<<

Spatiotemporal variation in the - pre- and post- extinction of the Bowen and Galilee basins ()

Alexander Thomas Wheeler BSc. (Hons.) () – Rhodes University MSc. (Geology) – University of Pretoria

https://orcid.org/0000-0003-1677-6789

A thesis submitted for the degree of Doctor of Philosophy at The University of Queensland in 2020 School of Earth and Environmental Sciences Abstract

Palynology has been used as an important tool for exploration and correlation in the late Permian measures of the Galilee and Bowen basins. It has seen less application in studying the floras of these basins and the significant effect of the end-Permian extinction (EPE) on the environment and vegetational ecology. The main objective of this thesis is to understand how depositional environment controlled the distribution and variation in floral assemblages in the late Permian, during the EPE and in the earliest Triassic using palynology as a for vegetation. A marker mudstone was previously examined in the Bowen Basin and is thought to represent a P-T Boundary . This thesis also examined a potential correlative of this marker mudstone in the Galilee Basin and assesses its temporal relationship to the EPE and its significance for the interpretation of the end-Permian palaeoenvironment.

A secondary objective of this work was to test the efficacy of an acid-free palynological processing technique against the conventional processing that employs strong acids (HCl and HF). This technique was previously only used on Cenozoic-aged material and this represents the first application of it on Palaeozoic-aged samples. Samples were sent for conventional processing to a commercial palynological laboratory and splits of the same samples were processed using the acid- free technique at The University of Queensland. Results show large differences in the palynofacies assemblages between the samples, however both processing techniques yielded relatively comparable and well preserved palynomorph assemblages. The acid-free processing appears to yield lower concentrations of palynomorphs relative to organic debris, but also preserves a higher abundance and more diverse assemblage of algae in these samples.

To examine floral variation in the late Permian, samples were selected from five localities representing a shift from a proximal, terrestrial environment to a distal, marginal-marine and shallow- marine environment. Palynological samples were collected from the last Permian marine incursion at the base of the Black Alley Shale up to the top of the Bandanna Formation, which is thought to represent the Permian-Triassic Boundary. Palynological assemblages were assigned putative botanical affinities and examined relative to the lithostratigraphy of each section in order to understand temporal and spatial variance in floras at each locality. are dominant in lower delta plain settings along with accessory horsetails and lycophyte spores, suggesting that these pteridophytes formed the pioneer flora on prograding delta lobes, and also made up the understory vegetation in a paludal/lacustrine setting. Glossopteris becomes more dominant on the upper

I delta plain and alluvial plain, and also makes up a significant part of paludal palynofloras, suggesting that Glossopteris had a relatively broad environmental tolerance. The distribution of other gymnosperm pollen associated with conifers, cordaitaleans, cycads and potentially peltaspems suggest that woodlands developed on better drained soils on the alluvial and coastal plains, but conifers and cordaitaleans also occur as minor components of forest swamp floras. Dulhuntyispora parvithola, previously identified as an important index taxon for late Permian Australia, appears to be more prevalent in coastal regions, which might limit its utility in more proximal parts of the Bowen and Galilee basins.

To examine the marker mudstone in the Galilee Basin, samples were taken from below, within and above this 1 m thick carbonaceous shale at the interface of the Bandanna and Rewan formations for palynological processing and stable carbon isotope . Palynofacies results show a distinct shift from the base to the top of the marker mudstone. A well-preserved translucent-phytoclast- dominated assemblage shifts to a near-barren opaque-phytoclast-dominated assemblage, indicating a potential increase in terrestrial input from the hinterland due to higher erosion rates, widespread wildfires or redox changes within the water column. Acanthomorph acritarchs and fungal spores were observed within and above the carbonaceous mudstone. Palynostratigraphy places this marker mudstone in the APP5 (Late Permian) zone with the transition to the APP6 zone occurring several centimetres above the marker mudstone. In the Bowen Basin, the marker mudstone was previously placed in the APP6 zone which suggests that the carbonaceous mudstone in the Galilee Basin is either not a direct correlative or that this unit is time-transgressive across the eastern Australian basins.

To examine the variation in the earliest Triassic floras, samples were collected from four localities around the Bowen Basin. Samples taken from the Rewan Group generally yielded low abundances or poorly preserved palynological assemblages and many samples were barren. However, yields were good enough in a sufficient number of samples to show that the shift from the APP5 zone to the APP6 and potentially APT1 zones occurs synchronously above the uppermost Permian coal seam and represents the transition from the Glossopteris-flora to a post-extinction flora. The post-extinction vegetation appears to be made up of peltasperm and conifer-dominated woodlands which sit on the more well-drained, sandier soils of the Rewan Group and a fern-dominated lowland which dominates the margins of water bodies created by base-level rise. Variable algal and assemblages as well as pyrite pockmarked palynomorphs suggest that these water bodies featured fluctuating salinity, localised anoxia and algal blooms caused by increased nutrient input due to high erosion rates causing increased deposition of organic detritus from the hinterland. Increased levels of UV-B radiation are also inferred due to the presence of mutated pollen grains and unseparated tetrads.

II

The findings of this work indicate that the diverse and widespread Glossopteris-flora was quickly and near-synchronously devastated during the EPE. The new flora consisted of hardy survivors and colonisers previously limited to extrabasinal areas. If the marker mudstone in the Galilee Basin is a correlative of the uppermost Permian coal seam, it would suggest that environmental perturbation may have already begun to occur immediately preceding the main extinction event. However, more detailed correlations based on densely spaced palynological sample sets from the uppermost Permian coal seams across the basins would be required to fully examine this.

III

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, financial support 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 higher degree by research 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 and have sought permission from co-authors for any jointly authored works included in the thesis.

IV

Publications included in this thesis

The following publication has been incorporated as Chapter 5: Wheeler, A., Van de Wetering, N., Esterle, J. S. & Götz, A. E. (2020). Palaeoenvironmental changes recorded in the palynology and palynofacies of a Late Permian marker mudstone (Galilee Basin, Australia). Palaeoworld, 29(2), 439-452. DOI: 10.1016/j.palwor.2018.10.005.

Contributor Statement of contribution %

Project conception and design 60 . Alexander Wheeler Core logging and sample collection 100

(Candidate) Analysis of palynological samples 100

Preparation of Figures 100

Writing of Manuscript 90 Nikola Van de Wetering Stable carbon isotope analysis 100 Writing of Manuscript 10 Reviewed and edited the manuscript 20 Joan Esterle (Principal Supervisor) Project conception and design 40 Reviewed and edited the manuscript 50 Annette E. Götz Reviewed and edited the manuscript 30

The following publication has been incorporated as Chapter 3: Wheeler, A., Moss, P. T., Götz, A. E., Esterle, J. S., & Mantle, D. (2021) Acid-free palynological processing: a Permian case study. Review of Palaeobotany and Palynology, 284, 104343. DOI: 10.1016/j.revpalbo.2020.104343.

Contributor Statement of contribution % Project conception and design 80 Alexander Wheeler Sample collection and acid-free processing 80 (Candidate) Analysis of palynological samples 100 Preparation of Figures 100 Writing of Manuscript 80 Patrick Moss (Associate Supervisor) Project conception and design 20 Reviewed and edited the manuscript 40 Annette E. Götz Reviewed and edited the manuscript 40 Joan Esterle (Principal Supervisor) Reviewed and edited the manuscript 20 Dan Mantle Acid sample processing 20 Writing of Manuscript 20

V

Other publications during candidature

Conference abstracts Wheeler, A., Esterle, J. & Van de Wetering, N. Palynological records of marine incursions in the Sydney-Gunnedah-Bowen Basin – preliminary investigation. 40th Sydney Basin Symposium (2017). Hunter Valley.

Wheeler, A., Esterle, J. & Van de Wetering, N. & Götz, A.E. Palynofacies of an end-Permian marker mudstone in the Galilee Basin, Australia. 50th Annual AASP Meeting (2017). Nottingham.

Wheeler, A., Van de Wetering, N., Esterle, J. & Götz, A.E. Palynology and palynofacies of the Late Permian Galilee Basin: Implications for the end-Permian palaeoenvironment. 51st Annual AASP Meeting (2018). Houston.

Wheeler, A., Van de Wetering, N., Esterle, J. & Götz, A.E. Palynology of a Late Permian marker mudstone in the Bowen and Galilee basins: Implications for inter-basinal correlation. Australian Geoscience Council Convention (AGCC) (2018). Adelaide.

Wheeler, A., Esterle, J.S. & Götz, A.E. Proximal-distal palaeoenvironmental patterns recorded by aquatic palynomophs in the Galilee Basin, Australia. 19th International Congress on the and Permian (XIX ICCP 2019). Cologne.

VI

Contributions by others to the thesis

In addition to the contributions of co-authors to published and submitted manuscripts, the following people have contributed to this thesis:

Professor Joan Esterle (principal supervisor) provided advice, edits and comments to drafts of this thesis. Professor Esterle also provided funding for research costs (sample collection, processing and analysis software) from the Vale-UQ Coal Geoscience Program.

Professor Patrick Moss (associate supervisor) provided advice, edits and comments to drafts of this thesis. Professor Moss also provided training to the candidate in order to conduct palynological sample processing at The University of Queensland.

Dr. Daniel Mantle and Dr. John Lignum provided advice, edits and comments to the palynological processing techniques and analysis presented in Chapter 3. Dr. Mantle wrote the conventional processing technique methodology. Palynological sample processing was partially conducted by MGPalaeo Pty Ltd under the supervision of Dr. Lignum. Analysis and interpretation of palynological data was conducted by the candidate.

Nikola Van de Wetering and Kim Baublys performed the stable carbon isotope analysis presented in chapter 5. Ms. Van de Wetering wrote the stable carbon isotope analysis methodology and results. Integration and interpretation of the stable carbon isotope analysis was conducted by the candidate, but Ms. Van de Wetering provided editorial comments on the discussion.

Statement of parts of the thesis submitted to qualify for the award of another degree No works submitted towards another degree have been included in this thesis.

Research involving human or animal subjects No animal or human subjects were involved in this research

VII

Acknowledgments

As my principal supervisor, Professor Joan Esterle has been a pillar of support for this project and myself over the past few years. I express my utmost thanks for her guidance through all aspects of my candidature.

Many thanks also to Professor Patrick Moss as my associate supervisor and Professor Annette Götz as an external supervisor for helping me at all stages of my research and for in-depth discussions on any questions I might have or any training I required.

I also express gratitude to colleagues who have assisted with my research, particularly Ms. Nikola van de Wetering and Ms. Kim Baublys from UQ, as well as Dr. Daniel Mantle and John Lignum from MGPalaeo. Thanks also to Bronwyn Leonard of Stanmore Coal, Martin Clarke and Max Ayliffe of Anglo-American, Luke Faulkner and Gregory Peters of Santos, Dr John McKellar and the Geological Survey of Queensland, as well as the staff of the Exploration Data Centre in Zillmere for lending their time and support in helping me collect samples, data and for the use of their facilities and equipment for this project.

I would like to thank The University of Queensland and the Vale-UQ Coal Geoscience Program for financial support over the course of my research. Thanks also to the AASP for additional funding and for welcoming me into the wider community of palynologists.

Many thanks go to my friends and colleagues in the School of Earth and Environmental Sciences, particularly my fellow PhD students for encouragement and stimulating discussion, often over a few pints. Special mention must be made to Dr. Linda Nothdurft, Dr. Gang Xia and Dr. Sandra Rodrigues for support and training in the labs and Dr. Laura Phillips, my mentor for all things regarding the Galilee Basin.

Lastly, I would like to thank my family, particularly my parents, Joe and Beryl, for their continued support throughout my life and my academic career. I cannot fully express the love and gratitude I feel for all that they have given me.

VIII

Financial support

Tuition and scholarship funding for the candidate was provided by the UQ International Scholarship (UQI).

Research funding for sample collection, processing and analysis as well as travel expenses were provided by the Vale-UQ Coal Geoscience Program, which also shared tuition funds in the UQI scheme.

Additional funding for laboratory processing expenses was provided by the AASP – The Palynological Society Student Research Award.

Keywords

Bowen Basin, Galilee Basin, , palynology, Permian-Triassic Boundary, end-Permian extinction, palaeoenvironment, palaeoecology, Early Triassic, floral turnover

IX

Australian and Standard Research Classifications (ANZSRC)

ANZSRC code: 040308 Palaeontology (incl. Palynology) (70%) ANZSRC code: 040311 (incl. and Sequence Stratigraphy) (20%) ANZSRC code: 040310 (10%)

Fields of Research (FoR) Classification

FoR code: 0403 Geology (100%)

X

Table of Contents

Abstract ...... I

Declaration by author ...... IV

Publications included in this thesis ...... V

Other publications during candidature ...... VI

Contributions by others to the thesis ...... VII

Acknowledgments ...... VIII

Financial support ...... IX

Keywords ...... IX

Australian and New Zealand Standard Research Classifications

(ANZSRC) ...... X

Fields of Research (FoR) Classification ...... X

Table of Contents ...... XI

List of Figures ...... XIX

List of Tables ...... XXVII

List of Abbreviations ...... XXVII

1. Introduction ...... 1

1.1 Overview...... 1

1.2 Tectonics and Geological Setting ...... 6

1.3 Research Objectives, Questions and Methodology ...... 6

Component 1 (Chapter 3) ...... 7

Component 2 (Chapter 4) ...... 7

Component 3 (Chapter 5) ...... 8

Component 4 (Chapter 6) ...... 8

1.4 Research Significance ...... 9

XI

1.5 Research Outcomes and Structure of Thesis ...... 10

2. Literature Review ...... 12

2.1 Introduction ...... 12

2.2 Upper Permian Stratigraphy of the Bowen and Galilee Basins ...... 12

2.2.1 Bowen Basin ...... 12

2.2.2 Galilee Basin ...... 15

2.2.3 Summary ...... 16

2.3 Palynology of the Galilee and Bowen Basins ...... 16

2.3.1 Introduction ...... 16

2.3.2 Palynostratigraphy in Australia ...... 17

2.3.3 Summary ...... 20

2.4 The Permian-Triassic Boundary ...... 23

2.4.1 Introduction ...... 23

2.4.2 The Meishan Section ...... 23

2.4.3 Dating the Boundary ...... 23

2.4.4 The end-Permian Extinction (EPE) ...... 24

2.4.5 Extinction Mechanisms ...... 25

2.4.6 Proxies for the EPE and P-T Boundary ...... 27

2.4.7 The Boundary in Australia ...... 28

2.4.8 Palynology of the Boundary in Eastern Australia ...... 28

2.4.9 Summary ...... 30

3. Acid-free palynological processing: a Permian case study ...... 32

3.1 Abstract ...... 32

3.2 Introduction and aims ...... 32

3.3 Materials and methods ...... 33

XII

3.3.1 Samples ...... 33

3.3.2 Palynological Processing Techniques ...... 34

Conventional processing (MGPalaeo) ...... 34

Acid-free processing (University of Queensland) ...... 35

3.3.3 ...... 35

3.4 Results...... 36

3.4.1 Palynofacies ...... 36

3.4.2 Palynology ...... 37

3.5 Discussion ...... 40

3.5.1 Is the acid-free process effective? ...... 40

3.5.2 How does the acid-free process compare to the conventional process? ... 40

3.5.3 How does this acid-free technique compare with other non-standard processing methods? ...... 41

3.6 Conclusions and outlook ...... 42

4. Linking lithofacies and palynology to examine proximal-distal palaeofloral patterns in the late Permian of the Bowen and Galilee basins

...... 46

4.1 Abstract ...... 46

4.2 Introduction and Aims ...... 46

4.2.1 Motivation ...... 46

4.2.2 Aims ...... 47

4.2.3 Botanical Affinities ...... 48

4.2.4 Geological Setting ...... 51

4.3 Material and Methods ...... 53

4.3.1 Core Logging, Sampling and Lithofacies Assignment ...... 53

XIII

4.3.2 Palynology ...... 53

4.4 Results...... 54

4.4.1 Lithofacies ...... 54

Mantuan Productus Beds and Peawaddy Formation ...... 54

‘Fair Hill Formation’ equivalent and Black Alley Shale ...... 55

‘Burngrove Formation’ equivalent and Bandanna Formation ...... 55

4.4.2 Palynology ...... 57

Galilee Basin – Montani 1 ...... 57

Galilee Basin – Glue Pot Creek 1 ...... 59

Galilee Basin – Tambo 1-1A ...... 61

Bowen Basin – Springsure 19 ...... 63

Bowen Basin – Taringa 7 ...... 65

4.4.3 Palaeofloral composition ...... 67

Tambo 1-1A, Springsure 19 and Taringa 7 ...... 68

Montani 1 and Glue Pot Creek 1 ...... 70

4.5 Discussion ...... 70

4.5.1 Floral Interpretations of Microflora Assemblages ...... 70

4.5.2 Palaeoenvironmental and Palaeofloral Reconstruction ...... 71

4.5.3 Algae and Acritarchs ...... 76

4.6 Conclusions ...... 77

5. Palaeoenvironmental changes recorded in the palynology and palynofacies of a late Permian marker mudstone (Galilee Basin, Australia)

