and semi-quantitative interpretations because the numerical techniques for creating transfer functions were still in an early stage of development. In Australia, two previous chironomid species records from Lake Barrine, Atherton Tableland and Blue Lake, Mount Kosciuszko (Dimitriadis and Cranston, 2001; Schakau 1993 respectively) have demonstrated the applicability of subfossil chironomids as a proxy for unravelling the regional paleoclimate history from tropical to alpine climate respectively. With the transfer function established in this thesis, both of these sites can be revisited and the transfer function can be applied to the chironomid data provided in the published work. Quantitative reconstructions can be developed and this will provide valuable information to understand climate changes and to identify abrupt climate events (if there are any) in the late glacial to the Holocene from these two very different climate zones of Australia. A two-metre sediment core was recovered from Lake Cootaptamba, Mount Kosciuszko in December 2014, and radiocarbon dates will be obtained from this core. Chironomid analysis will be performed under a recently funded project (ANLGRA15008, in collaboration with CI: Shulmeister, J), to reconstruct the climate history from the site. Other potential sites that the transfer function could be applied to reconstruct the climate history extend back to the last glacial maximum (LGM) are Mountain Lagoon, Blue Mountains (New South Wales), Lake Frome, South Australia (in collaboration with Cohen, T) and Hazards Lagoon, eastern Tasmania (in collaboration with Mackenzie, L and Moss, P). Sediment cores of the above sites were obtained from various working groups and have established radiocarbon age chronology.

143 | P a g e

Conclusions and future work 7.4.3 Further exploration and application of the chironomid δ18O and δ2H method

In Chapter 6 (Objective 4) of this thesis, the overall conclusion made was that both δ18O and δ2H are potentially valuable tools for reconstructing temperature in low nutrient lakes in humid parts of Australia (e.g. Tasmania and the southeastern highlands). To further develop and test this method, future studies should be conducted for lakes located in regions that have similar environmental and climatic conditions to southeastern Australia, such as some areas in South America, as potential evapo-transpiration (PET) plays a key role in the application of the stable isotope methods. The method will be better understood when it is tested in both arid (Precipitation < PET) and non-arid (Precipitation > PET) conditions. In this thesis it was concluded that δ18O of Chironomus spp. and mean annual temperature (MAT) are strongly correlated (r2 = 0.78) and the relationship obtained in the nutrient low southeastern lakes were very similar to Verbruggen et al.’s (2011) European dataset, so it is reasonable to apply the method to Mount Kosciuszko or Tasmanian sites. Specific potential sites for isotope work include Lake Cootaptamba and Eagle Tarn (Rees and Cwynar, 2010), as the transfer function application has been (or will be) established from these lakes. In the future, the stable isotopes method can be validated for temperature reconstructions, in adjunct with applying the transfer function method. On the other hand, the collection of enough material from paleo-sites is challenging and the use of stable isotopes is recommended for late Holocene and modern studies, perhaps focussed on nutrient changes rather than for paleo-temperature reconstructions.

7.4.4 Regional patterns and hemispheric to global connections

Many literature reviews have addressed the fact that there are inadequate quantitative and multi- proxy based paleo-temperature data in the Southern Hemisphere, for us to gain a systematic understanding of the past climate change over the last 20,000 years. This has hindered us from advancing our knowledge of hemispheric and global climate teleconnections. Within the interpretation of the limited paleoecological records, there are divergent views concerning climate dynamics during key periods of large-scale climate fluctuations, such as the Younger Dryas (YD) and Antarctic Cold Reversal (ACR). The disagreement stems from the disparities among the terrestrial records based on pollen, beetles, chironomids and glaciers from the Southern Hemisphere. This discrepancy may be due to different proxy responses to temperature changes, chronological issues or sampling resolution. However, these inconsistencies could also reflect the actual regional climate variability. For instance, there could be differential influences of the Southern Hemisphere Westerly Winds on the east or west side of the Andes, New Zealand and Tasmania, at similar latitudes.

144 | P a g e

Conclusions and future work

The future constructed chironomid-based high resolution, detailed quantitative temperature records of the same region and regions from other parts of Southern Hemisphere can be compared and synthesized to reveal if there are regional and hemispheric coherent patterns. An inter-proxies approach, such as using chironomid data in conjunction with the pollen and glacial records, could then be used to elucidate the regional and hemispheric paleoclimate conditions (e.g. seasonality and rainfall). This has been carried out in New Zealand for the INTIMATE project (e.g. Alloway et al., 2007) but is not yet possible at a high level of resolution for Australia (e.g. Reeves et al., 2013) and is definitely worth working towards.

145 | P a g e

References

References

Abell, J. M., Ozkundakci, D., and Hamilton, D. (2010). Nitrogen and phosphorus limitation of phytoplankton growth in New Zealand lakes: implications for eutrophication control. Ecosystems 13, 966–977. Alloway, B.V., D.J. Lowe, D.J.A. Barrell, R.M. Newnham, P.C. Almond, P.C. Augustinus, N.A.N. Bertler, L. Carter, N.J. Litchfield, M.S. McGlone, J. Shulmeister, M.J. Vandergoes, P.W. Williams & NZ-INTIMATE members (2007). Towards a climate event stratigraphy for New Zealand over the past 30,000 years (NZ-INTIMATE project).Journal of Quaternary Science 22, 9-35 Armitage, P.D. 1995. Behaviour and ecology of adults. In: Armitage, P.D., Cranston, P.S. and Pinder, L.C.V., (eds.), The Chironomidae. The biology and ecology of non-biting midges. Chapman and Hall, London, pp. 194-224 Anderson, P.M., Bartlein, P.J., Brubaker, L.B., Gajewski, K., Ritchie, J.C. (1989). Modern analogues of late quaternary pollen spectra from western interior of North America. Journal of Biogeography. 16, 573-596. Australian National University Dept of Archaeology and Natural History (2000). Palaeoworks [electronic Resource]. http://nla.gov.au/nla.arc-61389 (accessed 13.03.15). Baker, P. D., and Humpage, A. R. (1994). Toxicity associated with commonly occurring cyanobacteria in surface waters of the Murray–Darling Basin, Australia. Marine and Freshwater Research 45, 773–786. Barley, E.M., Walker, I.R., Kurek, J., Cwynar, L.C., Mathewes, R.W., Gajewski, K., Finney, B.P. (2006). A northwest North American training set: distribution of freshwater midges in relation to air temperature and lake depth. Journal of Paleolimnology. 36, 295–314. Barr, C.S., Tibby, J., Marshall, J.C., McGregor, G.B., Moss, P.T., Halverson, G.P., Fluin, J. (2013). Combining monitoring, models and palaeolimnology to assess ecosystem response to environmental change at monthly to millennial timescales: the stability of Blue Lake, North Stradbroke Island, Australia. Freshwater Biology 58, 1614-1680. Barrows, T.T., Stone, J.O., Fifield, L.K., Cresswell, R.G. (2002). The timing of the last glacial maximum in Australia. Quaternary Science Reviews 21, 159-173. Barton, A.B., Herczeg , A.L., Dalhaus, P.G., Cox, J.W. (2007). A geochemical approach to determining the hydrological regime of wetlands in a volcanic plain, south-eastern Australia, p. 8, IAH Congress: Groundwater and ecosystems, Lisbon Bayly, I. A. E. (1964). Chemical and biological studies on some acidic lakes of east Australian sandy coastal lowlands. Marine and Freshwater Research 15, 56–72. 146 | P a g e

References Bayly, I. A. E., Ebsworth, E. P., and Wan, H. F. (1975). Studies on the lakes of Fraser Island, Queensland. Marine and Freshwater Research 26, 1–13. Behar, S. (1997). Definition of water quality parameters, Testing the waters: chemical and physical vital signs of a river. River Watch Network, Montpelier. Beklioglu, M., Meerfhoff, M., Søndergaard, M., and Jeppesen, E. (2011). Eutrophication and restoration of shallow lakes from a cold temperate to a warm mediterranean and a (sub)tropical climate. In ‘Eutrophication: Causes, Consequences and Control’. (Eds A. A. Ansari, G. S.Singh, G. R. Lanza and W. Rast.) pp. 91–129. (Springer: Dordrecht, The Netherlands.) Bell, D. M., Hunter, J. T., and Haworth, R. J. (2008). Montane lakes (lagoons) of the New England tablelands bioregion. Cunninghamia 10, 475–492. Bennett, K.D. (2002). Psimpoll 3.10: C Programs for Plotting Pollen Diagrams and Analysing Pollen Data. Uppsala University, Uppsala, p. 117. Berger, A., Loutre, M.F. (1991). Insolation values for the climate of the last 10 million years. Quaternary Science Reviews 10, 297-317. Bergstrom, A. K. (2010). The use of TN : TP and DIN : TP ratios as indicators for phytoplankton nutrient limitation in oligotrophic lakes affected by N deposition. Aquatic Sciences 72, 277– 281. Beutel, M. W., and Horne, A. J. (1999). A eeview of the effects of hypolimnetic oxygenation on lake and reservoir water quality. Lake and Reservoir Management 15, 285–297. Birks, H.J.B. (1995). Quantitative palaeoenvironmental reconstructions. In: Maddy, D., Brew, J.J. (Eds.), Statistical Modeling of Quaternary Science Data. Quaternary Research Association, Cambridge, pp. 161-254. Birks, H.J.B. (1998). Numerical tools in paleolimnology progress, potentialities, and problems. Journal of Paleolimnology. 20, 307–332. Birks, H.J.B. (2008). Palaeoecology. Encyclopedia of Ecology (ed. S.E. Jørgensen), pp. 2623–2634. Elsevier Ltd., Oxford Birks, H.J.B., Line, J.M. (1992). The use of rarefaction analysis for estimating palynological richness from Quaternary pollen-analytical data. The Holocene 2, 1–10. Birks, H.J.B., Line, J.M., Juggins, S., Stevenson, A.C., ter Braak, C.J.F. (1990). Diatoms and pH reconstruction. Philosophical Transactions of the Royal Society B 327, 263-278. Blockley, S.P.E., Lane, C.S., Hardiman, M., Rasmussen, S.O., Seierstad, I.K., Steffensen, J.P., Svensson, A., Lotter, A.F., Turney, C.S.M., Bronk Ramsey, C. & Intimate Members (2012) Synchronisation of palaeoenvironmental records over the last 60,000 years,

147 | P a g e

References and an extended INTIMATE event stratigraphy to 48,000 b2k. Quaternary Science Reviews 36, 2–10. BOM (Australian Bureau of Meteorology), (n.d). Australian climate influences. Retrieved 13.01.15., from http://www.bom.gov.au/watl/about-weather-andclimate/australian-climate- influences.shtml BOM (Australian Bureau of Meteorology), (2015). Summary Statistics: Dunwich Post Office. Climate Statistics for Australian Locations. Retrieved 10.01.15., from http://www.bom.gov.au/climate/averages/tables/cw_040070.shtml. Borror, D.J., Triplehorn, C.A., Johnson, N.F. (1976). An introduction to the study of . 6th edition. Thomson Learning Inc., pp. 852. Bostock, H.C., Opdyke, B.N., Gagan, M.K., Fifield, L.K. (2004). Carbon isotope evidence for changes in Antarctic intermediate water circulation and ocean ventilation in the southwest Pacific during the last deglaciation. Paleoceanography 19, PA4013. Bostock, H.C., Opdyke, B.N., Gagan, M.K., Kiss, A.E., Fifield, L.K. (2006). Glacial/interglacial changes in the East Australian current. Climate Dynamics 26, 645-659. Boström, B., Andersen, J., Fleischer, S., Jansson, M. (1988). Exchange of phosphorus across the sediment-water interface, pp. 229-244. In: G. Persson, M. Jansson (eds.), Phosphorus in freshwater ecosystems. Springer Netherlands Boubee, J.A.P., 1983. Past and present Benthic Fauna of Lake Maratoto, with special reference to the Chironomidae. Ph.D. Thesis, University of Waikato, New Zealand, 151pp. Bowen, M.M., Wilkin, J.L., Emery, W.J. (2005). Variability and forcing of the East Australian Current. Journal of Geophysical Research: Oceans 110, C03019. Bowling, L. C. (1988). Optical properties, nutrients and phytoplankton of freshwater coastal dune lakes in south-east Queensland. Marine and Freshwater Research 39, 805–815. Bowling, L. C., and Tyler, P. A. (1988). Lake Chisholm, a polyhumic forest lake in Tasmania. Hydrobiologia 161, 55–67. Braconnot, P., Otto-Bliesner, B., Harrison, S., Joussaume, S., Peterchmitt, J.Y., Abeouchi, A., Crucifix, M., Driesschaert, E., Fichefet, Th, Hewitt, C.D., Kageyama, M., Kitoh, A., Laîn_ e, A., Loutre, M.F., Marti, O., Merkel, U., Ramstein, G., Valdes, P., Weber, S.L., Yu, Y., Zhao, Y. (2007). Results of PMIP2 coupled simulations of the Mid-Holocene and last glacial maximum e part 1: experiments and large-scale features. Climate of the Past 3, 261-277. Bridgman, H., and Timms, B. V. (2012). Australia, climate and lakes. In ‘Encyclopedia of Lakes and Reservoirs’. (Eds L. Bengtsson, R. W. Herschy and R. W. Fairbridge.) pp. 73–80. (Springer: Dordrecht, The Netherlands.)