...... 78

5.1 Abstract ...... 78

5.2 Introduction ...... 79

XIV

5.3 Geological Setting ...... 81

5.4 Permian–Triassic Boundary ...... 82

5.5 Materials and Methods ...... 83

5.5.1 Study Area ...... 83

5.5.2 Palynological Processing and Analysis ...... 84

5.5.3 Carbon Isotope Processing and Analysis ...... 85

5.6 Results...... 85

5.6.1 Palynostratigraphy ...... 85

5.6.2 Palynofacies ...... 88

5.6.3 Carbon Isotopes ...... 90

5.7 Discussion ...... 90

5.7.1 Age Determination – Unit APP6 ...... 90

5.7.2 Permian–Triassic Boundary ...... 91

5.7.3 Palaeoenvironment ...... 94

5.8 Conclusions and Outlook ...... 96

5.9 Acknowledgements ...... 97

6. Palynological Investigation of Variation in the Palaeoflora and Palaeoenvironment following the end-Permian Extinction (Bowen Basin,

Australia) ...... 101

6.1 Abstract ...... 101

6.2 Introduction and Aims ...... 101

6.3 Background ...... 102

6.3.1 Tectonics and Geological Setting ...... 102

6.3.2 The P-T Boundary ...... 105

6.4 Materials and Methods ...... 107

XV

6.4.1 Sampling ...... 107

6.4.2 Palynology ...... 107

6.5 Results...... 108

A41859 (Dawson Coal Mine - eastern Taroom Trough) ...... 108

CGIN0067 and CGIE0144 (Isaac Plains Coal Mine - Collinsville Shelf) ...... 110

Springsure 19 (Springsure Shelf) ...... 112

Taringa 7 (Roma Shelf) ...... 114

6.6 Discussion ...... 116

6.6.1 Preservation ...... 116

6.6.2 Biostratigraphy of the end-Permian Extinction ...... 116

6.6.3 Correlating the marker mudstone ...... 117

6.6.4 Composition and Distribution of the post-Extinction Flora ...... 118

6.6.5 P-T Boundary Palaeoenvironment in Eastern Australia ...... 120

6.7 Conclusions and Outlook ...... 124

7. Synthesis ...... 125

7.1 Synthesis of the Findings ...... 125

7.1.1 Acid-free Processing vs conventional Processing ...... 126

7.1.2 Palynology as a Tool for Palaeoecological Reconstruction in the late

Permian and Early Triassic ...... 126

7.1.3 The marker mudstone ...... 128

7.1.4 Post-EPE Palaeoflora and Palaeoenvironment ...... 129

7.1.5 Timing of the Extinction ...... 130

7.1.6 Comparison with wider ...... 131

7.2 Future Research ...... 132

7.2.1 Processing Techniques ...... 132

XVI

7.2.2 Palaeoflora and Palaeoenvironment ...... 132

7.2.3 The Permian-Triassic Boundary ...... 133

7.2.4 Biostratigraphy in Eastern Australia ...... 134

7.3 Summary ...... 134

References ...... 135

Appendices ...... 0

Appendix A - Oral and Written communication in conferences ...... 1

A1. Extended Abstract for oral presentation at the Sydney Basin Symposium

(2017) ...... 1

A2. Abstract for poster presented at 50th Annual AASP Meeting (2017) ...... 9

A3. Abstract for oral presentation at the 51st Annual AASP Meeting (2018) .. 11

A4. Abstract for poster presented at the Australian Geoscience Council

Convention (AGCC) (2017) ...... 13

A5. Abstract for poster presented at the 19th International Congress on the

Carboniferous and Permian (XIX ICCP 2019) ...... 13

Appendix B – Sample lists and raw data ...... 16

B1. Localities sampled for this study ...... 16

B2. List of samples collected for this samples ...... 16

B3. Stable carbon isotope data (δ13C org) ...... 22

B4. Palynological Data (Raw Counts) ...... 25

B4.1 Montani 1 ...... 25

B4.2 Glue Pot Creek 1 ...... 29

B4.3 Tambo 1-1A (standard processing) ...... 32

B4.4 Tambo 1-1A (acid-free processing) ...... 40

B4.5 Springsure 19 ...... 44

XVII

B4.6 Taringa 7 ...... 49

B4.7 CGIE0144 ...... 53

B4.8 CGIN0067 ...... 55

B4.9 A41859 ...... 58

B5. Palynofacies data (raw counts) ...... 61

B5.1 Tambo 1-1A (samples 1 – 22) ...... 61

B5.2 Tambo 1-1A (Samples 23 – 32) ...... 63

Appendix C1 - Step-by-step methodology for the acid-free palynological processing technique ...... 65

XVIII

List of Figures

Figure 1.1: Map of the positions of significant Permian-Triassic sedimentary basins in Australia. Basins selected for sampling in this work are highlighted in green (modified from McLoughlin, 2011b)...... 2

Figure 1.2: Recalibration of the Australian biostratigraphic scheme (Price, 1997) to the global geological timescale using U-Pb zircon dates (from Laurie et al., 2016). 4

Figure 1.3: Stratigraphy of the Cooper, Bowen and Galilee basins correlated to the

Australian biostratigraphic zones (Price, 1997) (from Phillips et al., 2017a)...... 5

Figure 1.4: Schematic block diagram showing the distribution of the major Sporomorph Ecogroups (SEG): (A) Upland SEG; (B) Lowland SEG; (C) River SEG; (D) Pioneer SEG; (E) Coastal SEG; (F) Tidally-influenced SEG adapted from

Abbink et al. (2004)...... 9

Figure 2.1: Stratigraphy of the Bowen Basin calibrated to biostratigraphy and stratrigraphic supersequences (from Sliwa et al., 2017 modified from Fielding et al., 2001; Brakel et al., 2009 and others)...... 14

Figure 2.2: Tectonostratigraphic map of the Bowen and Galilee basins showing the main structural divisions. This also includes the approximate divisions between the

Rangal, Bandanna and Baralaba Coal Measures (Sliwa et al., 2017)...... 14

Figure 2.3: Revised stratigraphy of the Galilee Basin correlated to the Bowen

Basin (modified from Phillips et al., 2017a)...... 16

Figure 2.4: Eastern Australian biostratigraphic zonation scheme including index taxa for zones and subzones (adapted rom Smith & Mantle, 2013; after Price,

1997)...... 18

Figure 2.5: Appearances of Micrhystridium evansii as recorded in well completion reports and other government reports. The distribution of M. evansii appears limited to the western Bowen Basin and southern Galilee Basin (Dickins 1964; Norvick, 1981; Price, 1984; Filatoff, 1985; Pickering, 1985a; Pickering, 1985b; Pickering,

XIX

1985c; Jones, 1986; Filatoff & Price, 1991a; Filatoff & Price, 1991b; Purcell, 2011;

Smith & Mantle, 2013; this work)...... 19

Figure 2.6: Correlation of the various palynological schemes used in eastern and western Australia (modified from Backhouse 1991). Palynological zones are calibrated to the international (Laurie et al., 2016; Ogg et al.,

2016)...... 21

Figure 2.7: Pan-Gondwana palynological correlation scheme based on the new

FAD-based scheme developed for (from Barbolini et al., 2018)...... 22

Figure 2.8: Schematic depiction of potential interrelated processes occurring at the

P-T boundary (Algeo et al., 2011; modified from Wignall, 2001)...... 26

Figure 2.9: Correlation of the P-T boundary section from the borehole Bunnerong- 1 in the Sydney Basin, Australia to the Meishan section, China and calibrated using absolute age dates to the GSSP and the phases of volcanism from the Siberian Traps. This shows a temporal gap between the terrestrial extinction interval and the

GSSP marking the end of the Permian (modified from Fielding et al., 2019)...... 30

Figure 3.1: Map and lithological log in stratigraphic context showing the position of Tambo 1-1A in the Galilee Basin of eastern Australia and the samples taken from the borehole...... 34

Figure 3.2: Classification scheme for palynofacies (from Feist-Burkhardt et al.,

2008)...... 36

Figure 3.3: Relative abundance bar chart for the palynofacies data comparing the acid-free (left) and conventional (right) processing techniques for each sample. For raw palynofacies counts see Appendix B5.2...... 37

Figure 3.4: Relative abundance chart showing a comparison of the acid-free processing (red) and the conventional processing (blue) for each observed and counted. For raw counts see Appendices B4.3 and B4.4...... 39

Figure 3.5: Photomicrograph comparison of the conventional processing technique (left) and the acid-free technique (right). A – B: TAMP23; C – D: TAMP24; E – F:

XX

TAMP25; G – H: TAMP26; Notable features include: A – degraded phytoclasts; B, D, F & G – large opaque phytoclasts; C, E & G – well preserved bisaccate pollen grains...... 43

Figure 3.6: Photomicrograph comparison of the conventional processing technique (left) and the acid-free technique (right). A – B: TAMP27; C – D: TAMP28; E – F: TAMP29; Notable features include: A – small Botryococcus; B – large Botryococcus; C – degraded phytoclasts; D, E & F – large, well-preserved palynomorphs and phytoclasts...... 44

Figure 3.7: Photomicrograph comparison of the conventional processing technique (left) and the acid-free technique (right). A – B: TAMP30; C – D: TAMP31; E – F: TAMP32; Notable features include: A – degraded phytoclasts; B – large spores, bisaccate pollen grains and opaque phytoclasts; C & E – Micrhystridium evansii; D

& F – Micrhystridium evansii and large phytoclasts...... 45

Figure 4.1: Map of the Galilee and Bowen basins showing the placement of the studied wells...... 52

Figure 4.2: Stratigraphy of the Bowen and Galilee basins and placement of the studied wells (modified from Phillips et al., 2017b)...... 53

Figure 4.3: Correlation of the lithological logs of the studied boreholes, with Montani 1 representing the most proximal section and Taringa 7 representing the most distal section. Locations of palynological samples are marked in red with the corresponding sample name. Shell symbol indicates the presence of brachiopods. BAS – Black Alley Shale; ‘BF’e – ‘Burngrove Formation’ equivalent; ‘FHF’e –

‘Fair Hill Formation’ equivalent; MPB – Mantuan Productus Beds...... 56

Figure 4.4: Palynostratigraphic data for Montani 1 showing observed genera as a relative abundance (%). Barren samples are marked by a square symbol. For raw counts see Appendix B4.1...... 58

Figure 4.5: Palynostratigraphic data for Glue Pot Creek 1 showing observed genera as a relative abundance (%). Barren samples are marked by a square symbol. For raw counts see Appendix B4.2...... 60

XXI

Figure 4.6: Palynostratigraphic data for Tambo 1-1A showing observed genera as a relative abundance (%). Barren samples are marked by a square symbol. For raw counts see Appendix B4.3...... 62

Figure 4.7: Palynostratigraphic data for Springsure 19 showing observed genera as a relative abundance (%). Barren samples are marked by a square symbol. For raw counts see Appendix B4.5...... 64

Figure 4.8: Palynostratigraphic data for Taringa 7 showing observed genera as a relative abundance (%). Barren samples are marked by a square symbol. For raw counts see Appendix B4.6...... 66

Figure 4.9 (previous page): Relative proportions of different components of the late Permian flora for Tambo 1-1A, Springsure 19 and Taringa 7, plotted against the lithological signature of each hole. These represent the potential floral compositions in a progradational lower delta setting from the Mantuan Productus

Beds to the Bandanna Formation...... 68

Figure 4.10: Relative proportionsof different components of the late Permian flora for Montani 1 and Glue Pot Creek 1, plotted against lithological signature of each hole. These represent potential floral compositions in the alluvial plain and upper delta plain respectively...... 69

Figure 4.11: Block diagram of the distribution of different components of the late Permian flora in the Bowen and Galilee basins. Glossopteris is relatively ubiquitous forming dense forests on the floodplain and cohabitating with a fern, lycopsid and sphenophyte understory in -forming mires. The more proximal alluvial and upper delta plains feature xerophytic woodlands dominated by gymnosperms and a a hydrophytic lacustrine/paludal/coal-forming flora. Horsetails form reed-like stands in fens and marshes. The coastal plain features a coastal forest in more well-drained areas, whereas areas prone to flooding and incursion feature a coastal pioneer flora. Interdistributary bays feature a mix of prasinophytes and brackish-tolerant algae, while the open water features acanthomorph acritarchs.

...... 76

XXII

Figure 5.1: Map of the Galilee and Bowen basins with location of stratigraphically significant wells. GSQ Tambo 1-1A (marked with a star) occupies a position at the edge of the Springsure Shelf, which connects the two basins...... 80

Figure 5.2: W-E cross section of the Permo-Triassic infilling the Cooper, Galilee and Bowen basins overlain by the younger sediments of the

Eromanga Basin (after Hobday, 1987)...... 81

Figure 5.3: Lithostratigraphic scheme of the Galilee Basin and Denison Trough (modified from Phillips et al., 2017b). The relative positions of the major coal seams (A-F) are also displayed. Though there is evidence of a regional unconformity in many locations, the interface of the Bandanna and Rewan formations features the marker mudstone, and no unconformity is apparent. The interval being investigated in this study is marked between the two stars...... 82

Figure 5.4 (next page): Palynostratigraphic data showing selected species from Tambo 1-1A based on counts of 200 palynomorphs. Samples in which this count could not be reached are not included. The data indicates unit APP5 (?APP5006 subunit) continues into the marker mudstone. The transition to unit APP6 occurs above the mudstone (marked by the first sparse occurrence of Protohaploxypinus microcorpus). Cluster analysis suggests the assemblages in each unit are relatively distinct from one another even though some elements of unit APP5 are still present above the marker mudstone. For raw counts see Appendix B4.3...... 86

Figure 5.5: Line chart depicting palynofacies data from Tambo 1-1A. Lines measuring relative abundance (%) also mark sample locations on the lithology. Presence of key palaeoenvironmental indicators (algae/prasinophytes, acritarchs, fungal spores) are also marked. Cluster analysis of the samples suggests three main palynofacies patterns. Ternary diagrams plotting the data based on the major palynofacies components show three distinct palynofacies assemblages (A-C). The major palynofacies components are opaque phytoclasts (OP), translucent phytoclasts (TP) and terrestrial palynomorphs (PAL). For raw palynofacies counts see Appendix B5.1...... 89

XXIII

Figure 5.6: Carbon isotope trends with depth in Tambo 1-1A showing a low magnitude excursion within the marker mudstone, plotted alongside the pollen/spore ratio, bioevents, biostratigraphy and palynofacies cluster analysis. The current placement of the P-T boundary in Australia (Laurie et al., 2016) is marked, though evidence from this study is not sufficient to define it conclusively in this locality...... 93

Figure 5.7: Reconstruction of the potential palaeoenvironmental conditions of the Bandanna Formation (A) and the marker mudstone (B). Data from Tambo 1-1A supports the interpretation of a southwards-prograding delta which subsides as base level rises leading to the formation of a large waterbody (lake) across the Bowen and Galilee basins (Phillips et al., 2017a, 2018a). Blue arrows suggest areas where a potential marine transgression could have come in. The presence of both acanthomorph acritarchs and Botryococcus suggests increased salinity but not fully open marine conditions, leading to the interpretation of a marine influenced lacustrine environment...... 96

Figure 5.8: Phytoclasts, palynomorphs and fungal hyphae, borehole Tambo 1-1A (Galilee Basin). Taxon name is followed by sample number, slide number (brackets) and stage coordinates for a ZEISS Photomicroscope III. (A) Degraded phytoclasts, TAMP8 (b), 97.1/8.2. (B) Playfordiaspora crenulata, TAMP4 (a), 91.8/0.2. (C) Protohaploxypinus microcorpus, TAMP6 (a), 114.3/17.8. (D) Fungal hyphae?, TAMP10 (b), 97.2/2.4. (E) Microreticulatisporites bitriangularis,

TAMP18 (a), 100.3/12.2...... 98

Figure 5.9: Palynomorphs of borehole Tambo 1-1A (Galilee Basin). Taxon name is followed by sample number, slide number (brackets) and stage coordinates for a ZEISS Photomicroscope III. (A) Micrhystridium sp., TAMP10 (b), 99.7/1.2. (B) Cymatiosphaera gondwanensis, TAMP17 (a), 87.2/3.3. (C) Botryococcus sp., TAMP22 (a), 80.8/13.9. (D) Botryococcus?, TAMP3 (a), 104.2/16.7. (E)

Reduviasporonites chalastus, TAMP3 (a), 109.3/17.2...... 99

XXIV

Figure 5.10: Palynofacies of borehole Tambo 1-1A (Galilee Basin). (A) Palynofacies Assemblage A featuring a high proportion of palynomorphs, particularly bisaccate pollen. (B) Palynofacies Assemblage B featuring a moderate to high proportion of translucent phytoclasts (structured and unstructured) as well as a moderate proportion of opaque phytoclasts and some well-preserved palynomorphs. (C) Palynofacies Assemblage C featuring a high proportion of opaque phytoclasts and a low proportion of translucent phytoclasts; palynomorphs are rarely well preserved...... 100

Figure 6.1: Location of studied boreholes in the Bowen Basin. Borehole Tambo 1- 1A previously studied in the Galilee Basin is also indicated (Wheeler et al., 2020).