148 | P a g e

References Briggs, S. V. (1980). Chemical studies of four swamps on the northern tablelnds New South Wales. Marine and Freshwater Research 31, 729–736. Brodersen, K.P., Anderson, N.J. (2002). Distribution of chironomids (Diptera) in low arctic West Greenland lakes: trophic conditions, temperature and environmental reconstruction. Freshwater Biology 47, 1137–1157. Brodersen, K.P., Odgaard, B.V., Vestergaard, O., Anderson, N.J. (2001). Chironomid stratigraphy in the shallow and eutrophic Lake Søbygaard, Denmark: chironomid-macrophyte co- occurrence. Freshwater Biology 46, 253-267. Brodersen, K.P., Pedersen, O., Lindegaard, C., Hamburger, K. (2004). Chironomids (Diptera) and oxy-regulatory capacity: An experimental approach to paleolimnological interpretation. Limnology and Oceanography 49, 1549-1559. Brodersen, K.P., Pedersen, O., Walker, I.R., Jensen, M.T. (2008). Respiration of midges (Diptera; Chironomidae) in British Columbian lakes: oxy-regulation, temperature and their role as palaeo-indicators. Freshwater Biology 53, 593–602. Brodin, Y.W. (1986). The postglacial history of Lake Flarken, southern Sweden, interpreted from subfossil insect remains. Internationale Revue der gesamten Hydrobiologie. 71, 371-432. Brookes, J. D., and Hamilton, D. (2009). Lakes and reservoirs of Australia and New Zealand. In ‘Encyclopedia of Inland Waters’. (Ed. G. Likens.) pp. 513–524. (Academic Press: Oxford, UK.) Brooks, S.J. (2003) Chironomidae (Insecta: Diptera). In: MacKay A, Battarbee RW and Birks HJB (eds) Global Change in the Holocene. London: Arnold, pp. 328–341. Brooks, S.J., Axford, Y., Heiri, O., Langdon, P.G., Larocque-Tobler, I. (2012). Chironomids can be reliable proxies for Holocene temperatures: A comment on Velle et al. (2010). The Holocene 22, 1495 – 1500 Brooks, S.J., Birks, H.J.B. (2000). Chironomid-inferred Late-glacial air temperatures at Whitrig Bog, southeast Scotland. Journal of Quaternary Science 15, 759–764 Brooks, S.J., Birks, H.J.B. (2001). Chironomid-inferred air temperatures from late-glacial and Holocene sites in north-west Europe: progress and problems. Quaternary Science Reviews 20, 1723–1741. Brooks, S.J., Langdon, P.G., Heiri, O. (2007). The Identification and Use of Palaearctic Chironomidae Larvae in Palaeoecology. Quaternary Research Association. Buckney, R. T., and Tyler, P. A. (1973). Chemistry of Tasmanian inland waters. Internationale Revue der Gesamten Hydrobiologie und Hydrographie 58, 61–78. Carlson, R. E. (1977). A trophic state index for lakes. Limnology and Oceanography 22, 361–369.

149 | P a g e

References Caseldine, C. J.,Turney, C., Long, A.J. (2010). "IPCC and palaeoclimate – an evolving story?" Journal of Quaternary Science 25, 1-4. Chang, J.C., Shulmeister, J., Woodward, C. (2015a). A chironomid based transfer function for reconstructing summer temperatures in south eastern Australia. Palaeogeography Palaeoclimatology Palaeoecology 423, 109-121. Chang, J.C., Shulmeister, J., Woodward, C., Steinberger, L., Tibby, J., Barr, C., (2015b). A chironomid-inferred summer temperature reconstruction from subtropical Australia during the last glacial maximum (LGM) and the last deglaciation. Quaternary Science Reviews 122, 282-292. Chang, J.C., Woodward, C., Shulmeister, J. (2014). A snapshot of the limnology of eastern Australian water bodies spanning the tropics to Tasmania: the land-use, climate, limnology nexus. Marine and Freshwater Research 65, 872-883. Chanson, H., and James, D. P. (1999). Siltation of Australian reservoirs: some observations and dam safety implications. In ‘28th International Association for Hydro-Environment Engineering and Research (IAHR) Congress’. Session B5, 6 pages (Technical University: Graz, Austria) Chao, A. (1987). Estimating the population size for capture–recapture data with unequal catchability. Biometrics 43, 783–791. Chase, M., Bleskie, C., Walker, I.R., Gavin, D.G., Hu, F.S. (2008). Midge-inferred Holocene summer temperatures in Southeastern British Columbia, Canada. Palaeogeography Palaeoclimatology Palaeoecology 257, 244-259. Chivas, A.R., De Deckker, P., Cali, J.A., Chapman, A., Kiss, E., G. Shelley, J.M. (1993). Coupled stable isotope and trace element measurements of lacustrine carbonates as paleoclimatic indicators, pp. 113-121. In: P. K. Swart, K. C. Lohmann, J. Mckenzie and S. Savin. (eds.). Climate change in continental isotopic records. American Geophysical Union, Washington D.C. Clark, I.D., Fritz, P. (1997). Environmental isotopes in hydrogeology. CRC press. Clark, P.U., Dyke, A.S., Shakun, J.D., Carlson, A.E., Clark, J., Wohlfarth, B., Mitrovica, J.X., Hostetler, S.W., McCabe, A.M. (2009). The last glacial maximum. Science 325, 710-714. Coplen, T.B. (1995). Reporting of stable hydrogen, carbon, and oxygen isotopic abundances. Geothermics 24, 707-712. Constantine, J. A., and Dunne, T. (2008). Meander cutoff and the controls on the production of oxbow lakes Geology 36, 23–26.

150 | P a g e

References Cook, E.J., Van der Kaars, S. (2006). Development and testing of transfer functions for generating quantitative climatic estimates from Australian pollen data. Journal of Quaternary Science 21, 723–733. Costin, A. B., Gray, M., Totterdell, C. J., and Wimbush, D. J. (2000). ‘Kosciusko Alpine Flora.’ 2nd edn. (CSIRO Publishing: Melbourne.) Cranston, P.S. (1995). Chironomids: from Genes to Ecosystems, (ed.). CSIRO, Melbourne, pp. 482. Cranston, P.S. (1995). Introduction. pp. 1-7. In: Armitage, P.D., Cranston, P.S. and Pinder, L.C.V. (eds.). The Chironomidae. The biology and ecology of non-biting mideges. Chapman and Hall, London. Cranston, P.S. (2000a). Electronic Guide to the Chironomidae of Australia. http://www. science.uts.edu.au/sasb/chiropage/. Cranston, P.S. (2000b). Three new Chironomidae (Diptera) from the Australian wet tropics. Memoirs of the Queensland Museum 46, 107–127. Cranston, P.S. (2010). Chiro Key. http://chirokey.skullisland.info/. Cranston, P.S., Dimitriadis, S. (2004). The chironomidae (Diptera) larvae of Atherton Tableland lakes, North Queensland, Australia. Memoirs of the Queensland Museum 49, 573-588. Cranston, P.S., Edward, D.H.D. (1999). Botryocladius gen. n.: a new transantarctic genus of orthocladiine midge (Diptera: Chironomidae). System Entomology 24, 305–333. Cranston, P.S., Krosch, M.N. (2015). DNA sequences and austral taxa indicate generic synonymy of Paratrichocladius Santos-Abreu with Cricotopus Wulp (Diptera: Chironomidae). Systematic Entomology. doi: 10.1111/syen.12130 Cranston, P.S., Martin, J. (1989). Family Chironomidae. In: "Catalog of the Diptera of the Australasian and Oceanian regions." (Ed. N.L.Evenhuis) Bishop Museum Press, Honolulu & E.J.Brill, Leiden: 252-274. Davis, J. R., and Koop, K. (2006). Eutrophication in Australian rivers, reservoirs and estuaries – a southern hemisphere perspective on the science and its implications. Hydrobiologia 559, 23–76. D'Costa, D.M., Edney, P., Kershaw, A.P., Deckker, P.D. (1989). Late quaternary palaeoecology of Tower Hill, Victoria, Australia. Journal of Biogeography. 16, 461-482. D'Costa, D.M., Kershaw, A.P. (1997). An expanded recent pollen database from southeastern Australia and its potential for refinement of palaeoclimatic estimates. Australian Journal of Botany 45, 583–605. D'Costa, D.M., Grindrod, J., Ogden, R. (1993). Preliminary environmental reconstructions from late Quaternary pollen and mollusc assemblages at Egg Lagoon, King Island, Bass Strait. Australian Journal of Ecology 18, 351–366. 151 | P a g e

References De Deckker, P. (1983). Australian salt lakes: their history, chemistry, and biota – a review. Hydrobiologia 105, 231–244. De Deckker, P., and Williams, W. D. (1988). Physico-chemical limnology of eleven, mostly saline permanent lakes in western Victoria, Australia. Hydrobiologia 162, 275–286. Deines, P., Wooller, M.J., Grey, J. (2009). Unravelling complexities in benthic food webs using a dual stable isotope (hydrogen and carbon) approach. Freshwater Biology 54, 2243-2251. Development Core Team, R. (2010). R: A Language and Environment for Statistical Computing. R Foundation for Statistical Computing, Vienna. Dieffenbacher-Krall, A.C., Vandergoes, M.V., Denton, G.H. (2007). An inference model for mean summer air temperatures in the Southern Alps, New Zealand, using subfossil chironomids. Quaternary Science Reviews 26, 2487–2504. Dieffenbacher-Krall, A.C., Vandergoes, M.J., Woodward, C.A., Boothroyd, I.K.G. (2008). Guide to Identification and Ecology of New Zealand Subfossil Chironomids Found in Lake Sediment. Climate Change Institute, University of Maine, Orono, ME (http://www.climatechange.umaine.edu/Research/facilities/perl/nzguide.html). Dimitriadis, S., and Cranston, P. S. (2001). An Australian Holocene climate reconstruction using Chironomidae from a tropical volcanic maar lake. Palaeogeography, Palaeoclimatology, Palaeoecology 176, 109–131. Dunbar, G.B., Dickens, G.R. (2003). Massive siliciclastic discharge to slopes of the Great Barrier Reef platform during sea level transgression: constraints from sediment cores between 15°S and 16°S latitude and possible explanations. Sediment Geology 162, 141-158. Edney, P.A., Kershaw, A.P., De Deckker, P. (1990). A late Pleistocene and Holocene vegetation and environmental record from Lake Wangoom, Western Plains of Victoria, Australia. Palaeogeography Palaeoclimatology Palaeoecology 80, 325–343. Eggermont, H., Heiri, O., Russell, J., Vuille, M., Audenaert, L., Verschuren, D. (2010). Chironomidae (Insecta: Diptera) as paleothermometers in the African tropics. Journal of Paleolimnology 43, 413–435. Erez, J. (1978). Vital effect on stable-isotope composition seen in foraminifera and coral skeletons. Nature 273, 199-202. Fletcher, M.S., Thomas, I. (2010). A quantitative late Quaternary temperature reconstruction from western Tasmania, Australia. Quaternary Science Reviews 29, 2351–2361. Freeman, P. (1961). The Chironomidae (Diptera) of Australia. Australian Journal of Zoology 9, 611–738. Galloway, R.W. (1965). Late Quaternary climates in Australia. Journal of Geology 73, 603–618.