...... 104

Figure 6.2: Broad seam terminology used in each tile of the Bowen Basin and their stratigraphic equivalents, Local mines may apply their own terminology (from

Sliwa et al., 2017)...... 105

Figure 6.3: Late Permian stratigraphy of the Bowen Basin (modified from Ayaz et al., 2015)...... 106

Figure 6.4: Palynostratigraphy of A41859 (eastern Taroom Trough). Palynological data displayed as a relative abundane (%). AC: acritarchs; AL: freshwater algae; ALPR: prasinophytes; FU: fungi; PO: pollen; SP: spores. For raw counts see

Appendix B4.9...... 109

Figure 6.5: Palynostratigraphy of A) CGIE0144 and B) CGIN0067 (Collinsville Shelf). Palynological data displayed as a relative abundane (%). AL: freshwater algae; FU: fungi; PO: pollen; SP: spores. For raw counts see Appendices B4.7 and

B4.8...... 111

Figure 6.6: Palynostratigraphy of Springsure 19 (Springsure Shelf). Palynological data displayed as a relative abundane (%). AL: freshwater algae; ALBO: chlorophycean algae; ALPR: prasinophytes; FU: fungi; PO: pollen; SP: spores. For raw counts see Appendix B4.5...... 113

XXV

Figure 6.7: Palynostratigraphy of Taringa 7 (Roma Shelf). Palynological data displayed as a relative abundane (%). AL: freshwater algae; ALBO: chlorophycean algae; ALPR: prasinophytes; FU: fungi; PO: pollen; SP: spores. For raw counts see

Appendix B4.6...... 115

Figure 6.8: (A) Late Permian “Climax” flora; (B) Post-extinction flora consisting of resistant pioneer and colonisers from the upland areas...... 119

Figure 6.9: Key palynomorphs that act as palaeoenvironmental indicators: a) Unseparated spore tetrad; b) Tetrasaccate pollen grain; c) Quadrisporites horridus; d) Rediviasporonites chalastus; e) Botryococcus; f) Cymatiosphaera gondwanensis...... 121

Figure 6.10: SEM and light microphotographs of pyritic pockmarking in organic matter from A41859: a, d) pyritic pockmarking in a bisaccate pollen grain; b, c) pyritic pockmarking in woody debris (small, light particles covering organic matter are clay ); e) pyritic pockmarking in Playfordiaspora crenulata; f) pyritic pockmarking in a structured phytoclast...... 123

XXVI

List of Tables

Table 1: Botanical affinities (Balme, 1995; Grenfell, 1995; Lindström et al., 1997; Lindström, 2005; Hochuli et al., 2010; Schneebeli-Hermann et al., 2015 Gastaldo et al., 2015; Zavattieri et al., 2017; Mays et al., 2020b)...... 50

Table 2: Organic carbon isotope samples from GSQ Tambo 1-1A and their equivalent palynological samples...... 90

List of Abbreviations

AC – Acritarchs

AL – Algae

ALBO – Algae (Botryococcus)

ALPR – Algae (Prasinophycean)

AOM – Amorphous Organic Matter

CA-IDTIMS - Chemical Abrasion-Isotope Dilution Thermal Ionisation Mass Spectrometry

DP – Degraded phytoclasts

EPE – end-Permian extinction

FU – Fungi

PO – Pollen

P-T – Permian–Triassic

PVA – Polytopic Vector Analysis

QLD – Queensland

SEG – Sporomorph Ecogroup

SP – Spores

UQ – The University of Queensland

XXVII

1. Introduction

1.1 Overview

Our understanding of late Permian and early Triassic floral communities in Australia is limited mainly to coal measures and to relatively few localities with well-preserved macrofloral remains (e.g. Retallack, 1980; Taylor et al., 1989; Diessel, 2012; McLoughlin, 1992; 1993; 1994a; 1994b; Pigg & McLoughlin, 1997; Weaver et al., 1997; Anderson et al., 1999; Holdgate et al., 2005; Prevec et al., 2010; McLoughlin, 2011a; Van de Wetering et al., 2013b). While attempts have been made to use palynology as a proxy for macrofloral assemblages (Beeston, 1987; Van de Wetering et al., 2013b; Wagner et al., 2019) the botanical affinities of many Permian palynomorphs are unknown or limited to broad groups. This has significant implications for palaeoenvironmental reconstruction and studies of the end-Permian mass extinction (EPE).

Confirming correlations of the Permian-Triassic strata of eastern Australia has been consistently hampered due to the poor exposure of outcrop particularly in the Bowen and Gunnedah basins, part of the broad eastern Australian basins system that includes the Sydney Basin (Totterdell et al., 2009) (Fig. 1.1). Stratigraphic core drilling tends to be clustered in certain areas, usually around active mining zones, while there is relatively sparse drilling in other areas, limiting the available core from different parts of these basins. A biostratigraphic scheme developed by Price (1997) based on palynology has generally formed the basis of these correlations. While useful for broad intra- and inter-basinal correlation within Australia, this scheme defines its palynozones on species largely endemic to Australia, frustrating intra-Gondwanan correlation and attempts to calibrate the scheme to the global geological timescale (Nicoll et al., 2015; Laurie et al., 2016; Smith et al., 2018). Yet more challenges are experienced at the Permian-Triassic boundary and Early Triassic strata, where there are rapid changes in vegetation, but our understanding of the palaeoenvironmental processes that might affect our interpretation is limited. Factors like weathering, reworking and marine incursions can transport palynomorphs to different localities, control the types of sediments deposited and can influence preservation of , leading to a bias in our interpretation of palaeoenvironments.

1

Figure 1.1: Map of the positions of significant Permian-Triassic sedimentary basins in Australia. Basins selected for sampling in this work are highlighted in green (modified from McLoughlin, 2011b).

Recent work has attempted to alleviate correlation issues in Australia by applying high-resolution U- Pb zircon dates via CA-IDTIMS (chemical abrasion-isotope dilution thermal ionisation mass spectrometry) to calibrate the eastern Australian biostratigraphic scheme (Fig. 1.2) (Metcalfe et al., 2011; Smith & Mantle, 2013; Ayaz et al., 2015; Nicoll et al., 2015; Laurie et al., 2016; Phillips et al., 2018a) to the update Geological Time Scale 2012 (Gradstein et al., 2012). While this has proved a highly successful endeavour, correlation challenges are still faced at a formation-scale due to the relatively low resolution of the biostratigraphic scheme of Price (1997). For example, the APP5 zone(equivalent to the Dulhuntyispora parvithola Zone of Backhouse, 1991) covers a period of up to 5.7 million years (Laurie et al., 2016) and spreads across five formations in the Denison Trough of the Bowen Basin (Fig. 1.3) (Phillips et al., 2017a). Price (1997) did erect subzones but the relatively sporadic occurrence of the index taxa which define these subzones (e.g. Lycopodiumsporites “crassus”) has limited their utility.

For higher resolution correlations, improved palaeoenvironmental reconstructions, and a better understanding of the Permian-Triassic boundary, the distribution and environmental affinities of the

2 palynological assemblages needs to be examined. By attempting to link palynological assemblages to a sedimentological context, we can start to better define the spatial and temporal limits of key taxa, which in turn is controlled by their environmental preferences. This project will target the late Permian to Early Triassic interval in the Bowen and Galilee basins for sampling. This will capture a sequence from the last Permian marine incursion up to the Permian-Triassic boundary interval. Sampling localities will be selected to capture a variety of facies and depositional environments and will allow for examination of palaeofloras affected by differing palaeoenvironmental conditions. This work also follows up on recent work in the Galilee and Bowen basins, which has attempted large- scale correlations, biostratigraphy, chronostratigraphy and basin analysis (Ayaz et al., 2015; Ayaz et al., 2016a; Phillips et al., 2017b; Phillips et al., 2018a; Sliwa et al., 2017).

3

Figure 1.2: Recalibration of the Australian biostratigraphic scheme (Price, 1997) to the global geological timescale using U-Pb zircon dates (from Laurie et al., 2016).

4

Figure 1.3: Stratigraphy of the Cooper, Bowen and Galilee basins correlated to the Australian biostratigraphic zones (Price, 1997) (from Phillips et al., 2017a).

5

1.2 Tectonics and Geological Setting

The Bowen and Galilee basins cover a combined area of 317 000 km2 and are located in Queensland, Australia (Hulleatt 1991; Allen & Fielding 2007). The Bowen Basin forms part of an inferred foreland basin system alongside the Gunnedah and Sydney basins (Korsch & Totterdell 2009). It is bounded in the east by the New England Orogen and in the west by the Anakie inlier and overlain by the Surat Basin in the south (Draper, 2013). The basin developed from the early Permian to the mid-Triassic in three major phases: extension during the early Permian, thermal subsidence in the mid-Permian and foreland loading during the Late Permian to the Middle Triassic coinciding with the Hunter-Bowen Orogeny (Fielding et al., 2000; Fielding et al., 2001; Sliwa et al., 2017). It hosts significant coal and coal seam gas resources and produces both thermal and coking coal.

The late Carboniferous to Middle Triassic Galilee Basin is interpreted as an intracratonic basin that formed to the west of the Bowen Basin and east of the Cooper Basin. The basin is underlain by the Thomson Orogen in the centre, the -Carboniferous Drummond Basin in the east, Proterozoic cratonic rocks in the west and the Devonian Adavale Basin in the south (Van Heeswijck, 2010). The basin is overlain by the - Eromanga and Surat basins as well as Cenozoic cover significantly limiting surface outcrops of the basins rocks (Van Heeswijck, 2010). The tectonic of the basin is influenced by far-field tectonics occurring to the east and, thus, basin infill occurred in two phases of thermal subsidence (late Carboniferous to early Permian and late Permian to Middle Triassic), separated by a depositional hiatus in the middle Permian (Korsch & Totterdell, 2009; Van Heeswijck, 2010; Phillips et al., 2018a). The Galilee Basin also hosts valuable coal reserves and is currently being argued as a potentially significant coal resources for Australia in the near future.

1.3 Research Objectives, Questions and Methodology

The objective of this thesis is to examine variation and distribution of palynological assemblages in different depositional environments and examine the links between the composition of in-situ floras and environmental controls influenced by these different environments. This will be achieved by first examining proximal to distal vegetation changes in a relatively stable “climax” flora i.e. the late Permian Glossopteris flora and then examining spatial variation in a disturbed and stressed post- extinction flora. This thesis will also examine the palynology of an unusual marker mudstone, which sits above the extinction interval as it has been identified in the Bowen and Galilee basins and suggest how it may relate to the Early Triassic environment and flora. As a significant proportion of the data gathered in this thesis are from palynological samples taken from different basins and formations, an

6 additional goal of this thesis was to use acid-free processing techniques on at least a subset of the samples. This is to determine if there is significant variability between the different processing techniques as well as to test if the acid-free processing might produce better yields and preservations of palynomorphs that can show distinct environmental signatures. Standard acid (hydrofluoric and hydrochloric acid) processing of palynological samples is expensive and hazardous, and thus further developing an acid-free technique for Palaeozoic samples has the potential to minimise risks and costs to increase the number of samples able to be processed and examined. The main objectives of the thesis have been split into four smaller components.

Component 1 (Chapter 3)

Component 1 evaluates the acid-free technique of palynological processing and examines samples processed with both acid and acid-free processing techniques. This component answers the following research questions:

• Is an acid-free processing technique viable for processing Permian-aged palynological samples representing a fluvio-deltaic and shallow marine environment? • Are the results qualitatively and quantitatively comparable to samples processed with standard acid techniques? • Does the acid-free processing produce better yields and preservations of palynomorphs that are environmentally significant?

Component 2 (Chapter 4)

Component 2 examines proximal-distal variation in the microfloral assemblages from the last Permian marine incursion to the Permian-Triassic boundary as a proxy for spatio-temporal changes in plant communities in different depositional environments. This component seeks to answer the following questions:

• Do floral communities change from proximal to distal depositional settings in a way that can be observed in the palynological record? • Do floral communities change at individual localities as we observe up-section facies changes related to changing depositional environments (i.e. delta progradation)? • How is the last Permian marine incursion expressed in the palynological record?

7

Component 3 (Chapter 5)

Component 3 examines a carbonaceous mudstone in the southern Galilee Basin that, previously, was tentatively correlated with the so-called marker mudstone in the Bowen Basin. High-resolution biostratigraphy, palynofacies and stable carbon isotopes have been used to determine if this correlation is correct and to explore the implications for the end-Permian palaeoenvironment. This component seeks to answer the following questions:

• Is the marker mudstone Permian or Triassic in age based on biostratigraphy? • How does the position of the marker mudstone in the Galilee Basin compare to that of the marker mudstone of the Bowen Basin? • What do palaeoenvironmental indicators suggest about the palaeoenvironment at the time of the “marker mudstone’s” deposition?

Component 4 (Chapter 6)

Based on the results from component 3, component 4 examines the palynology of the Permian- Triassic boundary and earliest Triassic at several different localities around the Bowen Basin, to determine if there is a consistent environment around the basin. The aims are to reconstruct the palaeoenvironment based on flora and other palynological environmental indicators. This component attempts to answer the following questions:

• Is the floral turnover marked by the change from the APP5 to the APP6 zone synchronous throughout the Bowen Basin? • Is the vegetation consistent in different parts of the Bowen Basin immediately following the end-Permian extinction? • Does the palynology give any indication of the post-extinction palaeoenvironment?

The conceptual framework of this thesis is broadly similar to that of the Sporomorph Ecogroup Model (Fig. 1.4) (Abbink, 1998). The Sporomorph Ecogroup Model attempts to define Mesozoic palaeocommunities based on quantitative data from palynomorphs whose parent plants have known ecological preferences (Abbink et al., 2001; Abbink et al., 2004; Ruckwied et al., 2008; Kustatscher et al., 2010; Gedl et al., 2012). These ecological preferences are based on macrofloral data and extrapolation from modern plant communities into the past where applicable. Palaeoenvironmental conditions as they relate to stress and disturbance are used to define the different Ecogroups to which

8 the palynomorphs are assigned (Abbink et al., 2004; Grime, 1979). This work takes into account depositional environment as a primary driver for the distribution of floral elements as well as algae, acritarchs and fungi, which are useful environmental indicators to produce a broad palaeoenvironmental and palaeofloral model for the Late Permian and earliest Triassic.

Figure 1.4: Schematic block diagram showing the distribution of the major Sporomorph Ecogroups (SEG): (A) Upland SEG; (B) Lowland SEG; (C) River SEG; (D) Pioneer SEG; (E) Coastal SEG; (F) Tidally-influenced SEG adapted from Abbink et al. (2004).

1.4 Research Significance

The floral ecosystems of the late Permian and Early Triassic of Australia are poorly understood due to macrofloral remains being limited to coal measures (which are absent in the Early Triassic) and a number of fossil localities. In many cases, the coal seams themselves are of high rank and low liptinite content dissuading detailed palynological investigations. Large portions of these basins are not exposed to the surface, limiting the number of outcrops that can be examined. This thesis aims to improve our understanding of these floral communities by linking palynological and sedimentological datasets. This will improve our palaeoenvironmental models particularly in the resource-rich upper Permian sequences in the Galilee and Bowen basins. Currently, the late Permian coal deposits in the Bowen and Galilee basins fall into a single biostratigraphic zone (APP5). A better understanding of spatio-temporal changes in floral composition can assist in large-scale correlations of coal seams and marker beds, particularly with the assistance of zircon ages from ashfall tuff beds. It will assist future works in determining the spatial controls affecting facies-bound index taxa and the viability of biostratigraphic subzones. When combined with robust facies maps high-resolution palynological datasets can assist in determining the position of the palaeocoastline, which in turn will allow for more accurate mapping of potential energy resources. It will also improve our understanding of the post-extinction flora and palaeoenvironments immediately following the end-Permian mass extinction. Partly this will be through understanding the local composition of the post-extinction flora

9 and how that changes in different localities, and partly it will be done through examining palaeoenvironmental indicators, which can give some clues as to the palaeoenvironmental conditions and stresses affecting the ecosystem.

1.5 Research Outcomes and Structure of Thesis

This thesis is structured as a thesis by publications and as such chapters 3-6 have been or are currently being prepared for review or have been submitted and/or accepted for publication. Each chapter is based on the components outlined in section 1.3 and some repetition may be encountered in the background and figures in each chapter. In addition, some data may be presented in multiple chapters. A literature review is provided in chapter 2 to provide a broader background to the subsequent chapters and to outline the gaps in the literature that this thesis will attempt to address. Where relevant, reference will be made to the initial publication in which the dataset was published. The referencing style used is APA and citations are listed in chronological order.

The principle aims and outcomes of each research chapter are:

Chapter 3: To develop the utility of an acid-free palynological processing technique and compare it to the standard palynological processing (HF and HCl) used in many university and commercial laboratories.

This chapter aims to establish that an acid-free processing technique is effective on suitable Permian material and may be utilised on sets of samples presented in later chapters. Samples were collected from a single locality and splits were sent to a commercial lab for conventional processing as well as processed using the acid-free technique at The University of Queensland. Palynofacies and biostratigraphic data were collected and compared between each set of samples to identify similarities or differences in the assemblages yielded.

Chapter 4: To examine spatial and temporal variation in the palynological assemblages of the Late Permian in the context of a switch from marine to fully terrestrial conditions.

This chapter examines palaeoecological changes from the last Permian marine incursion to the Permian-Triassic Boundary in boreholes selected to represent a proximal to distal facies change. Varying compositions of sporomorph assemblages were then classified by their palaeobotanical affinites and examined in relation to the depositional environment to provide information as to the palaeoenvironmental affinities of these endmembers.

10

Chapter 5: To evaluate the marker mudstone in the Galilee Basins as a correlative of the one observed and mapped in the Bowen Basin as well as its palaeoenvironmental implications for the Permian-Triassic Boundary.

This chapter examines an organic-rich carbonaceous mudstone in the Galilee Basin that occupies the interface between the uppermost Permian Bandanna Formation and lowermost Triassic Rewan Formation. Samples were collected from the mudstone horizon itself as well as from the formations above and below. The mudstone was analysed using palynology, palynofacies and stable carbon isotope geochemistry. A biostratigraphic assessment and carbon isotope analysis was done to determine the position of the mudstone relative to the Permian-Triassic Boundary. A palaeoenvironmental model was developed using palynofacies changes and the presence of palaeoenvironmental indicators (algae, acritarchs, fungi).

Chapter 6: Examines variation in the immediate post-extinction flora of the Bowen Basin and use other palaeoenvironmental indicators (algae, acritarchs, fungi) to build a broad palaeoenvironmental model of the earliest Triassic in the Bowen Basin.

This chapter looks at the palaeoflora and palaeoenvironment of the earliest Triassic strata to examine the biostratigraphy of the boundary, the composition of the floras in different localities and the palaeoenvironment that these floras occupied. It also follows up on the results of the previous chapter to confirm or reject the correlation of the marker mudstone. Data are presented from samples collected from the uppermost Permian and lowermost Triassic interval in different parts of the Bowen Basin. A biostratigraphic assessment was made to determine the position of the Permian-Triassic Boundary and an ecological interpretation is presented based on the palynological assemblage. Palaeoenvironmental indicators were identified to determine palaeoenvironmental conditions and if these are local or regional scale signatures.

11

2. Literature Review

2.1 Introduction

While later chapters will examine the previous literature specific to that study, the aim of this chapter is to cover broader background topics and establish the knowledge gap that this thesis will attempt, in part, to fill. This involves an examination of: 1) the newly updated stratigraphic correlation scheme for the upper Permian in the Bowen and Galilee basins and the distribution and composition of the coal deposits therein; 2) historical work done on late Permian palynology and biostratigraphy in eastern Australia; and 3) the nature and definition of the Permian-Triassic Boundary at the Global Stratotype Section (GSSP), how it has been defined in eastern Australia and the complictions that arise from this.

2.2 Upper Permian Stratigraphy of the Bowen and Galilee Basins

2.2.1 Bowen Basin

The late Permian phase of sedimentation in the Bowen Basin is marked tectonically by the onset of foreland loading (Korsch et al., 2009). This period is marked by the development of widespread coal deposits and a series of marine transgressions and regressions. The general sedimentary style is clastic sedimentary material occasionally interbedded with volcanic ashfall tuffs. These tuffs have allowed for high-resolution age dating, which allows for intrabasinal correlation and calibration of the biostratigraphic scheme (Laurie et al., 2016). This thesis will focus on an interval from the base of the Black Alley Shale, and incorporating the Mantuan Productus Beds, up to the basal Rewan Group (Fig. 2.1). The interface of the Peawaddy Formation and Black Alley Shale is regarded as the last significant Permian marine transgression in the Bowen and Galilee basins (Fielding et al., 2000). This interval is sometimes occupied by the Mantuan Productus Beds, which is made up of brachiopod shells and coincides with the Micrhystridium evansii acritarch acme event (Price, 1997; Smith & Mantle, 2013). The restricted marine facies of the Black Alley Shale are progressively overlain by the Fort Cooper Coal Measures and Burngrove Formation, a prograding deltaic, coal-rich sequence (Fielding et al., 2000; Ayaz et al., 2016b). These coal measures also contain interbedded ashfall tuffs, marking significantly increased volcanic activity in the arc. The top of this sequence is marked by the Yarrabee Tuff (Fig. 2.1), a prominent marker horizon that has been dated and correlated across the Bowen Basin (Ayaz et al., 2016a). This work dated the Yarrabee Tuff at 252.69±0.16 Ma in the northern Taroom Trough and 253.07±0.22 Ma in the south-east Taroom Trough correlating strongly

12 with ages previously published for the Kaloola Tuff Member in the Roma Shelf (Metcalfe et al., 2015).

The final sequence of Permian deposition is made up of the Rangal Coal Measures and stratigraphic equivalents. At this point the depositional environment had switched to an expansive fluvial- dominated coastal plain, which drained southwards as a response to overfilling due to the increased supply (Fielding et al., 1993; Falkner & Fielding, 1993). The coal measures of this interval are loosely split into three geographic tiles: the Bandanna tile in the west, the Rangal tile in the north and the Baralaba tile in the east; named after the coal measures in each area (Fig. 2.2). Sliwa et al. (2017) defined three coal domains in the Rangal-Bandanna-Baralaba Coal Measures: Domain A in the main depocentre on the eastern side of the basin, which contains the thickest sedimentary sequence, highest net coal and the most seams; Domain B in the centre of the basin, which contains less coal but a large number of thinner seams; and Domain C in the west and north of the basin, which is made up of a relatively thin sedimentary sequence and contains less coal, and fewer, thinner seams. These domains also show the depocentre, which borders a significant thrust system to the west, likely extended northwards and has been uplifted and eroded. In certain areas, including within Domain C, the seams can coalesce into relatively thick (>25 m) pods of coal termed “crabs” or “squids” based on the way the seams split away from these pods (Sliwa et al., 2017). These pods represent near undisturbed peat formation for tens or hundreds of thousands of years while the splitting is controlled by subsidence and base-level changes marginal to these areas (Diessel 1992; Fielding et al., 1984). The splitting is greatest towards the south and east where accommodation and subsidence were highest, though there is also some increased splitting towards the north of the Nebo Synclinorium (Sliwa et al., 2017).

The Rewan Group overlies the uppermost Permian coal seams in the Bowen and Galilee basins. The Rewan Formation was upgraded to group status (Jensen, 1975), but the former nomenclature is still used often, particularly in the Galilee Basin where it has not been subdivided (Exon, 1970; Jell 2013). In the western part of the Bowen Basin, the basal part of the Rewan Group features a distinct carbonaceous mudstone termed the marker mudstone, which although not a continuous feature, is laterally widespread and variable in thickness (Michaelsen et al., 2000; Michaelsen 2002; Sliwa et al., 2017). Above this feature lies the Sagittarius Sandstone, which mainly consists of fining-up cycles of interbedded red and green mudstones and siltstones. The uppermost part of the Sagittarius Sandstone switches to a broadly coarsening up sequence into conglomeratic sandstones termed the Brumby Sandstone Member (Grech, 2001; Brakel et al., 2009; Sliwa et al., 2017). This sequence is devoid of coal and is thought to represent a fluvio-lacustrine environment in the Early Triassic (Grech, 2001).

13

Figure 2.1: Stratigraphy of the Bowen Basin calibrated to biostratigraphy and stratrigraphic supersequences (from Sliwa et al., 2017 modified from Fielding et al., 2001; Brakel et al., 2009 and others).

Figure 2.2: Tectonostratigraphic map of the Bowen and Galilee basins showing the main structural divisions. This also includes the approximate divisions between the Rangal, Bandanna and Baralaba Coal Measures (Sliwa et al., 2017).

14

2.2.2 Galilee Basin

In some respects the southern part of the Galilee Basin appears to have more in common with the western Bowen Basin than it does with the northern Galilee Basin. The late Permian deposits of the southern part of the Galilee Basin share nomenclature and often lithologies with the Springsure Shelf and Denison Trough of the Bowen Basin (Fig. 2.3). The presence of the Peawaddy Formation, Black Alley Shale and Mantuan Productus Beds also mark the extent of marine incursion into the Galilee Basin from the east along with the Micrhystridium evansii acme event, though these formations appear to thin significantly westwards due to decreasing accommodation (Norvick, 1981; Allen & Fielding, 2007). The northern Galilee Basin is separated from the south by a topographic high named the Barcaldine Ridge (Van Heeswijck, 2010). The upper Permian interval of this part of the basin is generally marked by a combination of fluvial deposits and significant coal deposits known as the Betts Creek Beds (Vine et al., 1964). Phillips et al., (2017a) recently reviewed and revised the stratigraphy and coal seam nomenclature of the Galilee Basin for more consistent correlation with the Bowen Basin (Fig. 2.3).

Deposition in the Galilee Basin occurred at much lower subsidence rates compared to the Bowen Basin, with the lower accommodation attributed to thermal subsidence rather than the foreland loading being experienced in the Bowen Basin at the same time (Allen & Fielding 2007). However, Van Heeswijck (2010) and Phillips et al. (2017b) suggest increased subsidence, particularly in the east and south of the basin, is a far-field response to the foreland loading, based on seismic data and increased coal seam splitting towards the south and east. The thermal-coal deposits in the Galilee Basin tend to contain higher inertinite and matter than the in the Bowen Basin and are of lower rank (Huleatt, 1991; Mutton, 2003). The splitting patterns and coarsening-upward sequences in the Bandanna Formation suggest that the southern Galilee coals were deposited in a lower deltaic setting while further north, thicker coals were deposited in an upper deltaic plain and alluvial setting (Phillips et al., 2017b).

15

Figure 2.3: Revised stratigraphy of the Galilee Basin correlated to the Bowen Basin (modified from Phillips et al., 2017a).

2.2.3 Summary

The palaeoenvironment of the Bowen and Galilee basins during the late Permian and Early Triassic was dynamic, controlled by the tectonic activity of the arc to the east. This provides an opportunity to study the reactions of the Gondwanan palaeoflora to the changing environment both spatially and temporally.

2.3 Palynology of the Galilee and Bowen Basins

2.3.1 Introduction

This section examines the previous work conducted to establish a palynostratigraphic scheme that could be applied in Australia. Much of the palynological data collected in the Galilee and Bowen basins is stored in industry well completion reports, government reports and other “grey literature”, which has not undergone formal peer review. It cannot, however, be overlooked, as it represents a significant proportion of research conducted in the basins and makes up the biostratigraphic framework applied in eastern Australia.

16

2.3.2 Palynostratigraphy in Australia

Some of the earliest palynological studies in the Bowen Basin were conducted by de Jersey (1946). This study broadly described the miospores taken from several coal measure localities in the basin and made a tentative correlation to the Greta Coal Measures in New South Wales (Reid, 1929; 1930). In parallel, Permian microspores were described in the Sydney and Tasmania basins (Dulhunty, 1945; 1946; Dulhunty & Dulhunty, 1949). A fully developed biostratigraphic scheme used in the Permian- aged basins of Queensland was initially developed by Evans (1969) as a series of “stages” based on palynological assemblages of certain key taxa. This was further refined by Paten (1969) who used index taxa to define subdivisions. Further development has sought to use the first appearance of stratigraphically important index taxa to further refine these zones (Price, 1976; Price, 1983; Price et al., 1985) culminating in the alphanumeric zones and subzones developed by Price (1997) (Fig. 2.4). While the major zones are easily recognisable and well-defined in different parts of Australia, the utility of some of the subzones has been questioned. The Micrhystridium evansii acme event has been identified as being limited to the Springsure Shelf, Denison Trough, Comet Ridge and the south-east of the Galilee Basin (Fig. 2.5). This has limited its wider utility but it is a useful stratigraphic marker when it can be observed. Schemes developed in the Sydney Basin have also been applied to eastern Australia at large. Helby (1970) defined a number of assemblage zones covering the Permian to Triassic transition in the Sydney Basin.

17

Figure 2.4: Eastern Australian biostratigraphic zonation scheme including index taxa for zones and subzones (adapted rom Smith & Mantle, 2013; after Price, 1997).

18

Figure 2.5: Appearances of Micrhystridium evansii as recorded in well completion reports and other government reports. The distribution of M. evansii appears limited to the western Bowen Basin and southern Galilee Basin (Dickins 1964; Norvick, 1981; Price, 1984; Filatoff, 1985; Pickering, 1985a; Pickering, 1985b; Pickering, 1985c; Jones, 1986; Filatoff & Price, 1991a; Filatoff & Price, 1991b; Purcell, 2011; Smith & Mantle, 2013; this work).

A key group used in defining the late Permian palynological zones of eastern and western Australia are those of the genus Dulhuntyispora. This group is nearly endemic to eastern Gondwana with only a single recorded occurrence in South Africa (Anderson, 1977). Other occurrences have been noted in the Cribas Formation in Timor Leste (McCartain et al., 2006), as well in in the Prince Charles Mountains (Lindström & McLoughlin, 2007) and Rennick Glacier area in the Transantarctic Mountains (Bomfleur et al., 2020). Reworked specimens of Dulhuntyispora have also been observed in Tertiary sediments in north-eastern India (Venkatachala & Kar, 1990). These unusual spores are well constrained in terms of their temporal range and are visually distinct featuring interradial ‘blisters’ of various sizes in each species. Price & Filatoff (1990) examined potential lineage relationships that would represent their potential evolutionary line from the more common Permian taxon Microbaculispora.

The Galilee Basin has been the subject of relatively few palynological studies compared to the Bowen and Sydney basins. Peer-reviewed palynological studies from the Galilee Basin, prior to this work,

19 were limited to Jones & Truswell (1992) who described the palynostratigraphy of the Carboniferous- Permian boundary and Phillips et al. (2017b) who used palynology and U-Pb zircon ages from tuffs to improve correlations of the late Permian strata in the basin. In the Galilee Basin, most palynological studies are part of stratigraphic drilling reports (Norvick, 1971; 1974; McKellar, 1975; 1977a; 1977b; 1978; 1979; De Jersey & McKellar, 1979; 1981; Brain et al., 1991; Green et al., 1991; McKellar et al., 1999;). A report published for the Bureau of Mineral Resource, Geology and Geophysics by Norvick (1981) represents the most comprehensive examination of the palynostratigraphy of the Galilee Basin. A PhD thesis by Millsteed (1997) also attempted to refine and correlate the Galilee Basin palynostratigraphy by defining a series of Oppel zones.

In Western Australia, Segroves (1970) defined five palynological zones in the marine and terrestrial sediments of the Perth Basin. Kemp et al. (1977) refined the eastern Australian scheme of Evans (1969) and applied it to the Canning Basin. Backhouse (1990; 1991; 1993) refined the palynostratigraphy of the Collie Basin and correlated it to the Officer and Perth basins in south- western Australia and later Mory & Backhouse (1997) adapted it for the Canning Basin (Fig. 2.6). Due to the position of these basins relative to the Tethys margin and other Gondwanan basins, this scheme has been found to be useful in developing large scale correlations across Gondwana, though it was observed that many of the stratigraphically significant index taxa are diachronous over such a large distance (Barbolini et al. 2016; 2018) (Fig. 2.7).

2.3.3 Summary

The palynological scheme of Price (1997) that is applied in the Bowen Basin is an informal scheme that was never formally published. The widespread utility of the zones (and especially the subzones) is not fully understood. In some cases, species that are used as index taxa to define subzones, such as Lycopodiumsporites “crassus”, the index taxon for the APP5006 subzone, have not been formally defined with taxonomic descriptions and photoplates, an issue that was previously discussed in Foster & Archbold (2001). While a formal codification of this scheme is beyond the scope of this thesis, a detailed examination of the ecological aspects of late Permian and earliest Triassic sporomorph assemblages can enhance the utility and applicability of the scheme by helping to identify taxa that are potentially facies-restricted or geographically limited.

20

Figure 2.6: Correlation of the various palynological schemes used in eastern and western Australia (modified from Backhouse 1991). Palynological zones are calibrated to the international geologic time scale (Laurie et al., 2016; Ogg et al., 2016).

21

Figure 2.7: Pan-Gondwana palynological correlation scheme based on the new FAD-based scheme developed for South Africa (from Barbolini et al., 2018).

22

2.4 The Permian-Triassic Boundary

2.4.1 Introduction

The Permian-Triassic Boundary marks the greatest known extinction of flora and fauna in the fossil record (Erwin, 1993; 1994). The end-Permian extinction (EPE) affected nearly all groups of living organisms in both marine and terrestrial habitats. In Gondwana, it marks the end of the widely- distributed Glossopteris-flora and precedes the Dicroidium-flora of the Triassic. The period after the boundary is marked by a global “coal gap” (Retallack et al., 1996), which is thought to represent harsh climatic and environmental conditions unsuitable for peat-forming plant communities and the preservation of peat.

2.4.2 The Meishan Section

The Permian-Triassic Boundary is formally defined by a GSSP (Global Stratotype Section and Point) in the Meishan section located in the Changxing County in southern China (Yin et al., 2001). The section was deposited in the north-eastern part of the North marginal basin of the Yangtze Platform (NMBY), which makes up a northward slope of the Yangtze Carbonate Platform and represents the deep water marine deposits (Yin et al., 2014). The boundary is formally recognised by the FAD (first appearance datum) of the conodont Hindeodus parvus in bed 27c within the Meishan section (Yin et al., 1996). The main phase of the extinction itself is recorded in beds 25 and 26 but there is still debate on whether the extinction occurred in a single phase (Jin et al., 2000) or in multiple phases with smaller extinctions occurring below and above the main phase (Yang et al., 1993; Yin et al., 2007).

Multiple episodes of environmental perturbation related to fluctuating atmospheric CO2 (Retallack et al., 2011) and marine anoxia (Zhang et al., 2018) have been suggested to have inhibited the recovery of global biotic systems. However, Rampino et al. (2000), Twitchett et al. (2001) and Angiolini et al. (2010) present evidence that the extinction was a single, rapid event.