152 | P a g e

References Gehre, M., Geilmann, H., Richter, J., Werner, R.A., Brand, W.A. (2004). Continuous flow 2H/1H and 18O/16O analysis of water samples with dual inlet precision. Rapid Communications of Mass Spectrometry 18, 2650-2660. Gentilli, J. (1971). Dynamics of the Australian troposphere. In ‘World Survey of Climatology: Climates of Australia and New Zealand’. (Ed. J. Gentilli.) pp. 53–117.(Elsevier: Amsterdam.) Gibson, J.J., Prepas, E.E., McEachern, P. (2002). Quantitative comparison of lake throughflow, residency, and catchment runoff using stable isotopes: modelling and results from a regional survey of Boreal lakes. Journal of Hydrology 262, 128-144. Glew, J. (1991). Miniature gravity corer for recovering short sediment cores. Journal of Paleolimnology 5, 285–287. Gouramanis, C., De Deckker, P., Switzer, A.D., Wilkins, D. (2013). Cross-continent comparison of high-resolution Holocene climate records from southern Australia - deciphering the impacts of far- field teleconnections. Earth Science Reviews 121, 55-72. Graham, D. J., and Midgley, N. G. (2000). Graphical representation of particle shape using triangular diagrams: an Excel spreadsheet method. Earth Surface Processes and Landforms 25, 1473–1477. Granéli, W., Solander, D. (1988). Influence of aquatic macrophytes on phosphorus cycling in lakes, pp. 245-266. In: G. Persson, M. Jansson. (eds.). Phosphorus in freshwater ecosystems. Springer Netherlands Green, K. (2012). Intra- and inter-annual changes in chemistry of Australian glacial lakes. Marine and Freshwater Research 63, 513–527. Grey J, Kelly A, Jones RI. (2004a). High intraspecific variability in carbon and nitrogen stable isotope ratios of lake chironomid larvae. Limnology and Oceanography 49, 239-240. Grey J, Kelly A, Ward S, Sommerwerk N, Jones RI. (2004b). Seasonal changes in the stable isotope values of lake-dwelling chironomid larvae in relation to feeding and life cycle variability. Freshwater Biology 49, 681-689. Gröcke, D.R., Schimmelmann, A., Elias, S., Miller, R.F. (2006). Stable hydrogen-isotope ratios in beetle chitin: preliminary European data and re-interpretation of North American data. Quaternary Science Reviews 25, 1850-1864. Guildford, S., and Hecky, R. (2000). Total nitrogen, total phosphorus, and putrient limitation in lakes and oceans: is there a common relationship? Limnology and Oceanography 45, 1213– 1223. Haberle, S. G. (2005). A 23 000-yr pollen record from Lake Euramoo, Wet Tropics of NE Queensland, Australia. Quaternary Research 64, 343–356. 153 | P a g e

References Hann, B. J., Warner, B. G. and Warwick, W. F. (1992) Aquatic invertebrates and climate change: A comment on Walker et al. (1991). Canadian Journal of Fisheries and Aquatic Sciences 49: 1274–1276. Harris, G. P. (2001). Biogeochemistry of nitrogen and phosphorus in Australian catchments, rivers and estuaries: effects of land use and flow regulation and comparisons with global patterns. Marine and Freshwater Research 52, 139–149. Harris, S., and Kitchener, A. (2005). ‘From Forest to Fjaeldmark: Descriptions of Tasmania’s Vegetation.’ (Department of Primary Industries, Parks, Water and Environment, Printing Authority of Tasmania: Hobart.) Heinrichs, M. L., Wilson, S. E., Walker, I. R., Smol, J. P., Mathewes, R. W. and Hall, K. J. (1997). Midge and diatom-based palaeosalinity reconstructions for Mahoney Lake, Okanagan Valley, British Columbia, Canada. International Journal of Salt Lake Research 6: 249-267. Heiri, O., Lotter, A.F. (2001). Effect of low count sums on quantitative environmental reconstructions: an example using subfossil chironomids. Journal of Paleolimnology 26, 343–350. Heiri, O., Brooks, S.J., Birks, H.J.B., Lotter, A.F. (2011). A 274-lake calibration data-set and inference model for chironomid-based summer air temperature reconstruction in Europe. Quaternary Science Reviews 30, 3445–3456. Heiri, O., Millet, L., 2005. Reconstruction of late Glacial summer temperatures from chironomid assemblages in Lac Lautrey (Jura, France). Journal of Quaternary Science 20, 33-44. Heltberg, R., Siegel, P.B., Jorgensen, S.L. (2009). Addressing human vulnerability to climate change: Toward a ‘no-regrets’ approach. Global Environmental Change 19, 89-99. Hembrow, S. C., and Taffs, K. H. (2012). Water quality changes in Lake McKenzie, Fraser Island, Australia: a palaeolimnological approach. The Australian Geographer 43, 291–302. Herczeg, A.L., Barnes, C.J., Macumber, P.G., Olley, J.M. (1992). A stable isotope investigation of groundwater-surface water interactions at Lake Tyrrell, Victoria, Australia. Chemical Geology 96, 19-32. Hill, M.O. (1973). Diversity and evenness: a unifying notation and its consequences. Ecology 54, 427- 432. Hofmann, W. (1986). Chironomid analysis. In: Berglund, B.E. (Ed.), Handbook of Holocene Palaeoecology and Palaeohydrology. John Wiley and Sons, Chichester, pp. 715–727. Hogg, A.G., Hua, Q., Blackwell, P.G., Niu, M., Buck, C.E., Guilderson, T.P., Zimmerman, S.R. (2013). SHCal13 Southern Hemisphere calibration, 0-50,000 years Cal BP. Radiocarbon 55, 1889-1903.

154 | P a g e

References Hughes, P.D., Gibbard, P.L., Ehlers, J. (2013). Timing of glaciation during the last glacial cycle: evaluating the concept of a global ‘Last Glacial Maximum’ (LGM). Earth Science Reviews 125, 171-198. IAEA/WMO (2015). Global Network of Isotopes in Precipitation. The GNIP Database. Accessible at: http://www.iaea.org/water; accessed October 2015. Jennings, J. N., and Banks, M. R. (1958). The Pleistocene glacial history of Tasmania. Journal of Glaciology 24, 298–303. Jeschke, J.M., Strayer, D.M. (2008). Usefulness of bioclimatic models for studying climate change and invasive species. Annals of the New York Academy of Sciences, 1134, 1–24. Johannsson, O.E. (1980). Energy Dynamics of the Eutrophic Chironomid f.semireductus from the Bay of Quinte, Lake Ontario. Canadian Journal of Fisheries and Aquatic Sciences 37, 1254-1265. Jones, G. J., and Poplawski, W. (1998). Understanding and management of cyanobacterial blooms in sub-tropical reservoirs of Queensland, Australia. Water Science and Technology 37, 161– 168. Jowsey, P.C. (1966). An improved peat sampler. New Phytologist 65, 245-248. Juggins, S. (2005). C2 Version 1.5: Software for Ecological and Palaeoecological Data Analysis and Visualisation. University of Newcastle, Newcastle-upon-Tyne. Kershaw, A.P. (1986). Climatic change and aboriginal burning in north-east Australia during the last two glacial/interglacial cycles. Nature 322, 47-49. Kershaw, A.P., Bulman, D., Busby, J.R. (1994). An examination of modern and pre-European settlement pollen samples from southeastern Australia- assessment of their application to quantitative reconstruction of past vegetation and climate. Review of Palaeobotany and Palynology 82, 83-96. Kershaw, A.P., Bulman, D. (1996). A preliminary application of the analogue approach to the interpretation of late quaternary pollen spectra from southeastern Australia. Quaternary International 33, 61-71. Kershaw, A.P., D'Costa, D.M., McEwen Mason, J.R.C., Wagstaff, B.E. (1991). Palynological evidence for quaternary vegetation and environments of mainland southeastern Australia. Quaternary Science Reviews 10, 391- 404. Kershaw, A. P., Bretherton, S. C., and Van der Kaars, S. (2007). A complete pollen record of the last 230 ka from Lynch’s Crater, north-eastern Australia. Palaeogeography, Palaeoclimatology, Palaeoecology 251, 23–45.

155 | P a g e

References Kim, N., Han, J.K. (1997). Assessing the integrity of cross-validation: a case for small sample- based research. Working Paper Series no. MKTG 97.096. Hong Kong University of Science and Technology, Department of Marketing. Kottek, M., Grieser, J., Beck, C., Rudolf, B., Rubel, F. (2006). World map of the Köppen–Geiger climate classification updated. Meteorologische Zeitschrift 15, 259–263. Krosch, M.N. (2011). Phylogeography of Echinocladius martini Cranston (Diptera: Chironomidae) in closed forest streams of eastern Australia. Australian Journal of Entomology 50, 258–268. Krosch, M.N., Cranston, P.S., Baker, A.M., Vink, S. (2015). Molecular data extend Australian Cricotopus midge (Chironomidae) species diversity, and provide a phylogenetic hypothesis for biogeography and freshwater monitoring. Zoological Journal of the Linnean Society. doi: 10.1111/zoj.12284 Kureck, A. (1979). Two circadian eclosio times in Chironomus thummi (Diptera), alternatively selected with different temperatures. Oecologia 40, 311–323. Lamb, H., Roberts, N., Leng, M., Barker, P., Benkaddour, A., van der Kaars, S. (1999). Lake evolution in a semi-arid montane environment: response to catchment change and hydroclimatic variation. Journal of Paleolimnology 21, 325-343. Larocque, I. (2001). How many chironomid head capsules are enough? A statistical approach to determine sample size for palaeoclimatic reconstructions. Palaeogeography Palaeoclimatology Palaeoecology 172, 133–142. Larocque, I., Finsinger, W. (2008). Late-glacial chironomid-based temperature reconstructions for Lago Piccolo di Avigliana in the southwestern Alps (Italy). Palaeogeography, Palaeoclimatology, Palaeoecology 257, 207-223. Larocque, I., Hall, R.I., Grahn, E. (2001). Chironomids as indicators of climate change: a 100-lake training set from a subarctic region of northern Sweden (Lapland). Journal of Paleolimnology 26, 307–322. Larocque-Tobler, I., Heiri, O., Wehrli, M. (2010). Late Glacial and Holocene temperature changes at Egelsee, Switzerland, reconstructed using subfossil chironomids. Journal of Paleolimnology 43, 649-666. Laycock, J.W. (1975). Hydrology of North Stradbroke Island. Proceedings of Royal Society of Queensland 86, 15-19. Ledru, M.P., Stevenson, J., 2012. The rise and fall of the genus Araucaria: a southern hemisphere climatic connection. In: Haberle, S.G., David, B. (Eds.), Peopled Landscapes: Archaeological and Biogeographic Approaches to Landscapes, Terra Australis, vol. 34. ANU ePress, Canberra, Australia, pp. 241-254.

156 | P a g e

References Leng, M. J., Lamb, A. L., Heaton, T. H., Marshall, J. D., Wolfe, B. B., Jones, M.D., Holmes, J.A., Arrowsmith, C. (2006), pp. 147-184. In: M.J. Leng (eds.). Isotopes in lake sediments, Isotopes in Palaeoenvironmental Research, Springer, Netherlands Leng, M.J., Marshall, J.D. (2004). Palaeoclimate interpretation of stable isotope data from lake sediment archives. Quaternary Science Reviews 23, 811-831. Levesque, A. J., Mayle, F. E., Walker, I. R. and Cwynar, L. C. (1993). A previously unrecognized late-glacial cold event in eastern North America. Nature 361: 623‐626. Little, J.L., Smol, J.P. (2001). A chironomid-based model for inferring late-summer hypolimnetic oxygen in southeastern Ontario lakes. Journal of Paleolimnology 26, 259–270. Lotter, A.F., Walker, I.R., Brooks, S.J., Hofmann, W. (1999). An intercontinental comparison of chironomid palaeotemperature inference models: Europe vs North America. Quaternary Science Reviews 18, 717–735. Luoto, T.P. (2009). Subfossil Chironomidae (Insecta: Diptera) along a latitudinal gradient in Finland: development of a new temperature inference model. Journal of Quaternary Science 24, 150–158. Marchant, R., Hirst, A., Norris, R., Metzeling, L. (1999). Classification of macroinvertebrate communities across drainage basins in Victoria, Australia: consequences of sampling on a broad spatial scale for predictive modelling. Freshwater Biology 41, 253–268. Massaferro, J., Larocque-Tobler, I. (2013). Using a newly developed chironomid transfer function for reconstructing mean annual air temperature at Lake Potrok Aike, Patagonia, Argentina. Ecological Indicators 24, 201–210. Massaferro, J., Larocque-Tobler, I., Brooks, S.J., Vandergoes, M., Dieffenbacher-Krall, A., Moreno, P. (2014). Quantifying climate change in Huelmo mire (Chile, Northwestern Patagonia) during the last Glacial termination using a newly developed chironomid-based temperature model. Palaeogeography Palaeoclimatology Palaeoecology 399, 214-224. Mata, M.M., Wijffels, S.E., Church, J.A., Tomczak, M. (2006). Eddy shedding and energy conversions in the East Australian Current. Journal of Geophysical Research: Oceans 111, C09034. Mayr, C., Laprida, C., Lücke, A., Martín, R.S., Massaferro, J., Ramón-Mercau, J., Wissel, H. (In press.) Oxygen isotope ratios of chironomids, aquatic macrophytes and ostracods for lake- water isotopic reconstructions – Results of a calibration study in Patagonia. Journal of Hydrology McComb, A. J., and Davis, J. A. (1993). Eutrophic waters of southwestern Australia. Fertilizer Research 36, 105–114.