2.4.3 Dating the Boundary

Age dates from the Meishan section were constrained even before the formal definition of the GSSP. U/Pb SHRIMP zircon dates from a bentonite below the H. parvus zone gave an age of 251.2 ± 3.4 Ma (Claoue-Long et al., 1991). Sanidine Ar/Ar age dates from the base of the Chinglung Formation were averaged with those from an equivalent P-T section in Shangsi, China, to give an age of 249.98 ± 0.95 Ma (Renne et al., 1995). More recent radiometric age dates for bed 25 are 252.3 ± 0.3 Ma

23

(Bowring et al., 1998) and 252.4 ± 0.3 (Mundil et al., 2004). More recently, Burgess et al. (2014) used a combination of U/Pb zircon age dates and calculated sediment accumulation rates to gather high resolution ages and durations of beds and events across the boundary section. To this end, the FAD of H. parvus was placed at 251.902 ± 0.024 Ma with the extinction interval occurring from bed 25 (251.941 ± 0.037 Ma) to bed 28 (251.880 ± 0.031 Ma) over a period of 0.061 ± 0.048 Ma. Baresal et al. (2017) published U-Pb ages from the Dongpan and Penglaitan sections in south China correlating them to the Meishan section.

2.4.4 The end-Permian Extinction (EPE)

The result of the significant climatic and environmental disturbance that occurred at the end of the Permian is reputed to have been the greatest mass extinction event in Earth’s history. Estimates of extinction rates based on data compilations have suggested that up to 54% of marine invertebrate families and up to 96% of marine species went extinct during the end-Permian extinction (Raup, 1979; Erwin, 1990), though extinction rates among different faunal and floral groups, as well as different kingdoms of life, are still being debated. Extinction rates among terrestrial tetrapods have been placed at between 74% and 82% (Maxwell, 1992; Benton et al., 2004). The extinction itself appears to have occurred relatively rapidly, with estimates for the extinction period ranging over 40000 to 500000 years (Bowring et al., 1998; Twitchett et al., 2001), but full recovery of global faunas only occurred by the Middle Triassic pointing to a significantly delayed recovery (Looy et al., 1999; De Wit et al., 2002; Irmis & Whiteside, 2012). The delay may have been due to the sheer magnitude of the extinction (Erwin, 1998) but also may be due to prolonged climatic and environmental instability (Retallack et al., 2011).

Most marine groups were significantly affected by environmental perturbation at the end of the Permian, and many went almost or completely extinct, including fusulinid , certain groups of corals and echinoderms, trilobites, articulate brachiopods, crinoids and stenolaemate bryozoans (Hallam & Wignall, 1997; Erwin et al., 2002; Benton & Twitchett, 2003). Terrestrial faunas were also significantly affected. The extinction that therapsids faced is particularly well documented in the Karoo Basin, South Africa (Smith & Ward, 2001; Retallack et al., 2003; Gastaldo et al., 2015). Plant communities also experienced significant turnover and disruption across the EPE. Nowak et al. (2019) suggests that this turnover does not constitute a true mass extinction event, based on relatively low decreases in taxonomic diversity. However, in Gondwana, the complete extinction of Glossopterids, which made up the most significant portion of plant biomass in the southern hemisphere, had severe and long-lasting effects on terrestrial ecosystems as well as the hydrological

24 and carbon cycles (McLoughlin, 2011a; Vajda et al., 2020). While the Gondwanan climate generally ameliorated during the Permian following the Late Palaeozoic Ice Age, the late Permian flora in Australia was relatively stable as indicated by the single biostratigraphic zone (APP5) that encompasses most of the late Permian sedimentary sequence in the Bowen Basin (Fig. 2.1) (Cuneo, 1996).

2.4.5 Extinction Mechanisms

The current widely accepted cause for the end-Permian extinction is thought to be the Siberian Traps, a large igneous province, which erupted contemporaneously with the extinction (Renne et al., 1995;

Reichow et al., 2009; Saunders & Reichow, 2009). In addition to degassing of CO2 from the eruption, it is also suggested that the Permian-aged coals and carbonaceous shales of the Tunguska Basin were thermally altered by contact metamorphism, significantly increasing CO2 and CH4 input into the atmosphere (Campbell et al., 1992; Svensen et al., 2009). There are, however, thought to be other extinction mechanisms that may be connected to the volcanism (Fig. 2.8). This includes H2S degassing and methane clathrate release from the leading to significant climate impacts and ozone damage (Berner, 2002; Grice et al., 2005a; Kump et al., 2005; Meyer & Kump, 2008). Many of these mechanisms have significant knock-on effects and create positive feedback loops with other mechanisms, such as runaway global warming caused by CO2 creating the conditions that allow for seafloor and permafrost methane clathrate release causing further global warming (Wignall, 2001).

25

Figure 2.8: Schematic depiction of potential interrelated processes occurring at the P-T boundary (Algeo et al., 2011; modified from Wignall, 2001).

A bolide impact has also been proposed as a potential trigger for the extinction event with the Bedout structure off the coast of Western Australia suggested as a potential impact crater (Gorter, 1996; Becker et al., 2004, Glikson et al., 2004). The timing of a potential impact and the nature of the Bedout structure as an impact crater has been disputed (Wignall et al., 2004, Renne et al., 2004). Kaiho et al. (2001) present sulphur and strontium isotope excursions as evidence for mantle-derived sulphur release caused by a bolide impact. However, the δ34S values may also be the result of fractionation caused by bacterial sulfate reduction in euxinic oceanic conditions (Koeberl et al., 2002). Retallack et al. (1998) presented evidence of shocked quartz and a low magnitude iridium anomaly from localities in Australia and Antarctica, though their placement of the P-T boundary in the Antarctic localities was disputed by Isbell et al. (1999). Microstructures within these shocked quartz grains from Graphite Peak, Antarctica were later re-analysed and re-interpreted as being inconsistent with an impact origin (Langenhorst et al., 2005).

26

2.4.6 Proxies for the EPE and P-T Boundary

When examining potential boundary sections, obtaining conodont or radiometric age dates is often difficult or not possible. Ascertaining the nature and position of the boundary is particularly difficult in terrestrial sections. Thus workers rely on a number of proxies that act as indicators of the boundary. More accurately, ones that reflect extreme environmental perturbations that we associate with the Permian-Triassic boundary. These proxies include: stable isotope spikes (mainly carbon and sulphur), evidence of macro- and microfloral turnover, faunal extinction, spikes in fungi and/or acritarchs, pyrite and indicators of oceanic anoxia.

A negative carbon-isotope excursion has been frequently noted at many Permian-Triassic boundary localities and is particularly useful for picking out the boundary. The excursion is generally observed to occur synchronously with the extinction horizon or slightly above it (Looy, 2000; Twitchett et al., 2001; Sephton et al., 2002), though there is some evidence of multiple excursions that occur either above or below the boundary (Xie et al., 2007). There are several proposed mechanisms for the negative carbon isotope excursion and multiple mechanisms may be responsible. These include increased erosion, reduced primary productivity, volcanism, oceanic anoxia, and the release of methane clathrates (Korte & Kozur, 2010).

The Permian-Triassic boundary is also associated with notable increases in so-called disaster taxa. A “fungal spike”, usually represented by an increase in the putative fungal spore taxon Reduviasporonites chalastus, has been noted in many localities and is thought to represent saprophytic fungi taking advantage of the decaying plant material (Eshet et al., 1995; Visscher et al., 1996; Elsik, 1999; Steiner et al., 2003; Peng et al., 2005; Spina et al., 2015; Bercovici et al., 2015). Some authors suggest R. chalastus actually has an affinity to shallow-water algae (Afonin et al., 2001; Foster et al., 2002; Hochuli, 2016), though chemical analysis of the spore wall has been inconclusive (Sephton et al., 2009). Acritarchs have also been noted to increase in marine and marginal-marine P- T boundary sections, which may be related to increased nutrient input from the terrestrial realm causing algal and/or acritarch blooms (Retallack, 1995; Vajda & McLoughlin, 2007; Gorter et al., 2009; Lei et al., 2013; Luo et al., 2013; Rampino & Eshet, 2018). Biomarkers linked to these acritarch blooms have been observed in this stratigraphic interval in basins in Western Australia and Greenland (Grice et al., 2005b; McIldowie & Alexander, 2005).

27

2.4.7 The Boundary in Australia

Even before the formal recognition of an end-Permian mass extinction (Erwin, 1993), Australian researchers noticed the rapid floral turnover at the end of the Permian from a Glossopteris-dominated community to one initially dominated by lycopods and later Dicroidium (Balme & Helby, 1973). Initially, this was regarded to be the result of marine transgression based on the expansion of post- extinction floras and abundance spikes in spinose and non-spinose acritarchs (Balme & Helby, 1973; Foster, 1982). Foster (1982) and Retallack (1995) examined the rapid changes in the flora and microflora in the Bowen and Sydney basins, respectively.

Several studies have focused on the marine sections of the Perth, Carnarvon, and Bonaparte basins off the coast of Western Australia due to the extensive exploration for hydrocarbons in that area. A particularly notable borehole is Hovea 3. This borehole was extensively studied using palynology, palynofacies, hydrocarbons and carbon-isotopes (Thomas et al., 2004; Grice et al., 2005a; Grice et al., 2007). Metcalfe et al. (2008) presented conodont data calibrated to the palynological biostratigraphy which allowed for the placement of the P-T boundary in the Sapropelic Interval of the Hovea Member in the Perth Basin. Gorter et al. (2009) collected and reassessed carbon-isotope and geochemical data in the Perth and Bonaparte basins to place the boundary in a stratigraphic and palynological context, and established that the P-T boundary occurs within the Protohaploxypinus microcorpus Oppel Zone (APP6). They also suggested that the rapid carbon-isotope excursion in the Hovea 3 well represents an unconformity at the top of the P. microcorpus Oppel Zone in the northern Perth Basin, whereas the more gradual isotope excursion in the Bonaparte Basin was coupled with seismic data to place the unconformity at the base of the P. microcorpus Oppel Zone.

2.4.8 Palynology of the Boundary in Eastern Australia

In the terrestrial basins of eastern Australia, linking the Permian-Triassic boundary to the biostratigraphic schemes has proved difficult. Environmental changes at the end of the Permian were not conducive to the preservation of organic material, and high erosion rates created unconformities that removed sections that may have contained the boundary (Brakel et al., 2009). Further complicating that is the rank of the uppermost Permian coals is commonly too high to extract and identify palynomorphs close to the boundary interval. In the sections that did yield palynomorphs, workers were able to define a number of biostratigraphic changes occurring in a short amount of time (Foster, 1982; Price, 1997). The zones immediately overlying the Dulhuntyispora parvithola Zone (APP5 in the scheme of Price, 1997) are APP6 and APT1. APP6 can be separated into APP6.1 (equivalent to the Playfordiaspora crenulata Zone of Foster, 1982) and APP6.2 (equivalent to the

28

Protohaploxypinus microcorpus Zone of Foster, 1982). APT1 is marked by the first appearance of Lunatisporites pellucidus.

The placement of the boundary within one of these zones or at the interface of two zones has changed over time. Initially, it was thought that the Playfordiaspora crenulata zone represented the last remnant of the Glossopteris flora (Foster, 1982) and that APT1 represented the first expression of the Triassic flora, with the P-T boundary placed at the base of APT1 (Michaelsen, 2002). Currently, the APP6 zone is recognised as representing a post-extinction flora (et alMays et al., 2020a; Vajda et al., 2020). However, Fielding et al. (2019) use absolute age dates to suggest that the terrestrial extinction interval (marked by the base of the APP6 zone above the uppermost Permian coal seam) occurs earlier than the GSSP-defined boundary, and is instead synchronous with a primary extrusive phase of the Siberian Traps (Fig. 2.9).

29

Figure 2.9: Correlation of the P-T boundary section from the borehole Bunnerong-1 in the Sydney Basin, Australia to the Meishan section, China and calibrated using absolute age dates to the GSSP and the phases of volcanism from the Siberian Traps. This shows a temporal gap between the terrestrial extinction interval and the GSSP marking the end of the Permian (modified from Fielding et al., 2019).

2.4.9 Summary

The current confusion around the position of the boundary lies in the disconnect between the boundary as defined by the EPE, the carbon-isotope excursion, and the chronostratigraphic age. Previous studies have been directed at only a few localities and do not examine potential lateral facies

30 changes and local variations in palaeoflora and palaeoenvironment, both of which can strongly influence the geochemical proxies. This thesis will also examine the palaeoenvironment in order to examine potential causes for the coal gap, whether that is related to aridity or perhaps factors that may otherwise inhibit peat formation, such as increased sediment input during orogenesis and development of a rapidly subsiding foreland system across Gondwana.

31

Chapter 3 was was published in 2021 in Review of Palaeobotany and Palynology (284, 104343) as:

3. Acid-free palynological processing: a Permian case study

Alexander Wheelera, Patrick T. Mossa, Annette E. Götzb, Joan S. Esterlea, Daniel Mantlec

aSchool of Earth and Environmental Sciences, The University of Queensland, St. Lucia, QLD 4072, Australia

bState Authority for Mining, Energy and Geology, Stilleweg 2, 30655 Hannover, Germany

cMGPalaeo PTY LTD, 5 Arvida Street, Malaga, WA 6090, Australia

3.1 Abstract

Acid-free palynological processing was performed on Permian-aged material from a borehole of the southern Galilee Basin, eastern Australia, to test the efficacy compared to standard processing techniques. Both techniques yielded well-preserved assemblages of terrestrial and aquatic palynomorphs. The proportion of phytoclasts in the acid-free preparations was much higher than in samples that underwent conventional acid processing, while spores and pollen grains occur in lower abundances. A slightly higher species count was gained from samples with acid-free treatment, but biostratigraphic index taxa were present in both sets of samples and, in general, the assemblages appeared to be quantitatively comparable. To improve the efficacy for future application, in particular more accurate statistical analyses with higher counts, absolute abundance calculations using a Lycopodium spike and refinement of the acid-free technique using peroxide treatment are suggested. Further work is required to determine if the higher abundances of algae are influenced by the processing technique or if they better reflect a true proportion within a given sample.

3.2 Introduction and aims

The process of liberating organic material, particularly microfossils, from sedimentary successions is necessary to interpret biostratigraphy as well as environments of deposition. Many papers examining pollen and spores use the term “Standard palynological processing” when referring to their processing technique. This generally concerns a series of acid-treatments (usually hydrochloric and hydrofluoric acid) that remove the mineral components and liberate the organic material within a sample. This is sometimes followed by a density separation (float/sink) using heavy liquid (zinc bromide or zinc chloride, and more recently lithium or sodium heteropolytungstate) and/or sieving. However, many researchers and commercial laboratories have developed their variations on this technique that works

32 for them and for the particular type of samples they examine (some of these techniques are summarised in Wood et al., 1996; Traverse, 2007; Brown, 2008; Lignum et al., 2008). This is undertaken because sediment and rock samples containing palynomorphs vary broadly in their mineral composition, thermal maturity, grain size and organic content. The palynomorphs themselves may also be in varying states of preservation and may require acetolysis or oxidation in nitric acid

(HNO3) for accurate examination and identification. Thus, techniques need to be tailored to the type of sample processed to ensure efficient extraction of the palynomorphs contained within.et al.

Because of the efficacy of acid treatments, the use of acid-free techniques is restricted to a few labs and is usually only applied to relatively young, i.e., Cenozoic samples that are not highly lithified. Acid treatments can be problematic however due to the inherent danger of using strong acids such as hydrofluoric acid (HF). Safety equipment and laboratories equipped with fume hoods are required when using these acids. Increasingly strict safety regulations applied to the storage, use and disposal of these have made it more difficult for palynological processing to be undertaken in university laboratories or in the field. Facing these challenges, it has become increasingly desirable to develop more effective acid-free palynological processing techniques (O’Keefe & Eble, 2012). However, acid-free palynological techniques have only been employed on Palaeozoic-aged material on very few occasions and never before on Permian-aged samples (Riding & Kyffin-Hughes, 2006; 2011; Riding et al., 2007).

This study investigated the efficacy of a processing technique previously used only on Cenozoic-aged samples (Van der Kaars, 1991; Moss et al., 2005; Moss & Kershaw, 2007; Moss, 2013; Moss et al., 2016) to Permian-aged material. Simple taxonomic and palynofacies counts were used to quantify and compare yields and compositions from the standard and acid-free approaches and to evaluate potential processing biases.

3.3 Materials and methods

3.3.1 Samples

Ten samples were selected from the borehole Tambo 1-1A located on the Springsure Shelf in the southern part of the Galilee Basin (Fig. 3.1). Samples were collected from the Mantuan Productus Beds, Black Alley Shale and ‘Burngrove Formation’ equivalent. These units were chosen to represent a progradational deltaic sequence (Phillips et al., 2017a). Samples represent a consistent biostratigraphic zone but have a variety of terrestrial and aquatic (freshwater and marine) palynomorphs. Approximately 3 to 5 cm thick half-core samples were collected preferentially

33 selecting siltstones and mudstones. Samples were then split and half were sent for conventional and the other half for acid-free processing.

Figure 3.1: Map and lithological log in stratigraphic context showing the position of Tambo 1-1A in the Galilee Basin of eastern Australia and the samples taken from the borehole.

3.3.2 Palynological Processing Techniques

Conventional processing (MGPalaeo)

Samples were sent to MGPalaeo, a commercial laboratory based in Perth, Western Australia for conventional acid processing (Wood et al., 1996). The core samples were washed to remove any drilling mud additives or modern pollen contaminants. Fifteen to 23 g of each sample was carefully crushed to 2–4 mm diameter pieces before being treated with 32% hydrochloric acid (HCl) to digest any carbonate minerals. The residues were then decanted and treated with 48% hydrofluoric acid (HF) to remove minerals. When the material had been sufficiently digested (after 24 hours for these preparations), the remaining HF was decanted, and the samples underwent repeated water washes with the centrifuge to neutralise any remaining acid. A small volume (10–15 ml) of lithium heteropolytungsate (LST; 2.1 specific gravity) was added to each sample and thoroughly mixed with the residue prior to further centrifuging. This heavy liquid separation technique concentrated the organic float with the inorganic residue settling to the base of the test tube or remaining in solution.

34

The organic float was then washed through a 10 µm polycarbonate filter to remove the unwanted fine fraction of palynodebris.

Acid-free processing (University of Queensland)

This technique is a modification of a process used by Van der Kaars (1991) to extract palynomorphs from deep-marine sediments (see Appendix C1). Samples undergoing the acid-free processing are crushed down to a fine sand to silt size fraction (<250 μm). Between 7 and 10 g of material is generally sufficient for processing. Forty ml of 10% sodium hexametaphosphate (commercially known as Calgon) is added to the samples which are heated and left to settle overnight to deflocculate the clays. Van der Kaars (1991) originally employed tetrasodium pyrophosphate as a deflocculant but sodium hexametaphosphate is an adequate replacement, the merits of which are discussed in Riding & Kyffin- Hughes (2004). Next, the material is sieved through a 250 μm sieve to remove sand and gravel particles and through an 8 μm mesh to remove clay particles. Then, the 8 to 250 μm size fraction undergoes a series of water washes using the centrifuge. 6 ml of sodium polytungstate (SPT; 1.9 specific gravity) is then added to the remaining material, which is centrifuged for 30 minutes at 2500rpm. Lastly, the organic material on the surface is collected, water-washed again and strew- mounted on a slide and dried using a slide water before applying the cover slip using Eukitt, a resin- based mounting medium.

3.3.3 Microscopy

Both sets of samples were examined on a Zeiss Photomicroscope III with an attached Leica MC190HD camera and photographed using the LAS software package. Taxonomic counts were made up to 200 specimens. Palynofacies counts of at least 500 particles were done and classified according to a modification of the scheme used by Feist-Burkhardt et al. (2008) (Fig. 3.2). Biostratigraphic counts were made and plotted using the Tilia palynological software package (Grimm, 2004). Identification of key index taxa for the late Permian was based on the scheme of Price (1997).

35

Figure 3.2: Classification scheme for palynofacies (from Feist-Burkhardt et al., 2008).

3.4 Results

3.4.1 Palynofacies

All samples processed using the acid-free method apart from TAMP28 show a much higher proportion of opaque phytoclasts than the standard processing. The abundance of opaque phytoclasts in the conventional samples is between 14.4% and 52.87% while the abundance in the acid-free samples is between 38.08% and 66.28%. The total relative abundance of phytoclasts (opaque and translucent) is higher in the acid-free samples. This can be seen in the counts (Fig. 3.3) and by visual inspection (Fig. 3.5; 3.6; 3.7). Larger opaque particles tend to clump together and may obscure palynomorphs below them. The classification degraded phytoclasts refers to clumps of fragmented phytoclast material. These are particularly abundant in the conventional samples TAMP23, TAMP28 and TAMP30. The high abundances of degraded phytoclasts in these samples made identification and counting of palynomorphs challenging and time consuming, potentially obscuring some key taxa. Spore and pollen grain abundances are also relatively low in the acid-free samples but yields and preservation are generally good enough for sporomorph counts. The conventional samples have pollen grain and spore yields that tend to vary but are highest in samples TAMP26 and TAMP27. TAMP23 produced very low yields of palynomorphs using both processing techniques. While some pollen grains and spores are observed in the conventional sample, they are too damaged or obscured

36 by degraded phytoclasts to count and thus both samples were regarded as “barren”. Freshwater and brackish algal abundances tend to be relatively low in all samples apart from a small peak in Botryococcus in the acid-free sample TAMP27. The conventional sample also features some Botryococcus, but these specimens are not as large or as abundant.

Figure 3.3: Relative abundance bar chart for the palynofacies data comparing the acid-free (left) and conventional (right) processing techniques for each sample. For raw palynofacies counts see Appendix B5.2.

3.4.2 Palynology

The acid-free processing yielded 85 different taxa in all 10 samples while the conventional processing yielded only 77 taxa (Fig. 3.4). The species counts also show some differences between the processing techniques. All of the major expected pollen and spore taxa appear in both sample sets. The common spore taxon Leiotriletes directus appears to occur at a higher relative abundance in the conventional samples, particularly in the lower part of the studied interval. Striate bisaccate pollen grains (Protohaploxypinus sp., Alisporites sp.) appear to occur at slightly higher abundances in the acid-free samples, but this can vary between different taxa. Protohaploxypinus limpidus is the most commonly observed bisaccate pollen taxon and tends to have very similar abundances for each processing technique in most samples. Dulhuntyispora parvithola, a key index taxon for the late Permian in

37

Australia (Price, 1997), appears in eight samples in the acid-free samples compared to six in the conventional samples. It appears at a particularly high abundance in the acid-free sample TAMP28 (17%). Another key index taxon for the late Permian in Australia is Microreticulatisporites bitriangularis. It appears in four of the acid-free samples compared to just one of the conventional samples. The abundance and diversity of algae appears to be different for each type of processing. The acid-free samples appear to yield a more diverse assemblage of algae and comparatively higher abundances of acritarchs in samples TAMP31 and TAMP32 as well as higher abundances of algae, particularly Botryococcus in TAMP27 (38%). However, the Micrhystridium evansii acritarch acme is well represented via both techniques. Micrhystridium evansii makes up 76% and 83% in the acid- free samples TAMP31 and TAMP32 respectively and 68.5% and 75.5% in the conventional samples. In samples TAMP28 and TAMP29, the abundance of Micrhystridium sp. is higher in the conventional samples (3.5% and 4.5% respectively) compared to the acid-free samples (2% and 0.5% respectively).

38

Figure 3.4: Relative abundance chart showing a comparison of the acid-free processing (red) and the conventional processing (blue) for each species observed and counted. For raw counts see Appendices B4.3 and B4.4.

39

3.5 Discussion

3.5.1 Is the acid-free process effective?

The acid-free process effectively yields a diverse range of well-preserved palynomorphs though there are some caveats. While there are apparent differences in species diversity between the two processing techniques, these extra species that appear in the acid-free processing all occur in very rare abundances and thus do not amount to any significant difference in species diversity. However, even without measuring absolute concentrations using a marker species such as Lycopodium (Mertens et al., 2009), it is clear that palynomorph yields are generally higher when employing the conventional processing. This is partly due to the numerous large phytoclast fragments overcrowding and covering palynomorphs, but also potentially because palynomorphs may struggle to work through the sediments during centrifuging and density separation. This is most apparent in samples which contain a high abundance of clay even after repeated sieving. The clays sit on top of the sand and silt particles during centrifuging, forming a cap that may potentially block palynomorphs from rising to the top of the sodium polytungstate. This problem can usually be alleviated with multiple treatments of sodium polytungstate to increase yields. But even with this issue, yields were suitable in the processed samples to easily count to 200 palynomorphs in every sample apart from TAMP23, which was barren of palynomorphs regardless of sample protocol.

3.5.2 How does the acid-free process compare to the conventional process?

Even small differences in processing techniques can have strong effects on the palynomorph yield and composition of the assemblage. Lignum et al. (2008) demonstrated that just changing the type of mesh used for sieving can greatly influence the cyst assemblage. In this work, each technique uses meshes of slightly different sizes (8 and 10 µm) and different materials (nylon vs polycarbonate). Characteristic differences between the mesh material such as flexibility might allow larger particles to squeeze through to inconsistent degrees. Sieving is also done at different stages of the process for different purposes. Sieving is done in the acid-free technique before density separation to remove clay particles, while it is used post-density separation in the conventional technique to remove fine organic material. Having said that, the palynological data presented herein show comparable results even though some differences in the composition of the assemblages could potentially be attributed to the different processing techniques. Any differences noted are also not surprising based on the relatively low number of palynomorphs counted, and the biostratigraphically important late Permian taxa for eastern Australia are represented in both assemblages. Sieving of the deflocculated clay particles potentially results in lower palynomorph yields than the conventional

40 processing technique (Riding & Kyffin-Hughes, 2011) and must be done carefully to remove as many fine clay particles as possible that can obscure the palynomorphs. Sodium polytungstate (SPT) has previously been used instead of the more toxic zinc bromide to density separate a variety of living and fossilised microorganisms (Savage, 1988; Munsterman & Kerstholt, 1996; Bolsch, 1997). While the initial costs of purchasing sodium polytungstate can be high, it can also be recycled and reused multiple times (Six et al., 1999). Tests on modern lake and peat samples using more statistically rigorous counts show different but relatively comparable pollen yields when comparing HF and sodium polytungstate (Leipe et al., 2019). Moss et al. (2005) achieved pollen yields from the mid Eocene Republic Formation in Washington, USA using the acid-free techniques, which had been previously described as barren using the standard acid technique. The high abundance of degraded phytoclasts in several samples processed using the conventional technique and the comparatively low abundance in their acid-free counterparts suggests that these are an artefact of the processing technique. They are easy to recognise and differentiate from amorphous organic matter (AOM) and true plant detritus (Tyson, 1995); however, their presence can be problematic as they can affect palynofacies counts and obscure palynomorphs. This is potentially the reason for the large difference in the abundance of the relatively fragile Brazilea scissa in TAMP30, whereas Leiotriletes directus is robust and easily identifiable.

3.5.3 How does this acid-free technique compare with other non-standard processing methods?

Riding & Kyffin-Hughes (2004; 2006) and Riding et al. (2007) developed and refined a technique employing sodium hexametaphosphate (NaPO6) and treatments of hydrogen peroxide (H2O2) to initially deflocculate the clays and then oxidise the phytoclast material to concentrate palynomorphs. This technique was tested on samples of a variety of ages (, Carboniferous, Jurassic and Palaeogene) with variable results. While Riding & Kyffin-Hughes (2004; 2006) employ sodium hexametaphosphate in a similar way to that presented in this work, their technique is not followed up by a density-seperation step. Hydrogen peroxide is a powerful oxidising agent and can also damage palynomorph assemblages (Hopkins & McCarthy, 2002). The technique employed by Riding et al. (2007) uses minimal exposure times of hot peroxide to sample material. This both disaggregates the sediment and floats the organic fraction. It must be done quite carefully though as hydrogen peroxide has also been noted to selectively damage dinoflagellate cysts (Hopkins & McCarthy, 2002). Algal bodies such as Botryococcus have relatively low preservation potential and are easily broken up during sieving and centrifuging (Tyson, 1995; Feist-Burkhardt et al., 2008). The acid-free technique employed in this work appears to favour preservation of Botryococcus, but further work is required

41 to determine if this is due to the processing technique, and if it accurately reflects the true assemblage. While the consistent appearance of algal taxa, such as Pilasporites, in multiple samples processed using the acid-free technique and not in the conventionally processed samples could be a random effect, it may suggest algal extraction is slightly favoured using the acid-free technique.

3.6 Conclusions and outlook

The conventional processing technique using HF and HCl remains an effective way to liberate palynomorphs from most types of pre-Cenozoic rock and sediments. However, this study shows that the acid-free treatment is also effective at liberating palynomorphs from Permian-aged samples. The advantages of this acid-free palynological processing technique lie in its safety relative to HF treatments and the simplicity of the technique, which offers wide applicability. However, further testing and refinement of this technique would be useful to improve its efficacy and to better understand how it compares to conventional methods. For more accurate statistical analysis higher counts of up to 1000 specimens would be more suitable particularly in the case of Permian samples, where common taxa such as Protohaploxypinus and Leiotriletes can overcrowd counts. A better test of this technique, though, would be further application on a wider variety of samples from different lithologies and ages. The addition of a Lycopodium tablet would also be useful to calculate the yield of palynomorphs per gram of sediment processed. This would allow us to estimate any losses and changes in proportion related to sieving and centrifuging, and would help better understand statistical variability between the two processing techniques. A further refinement of the acid-free technique could potentially involve the use of peroxide treatments to remove phytoclasts. The limits of this technique also require further examination. Testing this acid-free technique on the clay-rich, organic- poor samples such as those of the Early Triassic Rewan Group may prove informative as well on carbonate-rich rocks, which might require some addition of HCl to remove the minerals.

42

Figure 3.5: Photomicrograph comparison of the conventional processing technique (left) and the acid-free technique (right). A – B: TAMP23; C – D: TAMP24; E – F: TAMP25; G – H: TAMP26; Notable features include: A – degraded phytoclasts; B, D, F & G – large opaque phytoclasts; C, E & G – well preserved bisaccate pollen grains

43

Figure 3.6: Photomicrograph comparison of the conventional processing technique (left) and the acid-free technique (right). A – B: TAMP27; C – D: TAMP28; E – F: TAMP29; Notable features include: A – small Botryococcus; B – large Botryococcus; C – degraded phytoclasts; D, E & F – large, well-preserved palynomorphs and phytoclasts.

44

Figure 3.7: Photomicrograph comparison of the conventional processing technique (left) and the acid-free technique (right). A – B: TAMP30; C – D: TAMP31; E – F: TAMP32; Notable features include: A – degraded phytoclasts; B – large spores, bisaccate pollen grains and opaque phytoclasts; C & E – Micrhystridium evansii; D & F – Micrhystridium evansii and large phytoclasts.

45

4. Linking lithofacies and palynology to examine proximal-distal palaeofloral patterns in the late Permian of the Bowen and Galilee basins

4.1 Abstract

The composition and environmental controls affecting the Glossopteris-flora of Permian Gondwana is still poorly understood. Herein we examine proximal-distal trends in the distribution of floral assemblages in the Bowen and Galilee basins of eastern Australia using palynology as a proxy. This work compiles a palynological dataset from five boreholes interpreted to represent a transition from the alluvial plain more proximal to the upland through a deltaic environment to a distal shallow marine setting. By interpreting the composition of the palaeoflora based on the palynology and linking it with a depositional environment interpretation based on lithofacies data we were able to develop a high- resolution model of the composition and environmental preferences of the late Permian floras of eastern Australia. Glossopteris forests are widespread in both the delta and alluvial plains forming dense forests on the floodplain as well as a part of the coal-forming flora. occupy the understory in peat-forming communities but also act as colonisers and pioneers in the lower delta plain facies. Horsetails appear underrepresented in palynological datasets compared to their diverse macrofloras, in part due to the sheer abundance and diversity of fern spores, but this could also be due to the different reproductive styles of horsetails as they have the ability to reproduce asexually via progeneration from rhizomes. Forests occupying the coastal plain appear to occur as mix of Glossopteris, conifers and cordaitaleans and are more common in the sand-dominated paralic facies. The margins of the basins are occupied by a xerophytic upland flora made up of conifers, cordaitaleans and potentially peltasperms. The abundance of Dulhuntyispora parvithola appears consistently higher in coastal settings suggests that the parent plant preferably grew in paralic environments, which has widespread implications for its use as a late Permian index taxon.

4.2 Introduction and Aims

4.2.1 Motivation

Most palynologists studying the Galilee and Bowen basins of eastern Australia were focused on defining a biostratigraphic scheme for the region (Evans, 1969; Paten, 1969; Price, 1983; Price, 1997). While effective at a low time-resolution, problems are encountered when trying to correlate at a formation scale, as the index taxa for the subzones do not appear consistently throughout these basins

46

(Nicoll et al., 2015). The late Permian deposits of the Bowen and Galilee basins were deposited in a particularly complex and dynamic system of changing depositional environments. Obtaining a high level of detail in reconstructing these environments is challenging because of large data gaps arising from sparse core drilling and relatively few regional seismic surveys that attempt to connect these basins. Phillips et al. (2017a) used lithofacies and large-scale correlations to interpret and describe the palaeoenvironment of the Galilee Basin in high detail. Palynology can augment reconstructions of late Permian palaeoenvironments, particularly when used as part of a multi-proxy dataset with sedimentological and geochemical data or coal petrology (e.g., Van de Wetering 2013a; 2013b). By applying detailed palynological data alongside our understanding of lateral facies changes we here aim to create a detailed image of palynofloral assemblages related to specific depositional environments and facies as well as gain insight into the dominant processes controlling distribution of important index taxa.

4.2.2 Aims

The aims of this study are to:

1) examine temporal and spatial variation in palynological assemblages in a climatically stable environment dominated by Glossopteris vegetation, from the surrounding hinterlands through the proximal alluvial plain into the distal lower delta plain and nearshore setting; and

2) broadly reconstruct the floral distribution based on known and inferred botanical affinities and compare this dataset to the lithological changes in each section in order to better understand environmental controls on the in-situ palaeofloral composition and distribution

Carbon-isotope data and palaeoclimate models examining the climate of Gondwana during the Permian (Roscher et al., 2011) suggest the late Permian climate of the Bowen Basin is relatively stable, if warming slightly (Van de Wetering, 2013a; Ayaz et al., 2016b). These models generally place all of eastern Australia within the cool-temperate climate belt (Kutzbach & Ziegler, 1993; Roscher et al., 2011). A cooling event during the Changhsingian has been proposed based on climate- indicative brachiopods and bivalves, but further refinement of the timing and correlation of this event is needed to examine any potential effects on terrestrial floras (Waterhouse & Shi, 2013). We hypothesize that palaeoenvironment (water level, salinity, sedimentation, etc.) is the dominant control on the floral distribution in the basins and that this distribution will be observable in a palynological dataset. Distinctly climate-driven changes to the flora should only be observable at the Permian- Triassic boundary at which time at the latest, of course, also plant-evolutionary changes come into play in modifying vegetation composition (as opposed to just palaeoenvironment).. Glossopterids,

47 ferns and horsetails make up the majority of Gondwanan floral assemblages, with minor but consistent components of conifers, cordaitaleans, lycopods and cycads (Shi et al., 2010). The presence of ginkgos, Gnetales as well as corytosperm and peltasperm seed ferns is less conclusive based on a lack of strong macrofloral evidence of their presence in basinal settings in eastern Australia during the Permian, and the difficulty of using their dispersed polyphyletic pollen to infer their presence.