157 | P a g e

References McGowan, H.A., Petherick, L.M., Kamber, B.S. (2008). Aeolian sedimentation and climate variability during the late quaternary in southeast Queensland, Australia. Palaeogeography Palaeoclimatology Palaeoecology 265, 171-181. Medeiros, A.S., Quinlan, R. (2011). The distribution of the Chironomidae (Insecta: Diptera) along multiple environmental gradients in lakes and ponds of the eastern Canadian Arctic. Canadian Journal of Fisheries and Aquatic Sciences 68, 1511–1527. Miller, G.H., Magee, J.W., Jull, J.T. (1997). Low-latitude glacial cooling in the southern hemisphere from amino-acid racemization in emu eggshells. Nature 385, 241–244. Morris, D. P., and Lewis, W. M. (1988). Phytoplankton nutrient limitation in Colorado mountain lakes. Freshwater Biology 20, 315–327. Moss, B. (1998). ‘Ecology of Freshwaters. Man and Medium, Past to Future.’ (Blackwell: London.) Moss, P., Tibby, J., Petherick, L.M., McGowan, H.A., Barr, C. (2013). Late quaternary vegetation history of North Stradbroke Island, Queensland, eastern Australia. Quaternary Science Reviews 74, 257-272. Nielson, K.E., Bowen, G.J. (2010). Hydrogen and oxygen in brine shrimp chitin reflect environmental water and dietary isotopic composition. Geochimica et Cosmochimica Acta 74, 1812-1822. Oksanen, J., Blanchet, F.G., Kindt, R., Legendre, P., O'Hara, B., Simpson, G.L., Solymos, P., Stevens, M.H.H., Wagner, H. (2010). Vegan: Community Ecology Package. R Package Version 2.0-10. http://CRAN.R-project.org/package=vegan. Osmulski, P., Leyko, W. (1986). Structure, function and physiological role of chironomus haemoglobin. Comparative Biochemistry and Physiology, Part B: Biochemistry and Molecular Biology 85, 701-722. Olander, H., Birks, H.J.B., Korhola, A., Blom, T. (1999). An expanded calibration model for inferring lake water and air temperatures from fossil chironomid assemblages in northern Fennoscandia. The Holocene 9, 279–294. Olander, H., Korhola, A., Blom, T. (1997). Surface sediment Chironomidae (Insecta: Diptera) distributions along an ecotonal transect in subarctic Fennoscandia: developing a tool for palaeotemperature reconstructions. Journal of Paleolimnology 18: 45–59. Oliver, D.R. (1971). Life History of the Chironomidae. Annual Review of Entomology 16, 211-230. Page, M.C., Dickens, G.R., Dunbar, G.B. (2003). Tropical view of quaternary sequence stratigraphy: siliciclastic accumulation on slopes east of the Great Barrier Reef since the last glacial maximum. Geology 31, 1013-1016. Pearson, E.J., Juggins, S., Talbot, H.M., Weckstr€ om, J., Ros_ en, P., Ryves, D.B., Roberts, S.J., Schmidt, R. (2011). A lacustrine GDGT-temperature calibration from the Scandinavian 158 | P a g e

References Arctic to Antarctic: renewed potential for the application of GDGT-paleothermometry in lakes. Geochimica et Cosmochimica Acta 75, 6225-6238. Pedro, J.B., Rasmussen, S.O., van Ommen, T.D. (2012). Tightened constraints on the time-lag between Antarctic temperature and CO2 during the last deglaciation. Climate of the Past 8, 1213-1221. Petherick, L., Bostock, H., Cohen, T.J., Fitzsimmons, K., Tibby, J., Fletcher, M.S., Moss, P., Reeves, J., Mooney, S., Barrows, T., Kemp, J., Jansen, J., Nanson, G., Dosseto, A. (2013). Climatic records over the past 30 ka from temperate Australia — a synthesis from the Oz- INTIMATE workgroup. Quaternary Science Reviews 74, 58–77. Petherick, L.,M., McGowan, H.A., Moss, P.T. (2008). Climate variability during the last Glacial maximum in eastern Australia: evidence of two stadials? Journal of Quaternary Science 23, 787-802. Petherick, L.M., McGowan, H.A., Kamber, B.S. (2009). Reconstructing late quaternary dust transport pathways for eastern Australia using trace element composition of long travelled dusts. Geomorphology 105, 67-79. Petherick, L.M., Moss, P.T., McGowan, H.A. (2011). Climatic and environmental variability during the Late Glacial Maximum in eastern Australia: a review. Australian Journal of Earth Sciences 58, 563–577. Pinder, L.C.V., Reiss, F. (1983). The larvae of Chironominae (Diptera: Chironomidae) of the Holarctic region. In: Wiederholm, T. (Ed.), Chironomidae of Holarctic Region - Keys and Diagnoses (Part 1-Larvae), vol. 19. Entomologica Scandinavica Supplement, pp. 293-436. Porch, N., Jordan, G.J., Price, D.M., Barnes, R.W., Macphail, M.K., Pemberton, M. (2009). Last interglacial climates of south-eastern Australia: plant and beetle-based reconstructions from Yarra Creek, King Island, Tasmania. Quaternary Science Reviews 28, 3197–3210. Porinchu, D.F., MacDonald, G.M. (2003). The use and application of freshwater midges (Chironomidae: Insecta: Diptera) in geographical research. Progress in Physical Geography 27, 378–422. Porinchu, D.F., MacDonald, G.M., Bloom, A.M., Moser, K.A. (2002). The modern distribution of chironomid sub-fossils (Insecta: Diptera) in the Sierra Nevada, California: potential for paleoclimatic reconstructions. Journal of Paleolimnology 28, 355–375. Porinchu, D., Rolland, N., Moser, K. (2009). Development of a chironomid-based air temperature inference model for the central Canadian Arctic. Journal of Paleolimnology 41, 349–368. Potito, A.P., Woodward, C.A., McKeon, M., Beilman, D.W. (2014). Modern influences on chironomid distribution in western Ireland: potential for palaeoenvironmental reconstruction. Journal of Paleolimnology 52, 385–404. 159 | P a g e

References Quinlan, R., Smol, J.P. (2001). Setting minimum head capsule abundance and taxa deletion criteria in chironomid based inference models. Journal of Paleolimnology 26, 327–342. Rees, A.B.H., Cwynar, L.C.(2010a). Evidence for early postglacial warming in Mount Field National Park, Tasmania. Quaternary Science Reviews 29, 443–454. Rees, A.B.H., Cwynar, L.C.(2010b). A test of Tyler’s Line – response of chironomids to a pH gradient in Tasmania and their potential as a proxy to infer past changes in pH. Freshwater Biology 55, 2521-2540. Rees, A.B.H., Cwynar, L.C., Cranston, P.S. (2008). Midges (Chironomidae, Ceratopogonidae, Chaoboridae) as a temperature proxy: a training set from Tasmania, Australia. Journal of Paleolimnology 40, 1159–1178. Reeves, J.M., Bostock, H.C., Ayliffe, L.K., Barrows, T.T., De Deckker, P., Devriendt, L.S., Dunbar, G.B., Drysdale, R.N., Fitzsimmons, K.E., Gagan, M.K. (2013). Palaeoenvironmental change in tropical Australasia over the last 30,000 years - A synthesis by the OZ-INTIMATE group. Quaternary Science Reviews 74, 97-114. Renwick, W. H. (2009). Lake and Reservoirs of North America. In ‘Encyclopedia of Inland Waters’. (Ed. G. Likens.) pp. 524–532. (Academic Press: Oxford, UK.) Ridgway, K.R., Dunn, J.R. (2003). Mesoscale structure of the mean East Australian current system and its relationship with topography. Progress in Oceanography 56, 189-222. Ridgway, K.R., Godfrey, J.S. (1997). Seasonal cycle of the East Australian Current. Journal of Geophysical Research: Oceans 102, 22921 - 22936. Ridgway, K., Hill, K. (2009). The East Australian Current. In: Poloczanska, E.S., Hobday, A.J., Richardson, A.J. (Eds.), A Marine Climate Change Impacts and Adaptation Report Card for Australia 2009, ISBN 978-1-921609-03-9. NCCARF Publication 05/09. Riebeek, H. (2006). Paleoclimatology: Understanding past climate to predict future. Retrieved 20/6, 2015, from http://earthobservatory.nasa.gov/Features/Paleoclimatology_Understanding/. Roemmich, D., Gilson, J., Davis, R., Sutton, P., Wijffels, S., Riser, S. (2007). Decadal spinup of the South Pacific subtropical Gyre. Journal of Physical Oceanography 37, 162 - 173. Rosenzweig, C., Karoly, D., Vicarelli, M., Neofotis, P., Wu, Q., Casassa, G., Menzel, A., Root, T.L., Estrella, N., Seguin, B., Tryjanowski, P., Liu, C., Rawlins, S., Imeson, A. (2008). Attributing physical and biological impacts to anthropogenic climate change. Nature 453, 353-357. Rovira, J. L., and Pardo, P. (2006). Nutrient pollution of waters: eutrophication trends in European marine and coastal environments. Contributions in Science 3, 181–186.

160 | P a g e

References Roy, P.S., Thorn, B.G. (1981). Late quaternary marine deposition in New South Wales and southern Queensland d An evolutionary model. Journal of Geological Society of Australia 28, 471-489. Rozanski, K.,Araguás-Araguás, L., Gonfiantini. R. (1993). Isotopic patterns in modern global precipitation. Geophyscial Monograph Series 78, 1–36. Ryan, T. A., Joiner, B. L., and Ryan, B. F. (2004). MinitabTM. In ‘Encyclopedia of Statistical Sciences’. (Ed. S. Kotz., B.R. Campbell., N. Balakrishnan., B. Vidakovic.) (John Wiley & Sons, Inc.: Hoboken, NJ.) Sachs, H. M., Webb III, T. and Clark, D. R. (1977). Paleoecological transfer functions. Annual Review of Earth and Planetary Sciences, 5, 159 – 178 Scanes, P., Coade, G., Doherty, M., and Hill, R. (2007). Evaluation of the utility of water quality based indicators of estuarine lagoon condition in NSW, Australia. Estuarine, Coastal and Shelf Science 74, 306–319. Schakau, B. (1991). Stratigraphy of the fossil Chironomidae (Diptera) from Lake Grasmere, South Island, New Zealand, during the last 6000 years. Hydrobiologia 214, 213–221. Schakau, B. (1993). Palaeolimnological studies on sediments from Lake Grasmere, South Island, New Zealand, with special reference to the Chironomidae (Diptera). University of Canterbury, New Zealand, pp. 203–286 (Chapter 9). Schmidt, G. A. (2010). Enhancing the relevance of palaeoclimate model/data comparisons for assessments of future climate change. Journal of Quaternary Science 25, 79- 87. Shapiro, S. S., and Wilk, M. B. (1965). An analysis of variance test for normality (complete samples). Biometrika 52, 591–611. Samartin, S., Heiri, O., Vescovi, E., Brooks, S.J., Tinner, W. (2012). Lateglacial and early Holocene summer temperatures in the southern Swiss Alps reconstructed using fossil chironomids. Journal of Quaternary Science 27, 279-289. Schimmelmann, A., Miller, R.F., Leavitt, S.W. (1993). Hydrogen isotopic exchange and stable isotope ratios in cellulose, wood, chitin, and amino compounds, pp. 367-347. In: P. K. Swart, K. C. Lohmann, J. Mckenzie and S. Savin. (eds.). Climate change in continental isotopic records. American Geophysical Union, Washington D.C. Smith, V. (2003). Eutrophication of freshwater and coastal marine ecosystems a global problem. Environmental Science and Pollution Research International 10, 126–139. Smith, V. H., Tilman, G. D., and Nekola, J. C. (1999). Eutrophication: impacts of excess nutrient inputs on freshwater, marine, and terrestrial ecosystems. Environmental Pollution 100, 179– 196.

161 | P a g e

References Sniderman, J.M.K., Porch, N., Kershaw, A.P. (2009). Quantitative reconstruction of Early Pleistocene climate in southeastern Australia and implications for atmospheric circulation. Quaternary Science Reviews 28, 3185–3196. Søballe, D. M., and Kimmel, B. L. (1987). A large-scale comparison of factors influencing phytoplankton abundance in rivers, lakes, and impoundments. Ecology 68, 1943–1954. Speich, S., Blanke, B., de Vries, P., Drijfhout, S., D€ o€ os, K., Ganachaud, A., Marsh, R. (2002). Tasman leakage: a new route in the global ocean conveyor belt. Geophysical Research Letters 29, 55-51-55-54. Stenni, B., Buiron, D., Frezzotti, M., Albani, S., Barbante, C., Bard, E., Barnola, J.M., Baroni, M., Baumgartner, M., Bonazza, M., Capron, E., Castellano, E., Chappellaz, J., Delmonte, B., Falourd, S., Genoni, L., Iacumin, P., Jouzel, J., Kipfstuhl, S., Landais, A., Lemieux-Dudon, B., Maggi, V., Masson-Delmotte, V., Mazzola, C., Minster, B., Montagnat, M., Mulvaney, R., Narcisi, B., Oerter, H., Parrenin, F., Petit, J.R., Ritz, C., Scarchilli, C., Schilt, A., Schupbach, S., Schwander, J., Selmo, E., Severi, M., Stocker, T.F., Udisti, R. (2011). Expression of the bipolar see-saw in Antarctic climate records during the last deglaciation. Nature Geoscience 4, 46-49. Sturman, A. P., and Tapper, N. J. (2006) ‘The Weather and Climate of Australia and New Zealand.’ 2nd edn. (Oxford University Press: Melbourne.) Tejan-Kella, M.S., Chittleborough, D.J., Fitzpatrick, R.W., Thompson, C.H., Prescott, J.R., Hutton, J.T. (1990). Thermoluminescence dating of coastal sand dunes at Cooloola and North Stradbroke Island, Australia. Australian Journal of Soil Research 28, 465-481. Telford, R.J., Birks, H.J.B. (2011). A novel method for assessing the statistical significance of quantitative reconstructions inferred from biotic assemblages. Quaternary Science Reviews 30, 1272–1278. ter Braak, C.J.F., Juggins, S., 1993. Weighted averaging partial least squares regression (WA-PLS): an improved method for reconstructing environmental variables from species assemblages. Hydrobiologia 269, 485–502. ter Braak, C. J. F., and Smilauer, P. (2002). ‘CANOCO Version 4.5.’ (Biometris–Plant Research International: Wageningen, The Netherlands.) Tibby, J. (2004). Development of a diatom-based model for inferring total phosphorus in south- eastern Australian water storages. Journal of Paleolimnology 31, 23–36. Tibby, J., Haberle, S. (2007). A late glacial to present diatom record from lake Euramoo, wet tropics of Queensland, Australia. Palaeogeography Palaeoclimatology Palaeoecology 251, 46–56.

162 | P a g e

References Timms, B. V. (1973). A limnological survey of the freshwater coastal lakes of east Gippsland, Victoria. Marine and Freshwater Research 24, 1–20. Timms, B. V. (1974). Aspects of the limnology of Lake Tali Karng, Victoria. Marine and Freshwater Research 25, 273–279. Timms, B. V. (1977). A study of some coastal dune lakes in western Victoria. Proceedings of the Royal Society of Victoria 89, 167–172. Timms, B. V. (1982). Coastal dune waterbodies of north-eastern New South Wales. Marine and Freshwater Research 33, 203–222. Timms, B. V. (1992). ‘Lake Geomorphology.’ (Gleneagles: Adelaide.) Tokeshi, M. (1995). Life cycles and population dynamics. Chironomidae: biology and ecology of non-biting midges (Ed. by Armitage et al.), pp. 225-268. Chapman & Hall, London Townsend, S. A. (1999). The seasonal pattern of dissolved oxygen, and hypolimnetic deoxygenation, in two tropical Australian reservoirs. Lakes and Reservoirs: Research and Management 4, 41–53. Trabucco, A., and Zomer, R. J. (2009). ‘Global Aridity Index (Global-Aridity) and Global Potential Evapo-transpiration (Global-PET) Geospatial Database.’ (CGIAR Consortium for Spatial Information.) Available at http://www.csi.cgiar.org [accessed August, 2013]. (Washington, DC, USA) Troedson, A.L. (1997). Late Quaternary Sedimentation on the East Australian Continental Slope: Responses to Palaeoenvironmental Change (PhD thesis). Department of Geology and Geophysics, University of Sydney. Troedson, A.L., Davies, P.J. (2001). Contrasting facies patterns in subtropical and temperate continental slope sediments: inferences from east Australian late quaternary records. Marine Geology 172, 265-285. Turney, C.S.M., Haberle, S., Fink, D., Kershaw, A.P., Barbetti, M., Barrows, T.T., Black, M., Cohen, T.J., Corr ege, T., Hesse, P.P., Hua, Q., Johnston, R., Morgan, V., Moss, P., Nanson, G., van Ommen, T., Rule, S., Williams, N.J., Zhao, J.X., D'Costa, D., Feng, Y.X., Gagan, M., Mooney, S., Xia, Q. (2007). Integration of ice core, marine and terrestrial records for the Australian last Glacial maximum and Termination: a contribution from the OZ-INTIMATE group. Journal of Quaternary Science 21, 751-761. Turney, C.S.M., Kershaw, A.P., James, S., Branch, N., Cowley, J., Fifield, L.K., Jacobsen, G., Moss, P. (2006). Geochemical changes recorded in Lynch's Crater, northeastern Australia, over the past 50 ka. Palaeogeography Palaeoclimatology Palaeoecology 233, 187-203. Turney, C. S. M., Kershaw, A. P., Moss, P., Bird, M. I., Fifield, L. K., Cresswell, R. G., Santos, G. M., Di Tada, M. L., Hausladen, P. A., and Zhou, Y. (2001). Redating the onset of burning at 163 | P a g e

References Lynch’s Crater (north Queensland): implications for human settlement in Australia. Journal of Quaternary Science 16, 767–771. United Nations Environment Programme (UNEP) (1997). ‘World Atlas of Desertification.’ 2nd Edition (Routledge : London.) van Hardenbroek M, Heiri O, Grey J, Bodelier PLE, Verbruggen F, Lotter AF. (2010). Fossil chironomid δ13C as a proxy for past methanogenic contribution to benthic food webs in lakes? Journal of Paleolimnology 43, 235–245. van Hardenbroek, M., Lotter, A.F., Bastviken, D., Andersen, T.J., Heiri, O. (2014). Taxon-specific δ13C analysis of chitinous invertebrate remains in sediments from Strandsjön, Sweden. Journal of Paleolimnology 52, 95-105. Vandergoes, M.J., Dieffenbacher-Krall, A.C., Newnham, R.M., Denton, G.H., Blaauw, M., 2008. Cooling and changing seasonality in the Southern Alps, New Zealand during the Antarctic Cold Reversal. Quaternary Science Reviews 27, 589–601. Vanhoutte, K., Verleyen, E., Kilroy, C., Sabbe, K., Dasseville, R., and Vyverman, W. (2006). Catchment characteristics and chemical limnology of small lakes, tarns and mire pools in New Zealand (South Island) and Tasmania. Marine and Freshwater Research 57, 83–93. Velle, G., Brodersen, K.P., Birks, H.J.B., Willassen, E. (2010). Midges as quantitative temperature indicator species: Lessons for palaeoecology. The Holocene 20, 989-1002. Velle, G., Brodersen, K.P., Birks, H.J., Willassen, E. (2012). Inconsistent results should not be overlooked: A reply to Brooks et al. (2012). The Holocene 22, 1501 - 1508 Verbruggen, F, Heiri, O., Reichart, G.J., De Leeuw, J.W., Nierop, K.G.J., Lotter, A.F. (2010a).

Effects of chemical pretreatments on δ18O measurements, chemical composition, and morphology of chironomid head capsules. Journal of Paleolimnology 43, 857–872.

Verbruggen F, Heiri O, Reichart GJ, Lotter AF. (2010b). Chironomid δ18O as a proxy for past lake

water δ18O: a Lateglacial record from Rotsee (Switzerland). Quaternary Science Reviews 29, 2271-2279. Verbruggen F, Heiri O, Reichart GJ, Blaga C, Lotter AF. (2011). Stable oxygen isotopes in chironomid and cladoceran remains as indicators for lake-water δ18O. Limnology and Oceanography 56, 2071–2079. Von Grafenstein, U., Erlernkeuser, H. & Trimborn, P. (1999). Oxygen and carbon isotopes in modern fresh-water valves: assessing vital offsets and autecological effects of interest for palaeoclimate studies. Palaeogeography, Palaeoclimatology, Palaeoecology 148, 133-152. Walker, I.R. (1987). Chironomidae (Diptera) in paleoecology. Quaternary Science Reviews 6, 29– 40. 164 | P a g e

References Walker, I.R. (2002). In: Smol, J.P., Birks, H.J.B., Last, W.M. (Eds.), Midges: Chironomidae and Related Diptera Tracking Environmental Change Using Lake Sediments. Springer Netherlands. Walker, I.R., Cwynar, L.C. (2006). Midges and palaeotemperature reconstruction—the North American experience. Quaternary Science Reviews 25, 1911–1925. Walker, I. R., Levesque, A. J., Cwynar, L. C. and Lotter, A. F. (1997). An expanded surface-water palaeotemperature inference model for use with fossil midges from eastern Canada. Journal of Paleolimnology 18: 165‐178. Walker, I., Mathewes, R. (1989). Early postglacial chironomid succession in southwestern British Columbia, Canada, and its paleoenvironmental significance. Journal of Paleolimnology 2, 1- 14. Walker, I. R., Mott, R. J. and Smol, J. P. (1991). Allerod-Younger Dryas lake temperatures from midge fossils in Atlantic Canada. Science 253: 1010-1012. Walker, I.R., Paterson, C.G. (1985). Efficient separation of subfossil Chironomidae from lake sediments. Hydrobiologia 122, 189–192. Walker, I. R., Smol, J. P., Engstrom, D. R. and Birks, H. J. B. (1991). An assessment of Chironomidae as quantitative indicators of past climatic change. Canadian Journal of Fisheries and Aquatic Sciences. 48: 975-987. Walker, I. R., Wilson, S. E. and Smol, J. P. (1995). Chironomidae (Diptera): Quantitative palaeosalinity indicators for lakes of western Canada. Canadian Journal of Fisheries and Aquatic Sciences 52: 950-960. Walther, G.R., Post, E., Convey, P., Menze, 1, A., Parmesan, C., Beebee, T.J.C., Fromentin, J.M., Hoegh-Guldberg, O. & Bairlein, F. (2002) Ecological responses to recent climate change. Nature 416, 389–395. Wang, Y., Francis, D.R., O’Brien, D.M., Wooller, M.J. (2008). A protocol for preparing subfossil chironomid head capsules (Diptera: Chironomidae) for stable isotope analysis in paleoclimate reconstruction and considerations of contamination sources. Journal of Paleolimnology 40, 771-781. Wang, Y.V., O’Brien, D.M., Jenson, J., Francis, D., Wooller, M.J. (2009). The influence of diet and water on the stable oxygen and hydrogen isotope composition of Chironomidae (Diptera) with paleoecological implications. Oecologia 160, 225-233. Ward, W.T. (2006). Coastal dunes and strandplains in southeast Queensland: sequence and chronology. Australian Journal of Earth Sciences 53, 363 - 373. Warwick, W. F. (1989). Chironomids, lake development and climate: A commentary. Journal of Paleolimnology 2: 15–17. 165 | P a g e

References Wassenaar, L.I., Hobson, K.A. (2003). Comparative equilibration and online technique for determination of non-exchangeable hydrogen of keratins for use in migration studies. Isotopes in Environmental and Health Studies 39, 211-217. White, E., Law, K., Payne, G., and Pickmere, S. (1985). Nutrient demand and availability among planktonic communities – an attempt to assess nutrient limitation to plant growth in 12 central volcanic plateau lakes. New Zealand Journal of Marine and Freshwater Research 19, 49–62. Williams, M., Cook, E., van der Kaars, S., Barrows, T., Shulmeister, J., Kershaw, P. (2009). Glacial and deglacial climatic patterns in Australia and surrounding regions from 35,000 to 10,000 years ago reconstructed from terrestrial and near shore proxy data. Quaternary Science Reviews 28, 2398- 2419. Williams, W. D., Walker, K. F., and Brand, G. W. (1970). Chemical composition of some inland surface waters and lake deposits of New South Wales, Australia. Marine and Freshwater Research 21, 103–116. Wilson, S. E., Walker, I. R., Mott, R. J. and Smol, J. P. (1993). Climatic and limnological changes associated with the Younger Dryas period in Atlantic Canada. Climate Dynamics 8: 177- 187. WMO (World Meteorological Organization), (2015). Developing climate knowledge. Retrieved 05. 10. 15., from http://www.wmo.int/worldmetday/content/knowledge Woltering, M., Atahan, P., Grice, K., Heijnis, H., Taffs, K., Dodson, J. (2014). Glacial and Holocene terrestrial temperature variability in subtropical east Australia as inferred from branched GDGT distributions in a sediment core from Lake Mckenzi. Quaternary Research 82, 132-145. Woodward, C.A., Shulmeister, J. (2006). New Zealand chironomids as proxies for human induced and natural environmental change: transfer functions for temperature and lake production (chlorophyll a). Journal of Paleolimnology 36, 407–429. Woodward, C.A., Shulmeister, J. (2007). Chironomid-based reconstructions of summer air temperature from lake deposits in Lyndon Stream, New Zealand spanning the MIS 3/2 transition. Quaternary Science Reviews 27, 142–154. Wooller, M.J., Francis, D., Fogel, M.L., Miller, G.H., Walker, I.R., Wolfe, A.P. (2004). Quantitative paleotemperature estimates from δ18O of chironomid head capsules preserved in arctic lake sediments. Journal of Paleolimnology 31, 267–274. Wooller, M., Wang, Y., Axford, Y. (2008). A multiple stable isotope record of Late Quaternary limnological changes and chironomid paleoecology from northeastern Iceland. Journal of Paleolimnology 40, 63-77. 166 | P a g e

References Wright, I.A., Burgin, S. (2007). Species richness and distribution of eastern Australian lake chironomids and chaoborids. Freshwater Biology 52, 2354–2368. Zachos, J., Pagani, M., Sloan, L., Thomas, E., Billups, K. (2001). Trends, rhythms, and aberrations in global climate 65 Ma to present. Science 292, 686-693. Zar, J. H. (1999). ‘Biostatistical Analysis.’ (Prentice Hall: Upper Saddle River, NJ.) Zhang, Z., Sachs, J.P. (2007). Hydrogen isotope fractionation in freshwater algae: I. Variations among lipids and species. Organic Geochemistry 38, 582-608. Ziveri, P., Stoll, H., Probert, I., Klaas, C., Geisen, M., Ganssen, G., Young, J. (2003). Stable isotope ‘vital effects’ in coccolith calcite. Earth and Planetary Science Letters 210, 137-149.