4.2.3 Botanical Affinities

Even though taxonomic resolution is limited and inconsistent, published reports of in situ sporomorphs and their correlation to dispersed sporomorph taxa (as compiled in, e.g., Balme 1995) enable at least a partial identification of some vegetation components that are recorded in the palynomorph assemblages. In general, various major groups of algae, fungi, and land plants are represented by the palynomorph assemblages of late Permian Australia, and thus off some utility in the reconstruction of basinal floras when macrofloral remains are not present. In the late Permian formations of the Bowen and Galilee basins, terrestrial floras are mainly represented by spore- producing pteridophytes, pollen-producing gymnosperms and fungal remains. Algae and acritarchs represent the aquatic environment, both freshwater and marine.

Algal microfossils with modern-day equivalents such as Botryococcus and Cymatiosphaera can be confidantly assigned a Chlorophycean and Prasinophycean affinity respectively (Colbath & Grenfell, 1995). It is more difficult to assign groups such as the leiospheres or acritarchs, which are classified based on morphology and not on their taxonomic affinities. Acanthomorph acritarchs such as Micrhystridium, however, are distinct and useful environmental indicators and have thus been assigned separately. Maculatasporites is sometimes given a bryophyte affinity (Hochuli et al., 2010) but following Zavattieri et al. (2017) we have assigned it a zygnematacean algal affinity, along with Mehlisphaeridium, Brazilea, Peltacystia, Circulisporites and Tetraporina (Table 1) (Grenfell, 1995; Mays et al., 2020b).

Permian pteridophytes are diverse and comprise of lycopodiopsids (club mosses), equisetopsids (horsetails) and filicopsids (ferns). Selaginellalean and isoetalean lycopods are represented by the cavate trilete spores (Indotriradites, Gondisporites, Grandispora, etc.) (Balme, 1995; Hochuli et al., 2010) (Table 1). In Permian Gondwana, Calamospora and Laevigatosporites are usually assigned to horsetail, but have also been associated with a variety of other plant groups (Balme, 1995), and thus, for our purposes are assigned to the undifferentiated pteridophytes while Retusotriletes and Columnisporites have been assigned to the equisetopsids (Balme, 1995). Various orders of ferns

48

(Osmundales, Botryopteridales, Marattiales) are identifiable based on their spores (Table 1). Some spore genera, like Horriditriletes, Lophotriletes and Osmundacidites have been found along with in situ remains of putative Osmundales, Neomariopteris and Dichotomopteris in Gondwanan sections (Lele et al., 1981). Dictyotriletes, Microreticulatisporites, Camptotriletes and Raistrickia are attributed to the botryopterid ferns. Marattialean ferns share many spore taxa with other fern orders (Cyclogranisporites, Verrucosisporites) and are thus difficult separate within a palynological dataset (Balme, 1995). Leiotriletes (=Deltoidospora), a widely abundant taxon, has been associated with several pteridosperm groups. Permian forms have generally attributed it to several filicopsid orders (Balme, 1995; Lindström, 2005; Schneebeli-Hermann et al., 2015). The spore genus Dulhuntyispora is a key taxon for late Permian biostratigraphy and is endemic to Australia with only a single other occurrence recorded in South Africa (Anderson, 1977). The potential evolutionary lineage of the spore genus Dulhuntyispora has been described (Price & Filatoff, 1990), but as of yet no published records exist of in situ spores found with plant remains and, as such, no affinity can be assumed further than a broad pterophyte origin.

Glossopterid pollen-producing organs like Arberiella have been observed to produce different form genera and species of pollen within single sporangia (Lindström et al., 1997). Glossopterids have a widely reported association with taeniate bisaccate pollen grains (commonly Protohaploxypinus and Striatopodocarpites but also including Striomonosaccites and Weylandites) (Table 1) (Gould & Delevoryas, 1977; Zavada, 1991; Lindström et al., 1997; Nishida et al., 2014; Gastaldo et al., 2015). Non-glossopterid gymnosperms groups are represented by a diverse assemblage of pollen. Cordaitaleans are mainly represented by monosaccate pollen, while conifers are represented by non- striate bisaccate pollen (Balme, 1995). Vitreisporites and Vittatina have both previously been attributed to , but this association is somewhat tentative in this context as peltasperm macrofloral remains have not yet been identified in Permian Australia. Some non-striate bisaccate pollen grains such as Alisporites and Scheuringipollenites can potentially be attributed to multiple gymnosperm clades. The same is true for the monosulcate pollen genus Cycadopites, which in Permian Australia likely represents cycads but potentially also can be associated with peltasperms.

49

Group Affinity Palynomorph Genus Acritarchs Acanthomorphs Micrhystridium Algae Zygnamatacean Algae Brazilea Maculatasporites Circulisporites Tetraporina Peltacystia Mehlishphaeridium Incertae sedis Quadrisporites Rugaletes Pilasporites Chlorophycean Algae Botryococcus Prasinophycean Algae Cymatiosphaera Leiosphaeridia Fungi Undifferentiated fungal Reduviasporonites spores Pteridophytes Lycopodiopsida Indotriradites Secarisporites

Gondisporites Undifferentiated cavate trilete spores Densoisporites Lundbladispora

Foveosporites

Equisetopsida Retusotriletes Columinisporites Filicopsida Brevitriletes Microbaculispora

Dictyophillidites Dulhuntyispora

Didecitriletes Grandispora Interradispora Leiotriletes Concavissimisporites Microfoveolatispora

Cyathidites Phaselisporites Indospora Pseudoreticulatispora Filicopsida - Osmundales Osmundacidites Horriditriletes

Lophotriletes Baculatisporites

Filicopsida - Botryopteridales Dictyotriletes Microreticulatisporites

Raistrickia Camptotriletes

Filicopsida - Marrattiales Thymospora

Undifferentiated Calamospora Punctatisporites Pteridophytes Cyclogranisporites Verrucosisporites

Laevigatosporites Gymnosperms Glossopterids Marsupipollenites Striomonosaccites Protohaploxypinus Weylandites

Schizopollis Striatoabieites Striatopodocarpites Guttulapollenites

Peltaspermales Vittatina Vitreisporites

50

Coniferales Chordasporites Limitisporites Klausipollenites Triadispora Lunatisporites Potonieisporites Cordaitaleans Barakarites Parasaccites Cannanoropolis Plicatipollenites Florinites Undifferentiated Alisporites Tiwarisporites Gymnosperms Praecolpatites Scheuringipollenites Undifferentiated Striate Cycadopites bisaccate pollen grains

Bascanisporites

Table 1: Botanical affinities (Balme, 1995; Grenfell, 1995; Lindström et al., 1997; Lindström, 2005; Hochuli et al., 2010; Schneebeli-Hermann et al., 2015 Gastaldo et al., 2015; Zavattieri et al., 2017; Mays et al., 2020b).

4.2.4 Geological Setting

The intracratonic Galilee Basin and adjacent foreland Bowen Basin are Permian-Triassic-aged sedimentary basins that cover an area of nearly 317 000 km2 in Queensland, Australia (Huleatt 1991; Allen & Fielding, 2007). The sedimentary fill in both basins records a wide array of depositional environments, from alluvial to marine, and can reach a thickness of up to 10 km in the major depocentres. The basins are separated in the north by the Anakie Inlier, and are connected in the south across the Springsure Shelf (Fig. 4.1). The stratigraphy in both basins is correlatable to a degree that same names for many stratigraphic units are often used in both basins.

The Bowen and Galilee basins are commonly split into several structural domains which are used to place localities in a geographical context within the basin. The Collinsville Shelf and Nebo Synclinorium make up the northern part of the Bowen Basin; the Springsure Shelf, Denison Trough and Roma Shelf make up the west and south-western margin of the basin; the Comet Ridge and Taroom Trough make up the central, eastern and southern parts of the basin (Draper, 2013) (Fig. 4.1).

The Bowen Basin is informally split into three tiles based on the nomenclature used for the coal measures. These are: The Rangals tile (north), Bandanna tile (west) and Baralaba tile (east) (Fig. 4.1) (Sliwa et al., 2017). The Galilee Basin features three major depocentres (the Koburra Trough, Powell Depression and the Lovelle Depression) and is split into a northern and southern section by a palaeotopographic basement high called the Barcaldine Ridge (Van Heeswijck, 2010).

The depositional environment is of both basins generally made up of fluvio-deltaic systems interpreted to prograde southward into more open, often marine settings (Phillips et al., 2017a; Sliwa

51 et al., 2017). Large peat-forming bogs and lake systems developed in areas of lower sedimentation rates (Lindsay, 2000; Diessel, 2012). This work targets an interval from the base of the Black Alley Shale and Mantuan Productus Beds, which represent the last Permian marine incursion, through to the top of the Bandanna Formation, which broadly marks the end of the Permian and the extinction of the Glossopteris-flora (Fig. 4.2). The boreholes target for sampling are located in the north-east part of the Galilee Basin, the Springsure Shelf in both the eastern and western sides of the Nebine Ridge, and the Roma Shelf in the Bowen Basin (Fig. 4.1).

Figure 4.1: Map of the Galilee and Bowen basins showing the placement of the studied wells.

52

4.3 Material and Methods

4.3.1 Core Logging, Sampling and Lithofacies Assignment

In total, 102 samples were collected from five drill core sites, namely Montani 1 (15), Glue Pot Creek 1 (19), Tambo 1-1A (32), Springsure 19 (22), and Taringa 7 (14) (Fig. 4.1). Core logs for Montani 1, Glue Pot Creek 1 and Tambo 1-1A were adapted from Phillips et al. (2017a) while cores from Springsure 19 and Taringa 7 were examined and logged by the first author. Montani 1 represents a more proximal alluvial plain, while Glue Pot Creek represents the upper delta plain (Fig. 4.2). Tambo 1-1A, Springsure 19 and Taringa 7 represent more distal environments and feature facies ranging from the lower delta plain to a shallow-marine environment (Fig. 4.2). Sampling was focused on fine- grained organic-rich rocks (siltstones and mudstones), which would have the highest chance of yielding palynomorphs. Lithofacies assignments were made using the scheme applied by Phillips et al. (2017a), but modified to separate a subaqueous delta facies (prodelta, delta slope, interdistributary bay) and a subaerial delta facies (distributary channels, tidal flats).

Figure 4.2: Stratigraphy of the Bowen and Galilee basins and placement of the studied wells (modified from Phillips et al., 2017b).

4.3.2 Palynology

Samples from Tambo 1-1A, Isaac Plains and Taringa 7 were sent to MGPalaeo, a commercial palynological laboratory, for standard palynological processing. HCl and HF were added to 15 to 23 g of crushed sample to remove silicate and carbonate minerals. Samples then underwent a lithium

53 heteropolytungstate (s.g. 2.1) float-sink to remove any further minerals. The organic residue was then sieved through a 10 µm polycarbonate filter to remove very fine organic material. HNO3 was used on sample splits to “clean up” samples with a high abundance of plant debris. Samples from Montani 1, Glue Pot Creek 1 and Springsure 19 were processed at the University of Queensland using an acid- free treatment. This technique uses sodium hexametaphosphate (Calgon) to deflocculate the clays, which were then sieved through an 8 μm mesh. Density separation was then achieved using sodium polytungstate (s.g. 1.9) to remove the remaining minerals. Samples were then mounted on slides using Eukitt, a commercial resin-based mounting medium. Both techniques gave relatively good yields (refer to chapter 3). Palynological data from samples from Tambo 1-1A (TAMP-1 to TAMP-22) were already presented in Wheeler et al. (2020) (refer to chapter 5). In this work we present additional data (TAMP-23 to TAMP-30) to augment the previous dataset. Microscopy of the palynological slides was conducted by the senior author using a Zeiss Photomicroscope III and a Leica MC190HD camera. Palynological counts were made to 200 specimens and plotted in the StratabugsTM software package.

4.4 Results

4.4.1 Lithofacies

Mantuan Productus Beds and Peawaddy Formation

The Peawaddy Formation was observed in Tambo 1-1A and Springsure 19 (Fig. 4.3). The upper part of the Peawaddy Formation generally comprises of alternating bands of fine-grained sandstone and mudstone drapes. Bedforms vary from horizontal and low angle lamination to wavy and flaser bedding with minor bioturbation. In both boreholes, the sequence transitions into a fine-grained lithology at the base of the Mantuan Productus Beds. Tambo 1-1A features some conglomeratic intervals which are not present in Springsure 19. The Mantuan Productus Beds unit is generally recognised by the presence of brachiopod shells and was observed in Tambo 1-1A, Springsure 19 and Taringa 7 (Fig. 4.3). This unit varies in thickness and the type of matrix in different parts of the Bowen and Galilee basins. In Tambo 1-1A, this unit is 30 cm thick with the brachiopod shells occurring in a mudstone matrix. Springsure 19 is 1 m thick and the matrix is composed of very fine to medium- grained sandstone. In Taringa 7, the Mantuan Productus Beds is up to 63.5 m thick and features fragmented brachiopods distributed throughout an organic-rich mudstone matrix.

54

‘Fair Hill Formation’ equivalent and Black Alley Shale

The ‘Fair Hill Formation’ equivalent (informal term proposed by Phillips et al., 2017b) (Fig. 4.2) is only observed in Montani 1 and Glue Pot Creek 1 (Fig. 4.3). The upper section (814.53 m to 829.41 m) is dominated by coal. This coal is relatively dull and interbedded with frequent tuffs and numerous mudstone and siltstone partings. The Black Alley Shale generally consists of organic-rich mudstones and shales with frequent tuffaceous intervals. In Tambo 1-1A, Springsure 19 and Taringa 7, the base of the Black Alley Shale tends to be organic-rich and relatively massive with little bedding or bioturbation. In Glue Pot Creek 1, the Black Alley Shale is mainly made up of interbedded fine- grained sandstones and mudstones and is highly bioturbated. An upwards coarsening sandstone unit is present in the Black Alley Shale in Tambo 1-1A at 815.6 m to 821.2 m and at Taringa 7 at 1044.12 m to 1058.85 m (Fig. 4.3).

‘Burngrove Formation’ equivalent and Bandanna Formation

The ‘Burngrove Formation’ equivalent (informal term proposed by Phillips et al., 2017b) (Fig. 4.2) is only observed in Montani 1, Glue Pot Creek 1 and Tambo 1-1A (Fig. 4.3). In Montani 1 it is dominated by coal with frequent tuffaceous interbeds. In Glue Pot Creek 1 it is made up of fine to medium-grained sandstones from the base up to 665 m, interbedded mudstones and sandstones with low-angle to horizontal lamination with abundant siderite nodules and a thin tuffaceous coal seam at the top of the formation. In Tambo 1-1A it mainly consists of coarsening-upward sandstone sequences topped by a highly bioturbated organic-rich mudstone interval interbedded with thin sandstones and occasional conglomerate horizons. The Bandanna Formation is highly variable. In Montani 1 it is dominated by mudstones and siltstones containing abundant plant material and small fining-up sandstones. In Glue Pot Creek, it features large fining-up sandstone sequences and is topped by a coal seam. In Tambo 1-1A, it features interbedded mudstones and sandstones with occasional conglomerate horizons. Overlying these are small fining-upwards sandstone sequences and the top of the formation is marked by a carbonaceous organic-mudstone known as the marker mudstone. In Springsure 19, the lower part (401.18 m to 445.28 m) is mudstone-dominated featuring wavy and lenticular lamination and occasionally sideritic horizons and nodules. Above this are thin coal seams separated by sandstones and mudstones with wavy lamination and with sideritic nodules. Root penetration and plant fragments were also observed. Tuffs are generally absent in the Bandanna Formation except in Taringa 7, which contains thick tuffaceous units (up to 2 m) between the coal seams (Fig. 4.3).

55

Figure 4.3: Correlation of the lithological logs of the studied boreholes, with Montani 1 representing the most proximal section and Taringa 7 representing the most distal section. Locations of palynological samples are marked in red with the corresponding sample name. Shell symbol indicates the presence of brachiopods. BAS – Black Alley Shale; ‘BF’e – ‘Burngrove Formation’ equivalent; ‘FHF’e – ‘Fair Hill Formation’ equivalent; MPB – Mantuan Productus Beds.

56

4.4.2 Palynology

Galilee Basin – Montani 1

Of the 15 samples collected in Montani 1, only eight yielded a high enough abundance of palynomorphs that counts of 200 specimens could be completed (Fig. 4.4). Sample MON-5 did have a small proportion of poorly preserved palynomorphs. Dulhuntyispora parvithola, the index taxon for the APP5 zone, is present from MON-14 to MON-10 and is absent upsection. The index taxon for APP5005 subzone (Price, 1997) is only present in MON-14, and the index species for APP5006, Lycopodiumsporites “crassus” is absent. MON-14 was counted up to 500 specimens due to the overwhelming abundance of palynomorphs which would not be adequately covered with a count of less particles. This sample is dominated by spores, especially Leiotriletes directus, Microbaculispora trisina and Dulhuntyispora parvithola. Pollen grains are relatively low in abundance except for Weylandites lucifer. MON-10 features the highest abundance of Calamospora and Brevitriletes. MON-6 to MON-8 feature a relatively high proportion of striate bisaccate pollen grains, particularly Protohaploxypinus and Striatopodocarpites. The spores are mainly represented by Horriditriletes ramosus. Species diversity decreases continually upsection.