167 | P a g e

Appendices

Appendices

APPENDIX 1. A SUMMARY OF PREVIOUS MODERN CALIBRATION TRAINING SETS DEVELOPED AND APPLIED FOR PALEO-RECONSTRUCTIONS USING CHIRONOMIDS

168 | P a g e

Appendix 1. Review of past training set Region No. of Lakes Application Other Correlations Number of Model Authors and Taxa coefficient Year of ( r2) publication Europe Across north- 30 lakes used Lake water temperature Maximum lake depth is 32 identified and r2= 0.5 (Olander et al., western Finland in lake surface transfer function strongly correlated (only 20 taxa used in (Jackknifing) 1997, Olander water developed based on limited environmental the lake water RMSEP = 1.5– et al., 1999) temperature chironomid assemblage variables were considered temperature 1.6°C transfer (Olander et al., 1997) in Olander et al., 1997). transfer function function Expanded calibration Sediment organic content, Extended to 38 Expanded to model for maximum lake depth, and taxa included for 53 lakes for inferring lake water and lake water temperature are July air July air July air temperatures the most important temperature temperature from fossil chironomid explanatory variables model model assemblages is developed (Olander et al., 1999) (Olander et al., 1999) Subarctic region 100 Mean July air temperature LOI, maximum lake depth 85 r2= 0.65 (Larocque et of northern transfer function was and mean January air 42 taxa used in (Jackknifing) al., 2001) Sweden developed based on temperature are also the model RMSEP = chironomid assemblage strongly correlated 1.13°C Greenland 43 Statistical analysis were The trophic variables 24 - (Brodersen and used to relate chironomid (total nitrogen and total Anderson, taxa to the environmental phosphorus (TN, TP))and 2002) variables Temperature Switzerland 81 A July air temperature July air temperature is the 76 r2= 0.81 (Heiri et al., (Swiss Alps) inference model (transfer only constraining variable 50 taxa included (Bootstrapping) 2003) function) was developed in the model RMSEP = based on chironomid 1.51°C assemblage Iceland 50 Mean July air temperature The chironomid 54 r2=0.66 (Langdon et al., transfer function assemblages is influenced 53 taxa used in (Jackknifing) 2008) developed by chironomid by Lake substrate (%TC) the model RMSEP = (Holmes et al., assemblage 1.095°C 2009).

169 | P a g e

Appendix 1. Review of past training set Region No. of Lakes Application Other Correlations Number of Model Authors and Taxa coefficient Year of ( r2) publication Finland 77 Mean July air temperature Sampling depth, dissolved 110 r2=0.78 (Luoto, 2009) inference model was oxygen and conductivity 72 taxa used in (Jackknifing) developed based on are relatively significant the model RMSEP = chironomid assemblage 0.721°C Sweden, 31 Relationships of Summer surface water 106 - (Verbruggen et Denmark, temperatures, trophic and July air temperature, al., 2011a) Germany, the state, and oxygen as well as total Netherlands, between chironomid phosphorus (TP) Austria, assemblages. concentrations, Switzerland, Stable oxygen isotopes in hypolimnetic oxygen France and Italy chironomid head capsules availability and is measured in all sites conductivity were found and a relationship was statistically significant developed between temperature and stable oxygen isotopic composition of chironomid head capsules. Europe 274 Mean July air temperature Larger range of both lake 131 taxa found r2= 0.87 (Heiri et al., (Norway and (Norway:153, models were developed types and pH are included in the Norwegian (Bootstrapping) 2011) Switzerland) Switzerland: for both Norwegian and lakes, 120 taxa RMSEP = 1.4°C (Largest 102) Swiss calibration data sets found in Swiss calibration set respectively, as well as a Alps so far) combined model of these 154 taxa two data sets. included North western Asia Tibetan Plateau 42 A chironomid-based Salinity is the most 30 r2 = 0.80 (Zhang et al., (North-western salinity inference model significant 24 taxa included (Jackknifing) 2007) China) was built. RMSEP = 0.29 log10 TDS 170 | P a g e

Appendix 1. Review of past training set Region No. of Lakes Application Other Correlations Number of Model Authors and Taxa coefficient Year of ( r2) publication North America South eastern 50 Transfer function for the Dissolved inorganic 67 r2 = 0.58 (Little and Ontario hypolimnetic oxygen carbon, the Anoxic 45 taxa included (Jackknifing) Smol, 2001) conditions of lakes, which Factor, maximum depth in the transfer RMSEP = 0.032 measured as average end- and total phosphorus function log(x + 19.88) of-summer hypolimnetic concentrations were also AvgDO(summ) dissolved oxygen correlated with (AvgDO(Summ) ) was assemblage composition constructed based on chironomid assemblages. Central, eastern 57 A taxon response model Surface water 68 - (Porinchu et al., Sierra Nevada of was used to assess the temperature, elevation, 44 taxa included 2002) California statistical relationship of depth, strontium, in the transfer each taxon to surface water particulate organic carbon function temperature. Quantitative is statistically significant transfer functions were variables. developed to estimate surface water temperature from the chironomid assemblages. Canadian Arctic 50 Chironomid assemblage lake morphometry, pH, 32 - (Gajewski et Islands was developed in these nutrients (TP) and al., 2005) lakes and statistical models temperature are identified were used to relate taxa to as statistical important the environment. factors Quebec, Canada 52 Mean August air Lake depth, dissolved 97 r2=0.67 Larocque, 2006 temperature transfer organic carbon (DOC), 64 taxa included (Jackknifing) function was developed are significant also in the model RMSEP = based on chironomid variables. development 1.17°C assemblage

171 | P a g e

Appendix 1. Review of past training set Region No. of Lakes Application Other Correlations Number of Model Authors and Taxa coefficient Year of ( r2) publication Eastern North 145 Chironomid taxa Lake depth, arctic tundra 80 r2=0.82 (Barley et al., America assemblage was developed vegetation, alpine tundra 62 taxa included (Bootstrapping) 2006) (Yukon/Northwest to the training set. Midge vegetation, pH, dissolved in the model RMSEP = *Expanded Territories, British inference models for mean organic carbon, lichen development 1.46°C results of Columbia and the July air temperature and woodland vegetation and Gajewski et al., Canadian transformed depth were surface area are all (2005) Arctic Islands) developed. significant to midge distribution Central Canadian 88 Quantitative midge based Maximum depth, pH, total 50 included in r2 = 0.77 (Porinchu et al., Arctic inference model was nitrogen-unfiltered (TN- the models (Jackknifing) 2009) developed for average July UF), Cl and Al are RMSEP = air temperature and statistically significant 1.03°C summer surface water temperature. Eastern Canadian 51 (Walker et Statistical analysis were Temperature and trophic 86 r2 = 0.88 (Walker et al., Arctic al., 1997) used to relate chironomid status were found to 32 (Jacknifing) 1997; 63 (Brodersn taxa assemblages to the strongly influence the RMSEP = Brodersen et et al., 2008) environmental variables chironomid assemblages 2.26°C al., 2008, (Walker et al., Medeiros and 1997) Quinlan, 2011) Southern Hemisphere New Zealand 46 The first chironomid based Chlorophyll a is also 50 r2 = 0.77 (Woodward (North and South transfer function build in significantly correlated. Retention of all (Jackknifing) and Island) the southern hemisphere. A Conductivity, Total species RMSEP = Shulmeister, temperature inference Nitrogen (TN) and pH are 1.31°C 2006; 2007) model was built with the also statistically For temperature most productive lakes significant. model removed and a Chlorophyll r2 = 0.49 a. transfer function based (Jackknifing) on chironomid RMSEP = 0.46 -1 assemblages was Log 10μg l 172 | P a g e

Appendix 1. Review of past training set presented. For Chlorophyll a model Region No. of Lakes Application Other Correlations Number of Model Authors and Taxa coefficient Year of ( r2) publication New Zealand 60 Taxon response models Chlorophyll a, 49 r2 = 0.8 (Dieffenbacher- (South Island) were constructed and a conductivity, organic (Jackknifing) Krall et al., mean summer air content of sediment and RMSEP = 2007; temperature transfer mean annual precipitation 1.29°C Vandergoes et function was developed are also significant al., 2008) variables Tasmania 54 Taxon response models pH, annual radiation, 49 non-rare taxa r2 = 0.72 (Rees et al., (Australia) were constructed and a magnesium, annual were used in (Jackknifing) 2008; Rees and midge transfer function precipitation, SiO2, and model RMSEP = 0.94 Cwynar, 2010) was developed for depth are also significant development temperature of the warmest factors. quarter (TWARM) of Tasmania. Uganda and 65 Taxon response models Water depth, conductivity 81 r2 = 0.81–0.97 (Eggermont et Kenya (Africa) and transfer functions of (salinity) and pH are 68 taxa used in (Jackknifing) for al., 2010) Mean Air Temperature important for EL-EM the models Mean Air (MAT) and Surface Water calibration data set. development Temperature Temperature (SWT) were DOC, LOI of the surface RMSEP = 0.61- developed. sediments and nutrients 1.50°C (defined as TN, TP) are important for the r2 = 0.77–0.95 Rwenzori-only Data set (Jackknifing) for (29 lakes). Surface Water Temperature RMSEP = 1.39- 1.98°C

173 | P a g e

Appendix 1. Review of past training set South America 46 Transfer function for Secchi depth and 49 r2 = 0.56 (Massaferro (Argentina and warmest months mean conductivity are also (Jackknifing) and Larocque- Chile) temperature was developed significantly correlated RMSEP = Tobler, 2012; and applied 1.69°C Massaferro et al., 2014)

174 | P a g e

Appendix 2. Photographs of chironomid taxa identified from modern eastern Australian waterways

APPENDIX 2. CHIRONOMID TAXA ENUMERATED FROM THE MODERN CALIBRATION DATA SET OF SOUTHEASTERN AUSTRALIA

• Chironomid taxa listed from 1 to 38 correspond to taxa specified in Table 4.5, Chapter 4. These are the taxa used in the South-eastern Australia transfer function development. • The ‘red box’ shows the key part of each taxa to be used for identification

175 | P a g e

Appendix 2. Photographs of chironomid taxa identified from modern eastern Australian waterways

1. Chironomus Meigen

a. Lake Cootapatamba, Kosciuszko, NSW b. Freshwater Lake, Victoria

Ecology: occurs in 94% of the lakes in the training set. It is abundant in both warm eutrophic and cool oligotrophic lakes; mostly confined to the profundal but may also be present in the littoral; tolerant of low oxygen concentrations and even anoxia; tolerant of low pH and high salinity; often an early coloniser after significant environmental change where it may occur in suboptimal conditions; detritivores and filter-feeders; associated with soft sediments (Brooks et al., 2007). These descriptions match with the obervations from this training set.

2. Polypedilum nubifer Skuse

a. Nuggetty Gully Reservoir, Victoria, 20 X b. Lake Poona, QLD, 100 X

Ecology: occurs in 82% of the lakes in the training set. It is often an indicator of temperate climatic conditions. It occurs in the litteral of eutrophic lakes and some species occur amongst vegetation (Brooks et al., 2007). It shows a strong correlation with temperature in the southeastern Australian training set. However, some morphotypes are also found in subtropical lakes.

176 | P a g e

Appendix 2. Photographs of chironomid taxa identified from modern eastern Australian waterways

3. Cladopelma Kieffer

b. Lake Cantani, Mt Buffalo, Victoria, 40 X a. Lake Poona, QLD, 40 X

Ecology: occurs in 70% of the lakes in the training set. It is a littoral taxon occurring in muddy and sand/gravel substrates. It is often a warm stenotherm but not particularly tolerant of high nutrient conditions, typical of mesotrophic lakes (Brooks et al., 2007). These descriptions match the observations from the southeast Australian training set but it shows a particularly strong correlation with water depth.

4. Dicrotendipes Kieffer

a. Nuggetty Gully Reservoir, Victoria, 20 X b. Lake Terangpom, Victoria, 40 X

Ecology: occurs in 70% of the lakes in the training set. It occurs in the littoral of lentic

environments where it is often associated with macrophytes and mesotrophic to eutrophic waters. Most of the training sets indicate that it is a warm indicator (Brooks et al., 2007). However in this training set, it is more clearly as a eutrophic and shallow water indicator.

177 | P a g e

Appendix 2. Photographs of chironomid taxa identified from modern eastern Australian waterways

5. Polypedilum spp. Kieffer

a. Thirlmere Lakes, NSW, 40 X b. Lake Lila, Tasmania, 40 X

Ecology: it is a morphotype of Polypedilum, occurs in 18% of the lakes in the training set. The ecology is not clearly known, it seems occur in a few reservoirs and a few natural lakes with in and outflows in the current training set. It is possibly related to flowing water or stream species.

6. Cryptochironomus Kieffer

a. Lake Elingamite, Victoria, 20 X b. Lake Tooliorook, Victoria, 40 x

Ecology: occurs in 24% of the lakes in the training set. It often occurs in the profundal, in nutrient rich waters and apparently can be acidophilic. Larvae occur in a variety of substrates in the littoral and sublittoral (Brooks et al., 2007). It shows a close relationship with conductivity in the southeastern Australian training set.

178 | P a g e

Appendix 2. Photographs of chironomid taxa identified from modern eastern Australian waterways

7. Parachironomus Lenz

a. Lake Fyans, Victoria, 40 X b. Lake Tooliorook, Victoria, 40 X

Ecology: occurs in 33% of the lakes in the training set. It usually occurs in the littoral of

standing waters and also running water, some morphotypes are asscociated with the presence of macrophytes. The genus is descrbed as acidophobic (Buskens 1987; Brooks et al., 2007). It shows correlation with temperature, macrophytes, lake productivity in the southeastern

Australian training set. It shows as a warm temperature indicator.

8. Kiefferulus martini Freeman

a. Freshwater Lake, Victoria, 20 X b. Lake Samuel, Tasmania, 20 X

Ecology: occurs in 50% of the lakes in the training set. Larvae of Kiefferulus inhabit

sediments of small water bodies. Three species of this genus is known from Australia (Cranston, 2010). It shows correlation with macrophytes and water depth in the southeastern Australian training set.

179 | P a g e

Appendix 2. Photographs of chironomid taxa identified from modern eastern Australian waterways

9. Procladius Skuse

a. Lake Albina, Kosciusko, NSW, 40 X b. Freshwater Lake, Victoria, 40 X

Ecology: occurs in 97% of the lakes in the training set. There are many species and morphotypes within this genus. The genus is abundant in mesotrophic and eutrophic lakes but may be absent from very cold lakes. It does not tolerant anoxia conditions. It tolerates acidic conditions (Brooks et al., 2007). These descriptions match the observations from the southeastern Australian training set in general. It shows as a warm temperature indicator in the training set.

10. Coelopynia pruinosa Freeman

a. Green Lake, Victoria, 10 X b. Grahamstown Lake, NSW, 40 X

Ecology: occurs in 25% of the lakes in the training set. Larvae are usually found in deep lakes (e.g. in Tasmania), billabongs in Australia, and in both standing and slow flowing waters (Northern Territory and Western Australia) (Cranston, 2010). It shows correlation with temperature in the southeastern Australian training set and appears as a warm stenotherm.

180 | P a g e

Appendix 2. Photographs of chironomid taxa identified from modern eastern Australian waterways

11. Pentaneurini undifferentiated

a. Eagle Tarn, Tasmania, 20 X b. Lake Hiawatha, NSW, 20 X

Ecology: ocurrs in 50% of the lakes in the training set. Cranston (2010) divided the genus to A-E. It has a wide range of distribution in Australia, from monsoonal tropics to Tasmania and west to the east. The ecology of its distribution is not well identified. In the southeastern Australian training set, it shows a correlation with conductivity.

12. Riethia Kieffer

a. sp. 1 - Grahamstown Lake, NSW, 40 X b. sp. 2 – Grahamstown Lake, NSW, 40 X

Ecology: occurs in 64% of the lakes in the training set. In Australasia Riethia larvae are found in depositional areas in low order streams, often naturally shaded by riparian vegetation. Records range from New Zealand’s temperate North Island, southern New

Caledonia, Tasmania and along the eastern margin of Australia to monsoonal tropical Northern Territory, and in temperate southern Western Australia. Some species occur in habitats characterised by high amounts of fine particle organic matter in both lotic and

lenthic systems (Cranston, 2010). It occurs in most of the reservoirs in the training set. It shows a correlation with conductivity in southeastern Australian lakes.

181 | P a g e

Appendix 2. Photographs of chironomid taxa identified from modern eastern Australian waterways

13. Tanytarsus lugens type

a. Lake Cantani, Mt Buffalo, Victoria, 20 X b. Green Lake, Victoria, 40 X

Ecology: it occurs in 85% of the lakes. The lugens type is a cold stenotherm and usually occurs in the profundal of oligotrophic lakes or in the littoral of cold subalpine and subarctic lakes (Brooks et al., 2007). The morphotype appears in the southeastern Australian lakes shows close features to the lugens type, but it does not show clear correlations with any particular environmental variables.

14. Tanytarsus Pallidicornis type

a. Lake Poona, QLD, 40 X b. Nuggetty Gully Reservoir, Victoria, 40 X

Ecology: occurs in 80% of the lakes in the training set. Most of the Tanytarsus morphotypes usually occur in the littoral of relatively warm, productive lakes and will tolerate acidic condtions (Brooks et al., 2007). The morphotype appears in the southeastern Australian lakes shows close correlation with lake productivity.

182 | P a g e

Appendix 2. Photographs of chironomid taxa identified from modern eastern Australian waterways

15. Tanytarsus glabrescens type

a. Grahamstown lake, NSW, 40 X b. Eagle Tarn, Tasmania, 40 X

Ecology: occurs in 50% of the lakes in the training set. As described above, the morphotype usually occurs in the littoral of relatively warm, productive lakes and will tolerate acidic condtions (Brooks et al., 2007). The morphotype appears in the southeastern Australian lakes

shows close correlation with nutrients loading, lake productivity and pH.

16. Paratanytarsus Skuse

a. Lake Terangpom, Victoria, 40 X b. Grahamstown Lake, NSW, 20 X

Ecology: occurs in 67% of the lakes in the training set. It may be abundant in warm or cold lakes. Some morphotypes occur in cold oligotrophic lakes at high latitude or altitude whereas, others are typical of warmer conditions. It is often assicated with macrophytes (Brooks et al., 2007). In the the southeastern Australian lakes, it shows close correlation with macrophytes, conductivity and appears as a warm indicator.

183 | P a g e

Appendix 2. Photographs of chironomid taxa identified from modern eastern Australian waterways

17. Tanytarsus undifferentiated

a. Lake Selina, Tasmania, 40 X b. Lake Lila, Tasmania, 40 X

It is required a more detailed identification guide to further refine the ecology of the genus of Tanytarsus in southeastern Australian lakes.

18. Tanytarsus lactescens type

a. Lake Poona, QLD, 40 X b. Lake Hiawatha, NSW, 40 X

Ecology: occurs in 27% of the lakes in the training set. It is usually more abundant in temperate, carbonate lakes (Heiri, 2001). In the the southeastern Australian lakes, the related morphotype shows close correlation with water depth, lake productivity and pH. It also

shows as a warm temperature indicator.

184 | P a g e

Appendix 2. Photographs of chironomid taxa identified from modern eastern Australian waterways

19. Tanytarsus nr chinyensis

a. Lake Lila, Tasmania, 40 X b. Lake Plimsoll Pond, Tasmainia, 40 X

Ecology: occurs in 15% of the lakes in the training set. It usually appears as a cold stenotherm from oligotrophic lakes but in Scandinavia it occurs at the warm end of the thermal gradient (Brooks et al., 2007). In the the southeastern Australian lakes, the related

morphotype shows close correlation with temperature and pH. It appears as a cold temperature indicator in southeastern Australia.

20. Stempellina Thienemann & Bause

a. Lake Selina, Tasmania, 20 X b. Lake Lea Pond, Tasmania, 20 X

Ecology: only occurs in two Tasmanian lakes in the training set. It is a warm stenotherm most abundant in oligotrophic lakes. Since it only appears in two lakes, it is difficult to summarize the ecology. However, it seems to be abundant in the Tasmanian training set (Rees et al., 2008).

185 | P a g e

Appendix 2. Photographs of chironomid taxa identified from modern eastern Australian waterways

21. Harnischia Kieffer

a. Reedy Lake, Victoria, 40 X b. Bamberang Reservoir, NSW, 40 X

Ecology: occurs in 12% of the lakes in the training set. It is usually associated with warm eutrophic lakes and the genus is described as acidophobic (Brooks et al., 2007). In the the southeastern Australian lakes, it shows a correlation with temperature and also as a warm stenotherm.

22. Paralimnophytes type 1 unofficial morphotype

a. Lake Mombeong, Victoria, 100 X b. Lake Selina, Tasmania 40 X

Ecology: occurs in 40% of the lakes in the training set. For the genus: It is usually associated with very shallow water in the littoral of lakes and with streams. The genus can sometimes be a useful indicator of lake level fluctuations. Some species are associated with aquatic macrophytes, others are terrestrial or semiterrestrial. The genus usually occurs in temperate waters. Larvae of Paralimnophyes can be found in eutrophic lowland pools (Brooks et al., 2007). This morphotype (type 1) described in the southeastern Australian training set is correlated with macrophytes, conductivity, pH and shows as a cold stenotherm.

186 | P a g e

Appendix 2. Photographs of chironomid taxa identified from modern eastern Australian waterways

23. Paralimnophytes type 2 unofficial morphotype

a. Lake Plimsoll Pond, Tasmania, 40 X b. Lake Wellums, NSW, 40 X

Ecology: occurs in 40% of the lakes in the training set. The description of the genus is shown in 22 above. In the the southeastern Australian lakes, this morphotype (type 2) shows as a close correlation with macrophytes.

24. Paralimnophytes type 3 unofficial morphotype

a. Thirlmere Lakes, NSW, 40 X b. Green Lake, Victoria, 40 X

Ecology: occurs in 27% of the lakes in the training set. The description of the genus is shown in 22 above. In the the southeastern Australian lakes, this morphotype (type 3) does not show any clear correlation with the tested environmental variables.

187 | P a g e

Appendix 2. Photographs of chironomid taxa identified from modern eastern Australian waterways

25. Parakiefferiella undifferentiated

a. Green Lake, Victoria, 40 X b. Lake Terangpom, Victoria, 40 X Ecology: occurs in 20% of the lakes in the training set. For the genus: species of Parakiefferiella are often abundant in the littoral of temperate lakes. Some morphotypes are associated with warmer and more eutrophic lakes than others. The distribution ranges from profundal to littoral or sublittoral of shallow lakes. Some morphotypes are cold stenotherms occurring in oligotrophic lakes (Brooks et al., 2007).

The undifferentiated species of the genus Parakiefferiella in the southeastern Australian lakes shows correlation with water depth.

26. Parakiefferiella type 1 unofficial type

a. Green Lake, Victoria, 40 X b. Lake Mombeong, Victoria, 40 X

Ecology: only occurs in 3 lakes in the training set. The description of the genus is shown in 25 above. In the the southeastern Australian lakes, this morphotype (type 1) does not show any clear correlation with the tested environmental variables.

188 | P a g e

Appendix 2. Photographs of chironomid taxa identified from modern eastern Australian waterways

27. Parakiefferiella type 2 unofficial type

a. Lake Selina, Tasmania, 40 X b. Green Lake, Victoria, 40 X

Ecology: occurs in 4 lakes in the training set. The description of the genus is shown in 25 above. In the the southeastern Australian lakes, this morphotype (type 2) shows correlation with temperature, macrophytes, productivity, pH and conductitivy. It is identified as a cold stenotherm.

28. Parakiefferiella type 3 unofficial type

a. Lake Lea Pond, Tasmania, 20 X b. Lake Selina, Tasmania, 40 X

Ecology: occurs in 4 lakes in the training set. The description of the genus is shown in 25 above. In the the southeastern Australian lakes, this morphotype (type 3) shows correlation with temperature, macrophytes, pH and conductivity. It is also identified as a cold stenotherm.

189 | P a g e

Appendix 2. Photographs of chironomid taxa identified from modern eastern Australian waterways

29. Botryocladius Cranston & Edward

a. Green lake, Victoria, 40 X b. Little Llangothlin Lagoon, NSW, 40 X

Ecology: occurs in 27% of the lakes in the training set. Larval Botryocladius occur in a range of mostly temperate habitats, ranging from glacial-fed lakes (Patagonia), to moderate elevation lakes (Patagonia, Tasmania, s.e. Australia), to rivers and creeks (Patagonia and s.e. Australia) to ephemeral streams (southern and western Australia) (Cranston and Edward, 1999). It shows correlation with temperature in the southeastern Australian training set and appears as a cold stenotherm.

30. Smittia Homgren

a. Lake Terangpom, Victoria, 40 X b. Green lake, Victoria, 40

Ecology: only occurs in 3 lakes in the training set. Most species of Smittia are terrestrial although the genus may be aquatic in the littoral and splash zone of lakes or its presence may be indicative of erosion events (Brooks et al., 2007).

This matches with the observation from the southeastern Australian lakes. This genus does not show any clear correlation with the tested environmental variables.

190 | P a g e

Appendix 2. Photographs of chironomid taxa identified from modern eastern Australian waterways

31. Gymnometriocnemus type 1 unofficial morphotype

a. Little Llangothlin Lagoon, NSW, 40 X b. Blue Lake, Kosciuszko, NSW, 40 X Ecology: occurs in 7 lakes in the training set. The larvae of the genus is said to be terrestrial, living in woodland soil (Cranston et al., 1983), so their presence in lake sediment may be indicative of erosion events. This matches with the observation from the southeastern Australian lakes, such as in Little Llangothlin Lagoon. This genus does not show any clear correlation with the tested environmental variables.

32. Genus Australia

a. Blue Lake, Kosciuszko, NSW, 40 X b. Lake Fortuna, Tasmania, taken by Andrew Rees (shared on October, 2013)

Ecology: only occurs in 2 lakes in the training set. This genus occurs sporadically in running waters, riffles and waterfalls of eastern Australia, from Far North Queensland south to Tasmania. Pupae have been found in New Zealand, in several moderate to large rivers in both the South Island and far northern North Island (Cranston, 2010). This genus does not show any clear correlation with the tested environmental variables.

191 | P a g e

Appendix 2. Photographs of chironomid taxa identified from modern eastern Australian waterways

33. Thienemanniella Kieffer

a. Spp 1. Little Llangothlin Lagoon, NSW, 40 X b. Spp 2. Lake Cartcarrong, Victoria, 40 X

Ecology: occurs in 4 lakes in the training set. The genus usually occurs in streams and rivers but is also found in the sediment of temperate lakes, often at cool temperature sites and can be relatively common in some Icelandic lakes. In the southeastern Australian training set, this genus shows correlation with temperature, macrophytes, water depth, lake productivity and conductivity. It is identified as a cold stenotherm.

34. Orthoclad type 1 unofficial morphotype

a. Little Llangothlin Lagoon, NSW, 40 X b. Lake Reedy, Victoria, 40 X

Ecology: occurs in 3 lakes in the training set. The three lakes are lowland sites in souteastern Australia. The species morphotype is not previously well described. The ecology is not well understood. In the southeastern Australian training set, it shows correlation with macrophytes.

192 | P a g e

Appendix 2. Photographs of chironomid taxa identified from modern eastern Australian waterways

35. Orthoclad type 4 unofficial morphotype

b. Wombat Pool, Tasmania, taken by Andrew a. Lake Cootaptamba, Kosciuszko, NSW, 40 X Rees (shared on October, 2013)

Ecology: occurs in 8 lakes in the training set. The species morphotype was not previously well recognized. It appears in glacial lakes of Tasmania and Kosciuszko, southeastern Australia. It does not show any clear correlation with the tested environmental variables in the southeastern Australian training set.

36. Pseudosmittia type 2 unofficial morphotype

a. Lake Lea Pond, Tasmania, 40 X b. Lake Cantani, Mt Buffalo, NSW, 40 X

Ecology: only occurs in 2 lakes in the training set. Many species in the genus are terrestrial or semi-terrestrial but some occur in the littoral and splash zone of lakes and are associated with aquatic macrophytes. In the southeastern Australian training set, this genus shows correlation with temperature,

lake productivity and conductivity. It is identified as a cold stenotherm.

193 | P a g e

Appendix 2. Photographs of chironomid taxa identified from modern eastern Australian waterways

37. Kosciuszko Orthoclad type 1

a. Lake Cootaptamba, Kosciuszko, NSW, 40 X b. Lake Cygnus, Tasmania, taken by Andrew Rees (shared on October, 2013)

Ecology: occurs in 3 lakes in the training set. It is recognized and named ‘SO2’ by Cranston (2010). It was found in eastern Australia and Tasmania (Rees et al., 2008). It appearance is restricted to montane tarns. The ecology is not described previously. It shows correlation

with lake productivity in the southeastern Australian training set.

38. Cricotopus ‘parbicintus’

b. Lake Esperance, Tasmania, taken by a. Blue Lake, Kosciuszko, NSW, 40 X Andrew Rees (shared on October, 2013) Ecology: only occurs in 2 lakes in the training set. At genus level, it is eurytopic and can be found in flowing and standing waters. The genus is ubiquitous in lake sediment samples. Most lentic taxa are associated with temperate, relatively eutrophic conditions. Many taxa are littoral and are associated with vegetation, some being leaf and stem miners. A number of species are cold stenotherms and occur in oligotrophic cool waters (Brooks et al., 2007). In the southeastern Australian training set, this genus shows correlation with temperature, macrophytes, water depth, lake productivity and pH. It is also identified as a cold stenotherm.

194 | P a g e

Appendix 3. Photographs of fossil chironomid taxa identified from Welsby Lagoon

APPENDIX 3. CHIRONOMID FOSSIL TAXA ENUMERATED FROM WELSBY LAGOON, NORTH STRADBROKE ISLAND, AUSTRALIA FROM THE LAST GLACIAL MAXIMUM TO NEAR PRESENT

• Chironomid taxa listed from 1 to 21 correspond to taxa specified in the chironomid stratigraphy diagram displayed in FIGURE 5.3, Chapter 5. • The ‘red box’ shows the key part of each taxon to be used for identification.

195 | P a g e

Appendix 3. Photographs of fossil chironomid taxa identified from Welsby Lagoon

1. Procladius

a. 20 X (Holocene) b. 20 X (Glacial) It is identified as a warm stenotherm in the modern training set and shows increase in abundance from the LGM to deglaciation in the Welsby Lagoon record.

2. Parachironomus

a. Sp 1. 40 X (Holocene) b. Sp 2. 20 X (Holocene) It is identified as a warm stenotherm in the modern training set and it is absent in the

samples that were inferred as cooler periods.

3. Cladopelma Kieffer

a. 40 X (Holocene) b. 40 X (Glacial)

It is identified as a warm and shallow water indicator in the modern training set and shows increase in abundance from the LGM to deglaciation in the Welsby Lagoon record.

196 | P a g e

Appendix 3. Photographs of fossil chironomid taxa identified from Welsby Lagoon

4. Paratanytarsus

a. 40 X (Glacial) b. 40 X (Glacial) It is identified as a warm stenotherm in the modern training set and shows increase in abundance from the LGM to deglaciation in the Welsby Lagoon record.

5. Tanytarsus lactescens type

a. 40 X (Glacial) b. 40 X (Glacial)

It is identified as a warm stenotherm in the modern training set and shows increase in abundance from the LGM to deglaciation in the Welsby Lagoon record.

6. Cryptochironomus Kieffer

a. 20 X (Glacial) b. 40 X (Glacial)

It is identified as correlated with conductivity in the modern training set. It is absent before

18 cal ka BP and appears from the deglaciation in the Welsby Lagoon record.

197 | P a g e

Appendix 3. Photographs of fossil chironomid taxa identified from Welsby Lagoon

7. Tanytarsus glabrescens type

a. 40 X (Holocene) b. 40 X (Holocene)

It is identified as correlated with macrophytes, productivity and pH in the modern training set. It shows apparent increase in abundance from the LGM to deglaciation in the Welsby Lagoon record.

8. Pentaneurini

20 X (Holocene) b. 40 X (Glacial) It is identified as correlated with conductivity in the modern training set and shows increase in abundance from the LGM to deglaciation in the Welsby Lagoon record.

9. Tanytarsus lugens type

a. 40 X (Glacial) b. 40 X (Holocene) It does not show clear correlation with the tested variables in the modern training set. It only appears after 19 cal ka BP in the Welsby Lagoon record.

198 | P a g e

Appendix 3. Photographs of fossil chironomid taxa identified from Welsby Lagoon

10. Tanytarsus pallidicornis type

a. 40 X (Holocene) b. 100 X (Glacial)

It is identified as correlated with lake productivity in the modern training set. It shows low abundance at 18.5and 22 cal ka BP and high abundance in the rest of samples of the LGM and deglaciation in the Welsby Lagoon record.

11. Chironomus

a. Spp 1. 40 X (Holocene) b. Spp 2. 40 X (Holocene) It does not show clear correlation with the tested variables in the modern training set. No apparent trend is observed from the LGM to the deglaciation in the Welsby Lagoon record, but the abundance is apparently high in the two Holocene samples.

12. Dicrotendipes

a. Spp 1. 40 X (Holocene) b. Spp 2. 40 X (Glacial) It is identified as correlated with lake productivity, macrophytes, depth and conductivity in the modern training set. It shows high abundance and is dominant the assemblages in the four Holocene samples but much lower abundance in the LGM and deglaciation samples.

199 | P a g e

Appendix 3. Photographs of fossil chironomid taxa identified from Welsby Lagoon

13. Polypedilum

a. 40 X (Holocene) b. 40 X (Glacial) It is often identified as a temperate climate indicator in modern training sets although some morphotype species appear in subtropical lakes in Australia. It shows decrease in abundance from the LGM to deglaciation in the Welsby Lagoon record.

14. Tanytarsus nr Chinyensis

a. 100 X (Glacial) It is identified as a cold stenotherm in the modern training set. It does not show apparent trend from the LGM to the deglaciation in the Welsby Lagoon record.

15. Stempellina

a. 100 X (Glacial) b. 100X (Glacial)

It is identified as correlated with macrophytes, productivity and pH in the modern training set and also a cold stenotherm. It shows decrease in abundance from the LGM to deglaciation and is absent from the Holocene in the Welsby Lagoon record. It was not found in modern subtropical lakes.

200 | P a g e

Appendix 3. Photographs of fossil chironomid taxa identified from Welsby Lagoon

16. Paralimnophyes unofficial morphotype 1

a. 40 X (Glacial) b. 40 X (Glacial) It is identified as correlated with macrophytes, conductivity and pH in the modern training set and also a cold stenotherm. It does not show apparent changes in abundance from the LGM to deglaciation in the Welsby Lagoon record, but it is absent throughout the Holocene and it was not found in modern subtropical lakes.

17. Paralimnophyes unofficial morphotype 2

a. 100 X (Glacial) b. 40 X (Glacial)

It is identified as correlated with macrophytes in the modern training set. It shows low abundance and only in 3 samples from the deglaciation in the Welsby Lagoon record. It is absent throughout the Holcene and it was not found in modern subtropical lakes.

18. Paralimnophyes unofficial morphotype 3

a. 40 X (Glacial) b. 100 X (Glacial) It does not show clear correlation with the tested variables in the modern training set. It shows low abundance in 4 samples from the LGM and deglaciation in the Welsby Lagoon record. It is absent throughout the Holcene and it was not found in modern subtropical lakes.

201 | P a g e

Appendix 3. Photographs of fossil chironomid taxa identified from Welsby Lagoon

19. Parakiefferiella nr. unofficial morphotype 3

a. Spp. 1. 40 X (Glacial) b. Spp .2. 100 X

It is identified as correlated with macrophytes,(Glacial) conductivity and pH in the modern training set and also a cold stenotherm. It does not show apparent changes in abundance from the LGM to deglaciation in the Welsby Lagoon record, but it is absent throughout the Holocene and it was not found in modern subtropical lakes.

20. Pseudosmittia type 2 unofficial morphotype

a. Spp .1. 40 X (Glacial) b. Spp. 2. 40 X (Glacial) It is identified as correlated with productivity and conductivity in the modern training set and also a cold stenotherm. It only appears from the start of deglaciation in the Welsby Lagoon record, but it is absent throughout the Holocene and it was not found in modern subtropical lakes.

21. Genus Australia

a. 40 X (Glacial) b. 40 X (Glacial)

It does not show clear correlation with the tested variables in the modern training set. It has low abundance in 4 of the LGM and deglaciation samples. It is absent throughout the Holocene.

202 | P a g e

Appendix 4. Published peer-reviewed journal articles during candidature

APPENDIX 4. PUBLISHED JOURNAL ARTICLES DURING CANDIDATURE

Paper 1, included as Chapter 3:

A snapshot of the limnology of eastern Australian water bodies spanning the tropics to Tasmania: the land-use, climate, limnology nexus

Jie Christine Chang A B , Craig Woodward A and James Shulmeister A

A School of Geography, Planning and Environmental Management, The University of Queensland, St Lucia, Brisbane, Qld 4072, Australia. B Corresponding author. Email: [email protected]

Marine and Freshwater Research 65(10) 872-883 http://dx.doi.org/10.1071/MF13265 Submitted: 10 October 2013 Accepted: 20 December 2013 Published: 7 July 2014 Hyperlink to full-text: http://www.publish.csiro.au/view/journals/dsp_journal_fulltext.cfm?nid=126&f=MF13265

Paper 2, included as Chapter 4:

Hyperlink to full-text: http://www.sciencedirect.com/science/article/pii/S0031018215000486 203 | P a g e

Appendix 4. Published peer-reviewed journal articles during candidature Paper 3, included as Chapter 5:

Hyperlink to full-text: http://www.sciencedirect.com/science/article/pii/S0277379115300147

204 | P a g e