Botryococcus is present throughout the sequence but relatively rare. It does, however peak in MON- 11 above a coal seam and at the base of a coarsening-up sequence. Tetraporina is present in MON- 14 and MON-12 in rare abundances, while Quadrisporites horridus is only present in MON-12. Pilasorites, Peltacystia, Brazilea, Maculatasporites, Cymatiosphaera gondwanensis and Mehlisphaeridium fibratum are present but very rare in the Black Alley Shale and “Fair Hill” equivalent but are absent in the Bandanna Formation.

57

Figure 4.4: Palynostratigraphic data for Montani 1 showing observed genera as a relative abundance (%). Barren samples are marked by a square symbol. For raw counts see Appendix B4.1.

58

Galilee Basin – Glue Pot Creek 1

Palynological samples from Glue Pot Creek 1 display a relative increase in bisaccate pollen grains, particularly Protohaploxypinus limpidus, up section from the base of the Black Alley Shale to the top of the Bandanna Formation (Fig. 4.5). The index species Dulhuntyispora parvithola is only present in GPC-17 in the Black Alley Shale. Microreticulatisporites bitriangularis is rare but present in the Bandanna Formation and Black Alley Shale. Pollen and spore diversity tends to decrease upsection, but there is a relative increase in the abundance of certain spore taxa associated with osmundacean ferns such as Horriditriletes, Cyclogranisporites and Osmundacidites. Spore taxa associated with horsetails such as Calamospora are relatively low in abundance with only a slight increase at the base of the A seam.

Botryococcus is the dominant species of aquatic palynomorphs in Glue Pot Creek 1 (Fig. 4.5). It reaches its greatest abundance in the Black Alley Shale (43.5%) and appears consistently throughout the section. Other lacustrine to brackish taxa are present in much lower abundances, with the highest diversity above the C seam at the base of the Black Alley Shale. Acanthomorph acritarchs, particularly Micrhystridium evansii are absent in these samples.

59

Figure 4.5: Palynostratigraphic data for Glue Pot Creek 1 showing observed genera as a relative abundance (%). Barren samples are marked by a square symbol. For raw counts see Appendix B4.2.

60

Galilee Basin – Tambo 1-1A

The palynological assemblage present in Tambo 1-1A is a well-preserved, typical late Permian APP5005 assemblage transitioning into an APP6 assemblage at the base of the Rewan Formation (Fig. 4.6). At the base of the Black Alley Shale, above the Mantuan Productus Beds, the palynological assemblage is dominated by Leiotriletes and Protohaploxypinus. Other significant components of the assemblage are Microbaculispora and Horriditriletes. The index taxon for APP5005, Microreticulatisporites bitriangularis, is present in rare abundances throughout the section up to the top of the marker mudstone. TAM-28 features a particularly high abundance of the APP5 index taxon Dulhuntyispora parvithola. It is a rare component in samples TAM25 and TAM24 and is absent above the Black Alley Shale. TAM24 and TAM25 also have high abundances of striate bisaccate pollen grains. Samples from the “Burngrove Formation” equivalent feature a striate-bisaccate pollen- dominated assemblage. Non-striate bisaccate pollen grains (Alisporites) and monosulcate pollen grains (Marsupipollenites) also make up a large component of this assemblage. In the marker mudstone itself, the assemblage features slightly higher proportions of trilete spores (Leiotriletes, Horriditriletes, Microbaculispora) than seen in the samples below.

TAM-31 and TAM-32 are dominated by the acanthomorph acritarch Micrhystridium evansii (Fig. 4.6). Both samples were taken from the Mantuan Productus Beds which are comprised of brachiopod shells in a sandy or muddy matrix. In the overlying samples the algal assemblage becomes more diverse featuring several species of acanthomorph acritarchs as well as a variety of algae including Brazilea, Peltacystia, Pilasporites and Leiosphaeridia. Botryococcus only appears along with Cymatiosphaera gondwanensis in the overlying ‘Burngrove Formation’ equivalent. Micrhystridium reappears within the marker mudstone and overlying Rewan Formation alongside Botryococcus.

61

Figure 4.6: Palynostratigraphic data for Tambo 1-1A showing observed genera as a relative abundance (%). Barren samples are marked by a square symbol. For raw counts see Appendix B4.3.

62

Bowen Basin – Springsure 19

Dulhuntyispora parvithola is present in minor amounts throughout the studied section in Springsure 19 with a significant abundance spike near the base of the Black Alley Shale (39.5%) (Fig. 4.7). Protohaploxypinus (particularly P. limpidus) is abundant throughout the section. Springsure 19 also marks the only section studied which features Lycopodiumsporites crassus, index species for APP5006, with SPR-11 featuring up to 11%. Spr-12 features unusual abundances of Verrucosisporites trisecatus, Baculatisporites comaumensis, Cyclogranisporites gondwanensis and Brevitriletes levis.

Springsure 19 features a varied and significant proportion of Botryococcus throughout the section (Fig. 4.7). The most significant proportion of Botryococcus occurs at the base of the Black Alley Shale (29.5%). Other algae (Peltacystia, Pilasporites, Brazilea) occur in low abundances in the Black Alley Shale and in the upper part of the Bandanna Formation, along with sporadic prasinophytes (Leiosphaeridia, Mehlisphaeridium). Micrhystridium is only present in the Mantuan Productus Beds and Micrhystridium evansii was not observed.

63

Figure 4.7: Palynostratigraphic data for Springsure 19 showing observed genera as a relative abundance (%). Barren samples are marked by a square symbol. For raw counts see Appendix B4.5.

64

Bowen Basin – Taringa 7

The Mantuan Productus Beds and the basal section of the Black Alley Shale is together dominated by a diverse assemblage of spores, particularly Dulhuntyispora parvithola, Didecitriletes, Leiotriletes, Microbaculispora and Horriditriletes (Fig. 4.8). Leiotriletes and Microbaculispora have their highest abundance in TAR-10. Upsection there is a gradual increase in Protohaploxypinus, Striatopodocarpites, Scheuringipollenites and Alisporites. TAR-5, which was sampled directly beneath a coal seam, features a very low abundance of pollen grains and a high abundance of spores (particularly Leiotriletes), though some pollen taxa, such as Triadispora and Marsupipollenites are also rare but consistent components of the assemblage. Microreticulatisporites bitriangularis was only observed in TAR-4 and TAR-5.

TAR 12-14 feature a relatively high abundance of Micrhystridium, particularly TAR-12, which sits beneath a large sandstone body (Fig. 4.8). Above this sandstone, in TAR-11, Micrhystridium is absent and Botryococcus appears in a relatively high abundance (12.5%). Other algal palynomorphs like Brazilea are present in the Mantuan Productus Beds and lower Black Alley Shale, and the Bandanna Formation.

65

Figure 4.8: Palynostratigraphic data for Taringa 7 showing observed genera as a relative abundance (%). Barren samples are marked by a square symbol. For raw counts see Appendix B4.6.

66

4.4.3 Palaeofloral composition

67

Figure 4.9 (previous page): Relative proportions of different components of the late Permian flora for Tambo 1-1A, Springsure 19 and Taringa 7, plotted against the lithological signature of each hole. These represent the potential floral compositions in a progradational lower delta setting from the Mantuan Productus Beds to the Bandanna Formation.

Tambo 1-1A, Springsure 19 and Taringa 7

Tambo 1-1A and Taringa 7 feature high proportions of acritarchs at the base of the studied sections (Fig. 4.9) associated with the Mantuan Productus Beds and the marine transgression. Springsure 19 on the other hand only feature a very small proportion of acritarchs in SPR-21 but has much higher proportions of algae (mainly Zygnematacean) throughout the section.

Tambo 1-1A, Springsure 19 and Taringa 7 all show a strong signal of an increasing proportion of gymnosperms upsection. In Tambo 1-1A, Springsure 19 and Taringa 7. The pteridophyte flora is dominated by ferns, however Taringa 7 features a much higher proportion of ferns attributable to Botryopteridales whereas Tambo 1-1A and Springsure 19 feature higher proportions of Osmundales (Fig. 4.9).

Horsetails and lycopods are rare but consistent components of all three sections, with horsetail spores generally having a higher proportion. In Tambo 1-1A, samples TAMP-30 to TAMP-24 feature slightly higher abundances of horsetails spores than other sections, however horsetail spores are absent in TAM-22 to TAMP-16 (Fig. 4.9). The undifferentiated pteridophytes are broadly dominated by Cyclogranisporites and Calamospora, however the higher proportion of this category in SPR-11 is due to high abundances of Lycopodiumsporites, and in SPR-12 by unusually high abundances of Verrucosisporites.

In Springsure 19 and Taringa 7, the gymnosperms mainly consist of Glossopteris, whereas Tambo 1- 1A displays a more diverse assemblage with an upsection increase in the undifferentiated gymnosperms. Conifers appear consistently in TAM-32 to TAM-24, whereas the cordiataleans and peltasperms appear more consistently from TAM-23 to TAM-9.

68

Figure 4.10: Relative proportionsof different components of the late Permian flora for Montani 1 and Glue Pot Creek 1, plotted against lithological signature of each hole. These represent potential floral compositions in the alluvial plain and upper delta plain respectively.

69

Montani 1 and Glue Pot Creek 1

Freshwater algae are present in varying abundances in both Montani 1 and Glue Pot Creek 1, with GPC-9, GPC-11 and particularly GPC-17 having a high abundance of Botryococcus (Fig. 4.10). Acanthomorph acritarchs are not present in either section.

Montani 1 features relatively high abundances of ferns with many taxa assignable to Osmundales. Ferns are dominant in sample MON-14, and lycopods and horsetails are also present in slightly higher abundances than other samples (Fig. 4.10). Glue Pot Creek on the other hand displays relatively low abundances of pteridophyte spores. Ferns are still the dominant pteridophyte group, however, horsetails spores are generally low, while lycopod spores are slightly higher in abundance in GPC-9 and GPC-11. The high abundance of undifferentiated pteridophytes in GPC-7 is mainly attributed to particularly high abundances of Calamospora (14.5%).

All samples in Montani 1 and Glue Pot Creek 1 apart from MON-14 and GPC-17 feature a high proportion (20-50%) of glossopterid pollen as well as moderate proportions (10-20%) of the undifferentiated gymnosperms (Fig. 4.10). Conifer pollen is rare in Glue Pot Creek 1 but consistent in all samples in Montani 1. Peltasperm pollen is present in rare abundances in some Montani-1 samples (MON -14, MON -7, MON -6) with cordaitaleans being very rare or absent. In Glue Pot Creek 1, cordaitalean pollen is higher in abundance relative to peltasperm pollen in samples GPC-17, GPC-13, and GPC-11, whereas peltaseperm pollen is more dominant in samples GPC-9, GPC-7 and GPC-6. The undifferentiated gymnosperms in both Montani 1 and Glue Pot Creek 1 mainly consist of the pollen taxa Alisporites and Scheuringipollenites.

4.5 Discussion

4.5.1 Floral Interpretations of Microflora Assemblages

Examination of vegetation communities using the palynological fossil record as a proxy can be influenced by a number of physical factors that need to be accounted for when utilising this proxy. Some factors that control the expression of microfloral data include pollen/spore durability and preservation potential, the absolute abundance of the living parent plants, dispersal and transportation distances and methods, absolute abundance of palynomorphs released from the parent plant, climate and depositional environment (Tyson, 1995; Slater & Wellman, 2015). Nevertheless, broad interpretations can still be valid and useful especially if linked with detailed sedimentological data which can give insight into palaeoenvironmental conditions such as transport, reworking, tidal influence and marine incursions.

70

The Sporomorph EcoGroup concept (Abbink, 1998) has been applied to reconstruct the palaeoclimate and environment of Mesozoic sections using palynological data based on botanical affinities and habitat preferences inferred from macrofloral remains (Abbink, 2004; Ruckwied et al., 2008). While Mesozoic palynological datasets can often be directly linked to parent plants, plotting the floral composition of each sample relative to the lithostratrigraphy offers insights into the palaeoenvironmental controls on the in-situ palaeoflora.

4.5.2 Palaeoenvironmental and Palaeofloral Reconstruction

The thick coals and fluvio-lacustrine sediments place Montani 1 in an alluvial plain setting for most of the late Permian, with only minor deltaic influence in the Black Alley Shale (Phillips et al., 2017a). Glue Pot Creek 1 also features little to no marine influence and thick sandstones and coals that represent an upper delta plain setting. Strata in Tambo 1-1A, Springsure 19, and Taringa 7 all represent a lower delta plain setting, however; Springsure 19 yielded few acanthomorph acritarchs and did not feature the M. evansii acritarch acme event. While the event itself may have been missed by the sampling, the significantly higher proportions of Botryococcus throughout Springsure 19 may also suggest a lower salinity, as well as calmer waters protected from wave action (Guy-Ohlson, 1992; Demetrescu, 1998). This might be indicative of deposition in a bay or lagoon that would have some freshwater input, while the high abundances of Micrhystridium in Tambo 1-1A and Taringa 7 might suggest an open water environment that is progressively prograded by deltaic sediments.

The palynomorphs identified in the late Permian assemblages in the Bowen and Galilee basins are typical of the Glossopteris flora of Gondwana and consistent with other palynological studies from eastern Australia (Helby, 1973; Price, 1997; Smith & Mantle, 2013). The floras represented by these palynomorphs are made up of a relatively small number of components (mainly glossopterids, ferns and sphenophytes, but also rare conifers, lycophytes, cycadophytes and corditaleans) (Archangelsky, 1990; Cuneo, 1996; Willis & McElwain, 2002; Bernardi et al., 2017). Glossopterid pollen is extremely abundant in nearly all formations and facies. The highest abundance of glossopterid pollen can be found in the delta plain facies of the ‘Burngrove Formation’ equivalent and the lower Bandanna Formation. A diverse array of fern spores likely representing the Osmundales, Botryopteridales and Marattiales can be as abundant as or more abundant than Glossopteris pollen. The highest abundance of fern spores is found in the lower Black Alley Shale within the restricted marine and sub-aqueous delta facies. These likely represent paralic fern communities which are more resistant to salinity and water-level changes. Dulhuntyispora parvithola is particularly abundant in these facies and appears only in very low abundance or not at all upsection. This suggests a paralic

71 affinity for the parent plant and would explain the strong facies control on its spores that has hindered its use as a biostratigraphic index taxon. Sphenophyta are consistent components of the deltaic- lacustrine facies in Glue Pot Creek 1 and mire-lacustrine facies of Montani 1 and likely represent reed-marsh communities on the margins of lakes and mires. They are also consistently found in Springsure 19 and may represent the resistant broadleaf communities on the margins of interdistributary bays as discussed in McLoughlin (1993). Lycopods are a minor but consistent component in all studied sections but only really increase in abundance and diversity in the aftermath of the Permian-Triassic extinction (Retallack, 1995). The index taxon Lycopodiumsporites “crassus” potentially has a bryophyte affinity but this is uncertain (Balme, 1995), and might be constrained to the Springsure Shelf, Denison Trough and Comet Ridge (Laurie et al., 2016). Their occurrence in Springsure 19 and absence in Tambo 1-1A and Taringa 7 may suggest a preference for freshwater habitats. The distribution of different components of the vegetation relative to the lithostratigraphy is discussed below along with a comparison to macrofloral datasets.

Glossopteris

Glossopterids were the dominant form of arborescent vegetation in Permian Gondwana, occupying a variety of environments (Fig. 4.12) (Shi et al., 2010). Antarctic localities have revealed high-density forests composed of mainly smaller, immature trees in frequently disturbed distal floodplain and lacustrine margin facies. Lower density mature forests with larger trees are found in stable bars in braided river systems, swamps and forest bogs (Knepprath 2006; Gulbranson et al., 2012). Examinations of stump density and growth rings from Glossopteris remains in a floodplain environment in Antarctica suggests that these immature Glossopteris forests could grow at quite high densities (potentially up to 2000 individuals per hectare and that temperature was not a limiting factor as the Late Permian high latitude climate was slightly warmer than that experienced by the boreal forests of the modern day (Taylor et al., 1992). These forests are highly productive in terms of biomass and individual trees could reach heights of over 19 m (Miller et al., 2016). Taylor (1994) proposes that the high-density Glossopteris forests occupying the floodplains in high latitude regions would have not supported an understory flora (Fig. 4.11). This would explain the low diversity of macrofloral remains recovered from these environments though it must be considered that could play a significant role too. Within the peat-forming mire environment, the understory vegetation that co-occurs with Glossopteris appears to be more diverse (Fig. 4.11) (McLoughlin et al., 2019).

72

Ferns

Ferns appears more dominant in the more distal facies near the base of the Black Alley Shale but is ubiquitous in all sections, except within the distributary channel facies in the ‘Burngrove Formation’ equivalent of Glue Pot Creek 1. This implies they are tolerant of a wide array of habitats and palaeoenvironmental conditions, acting as pioneer floras in the lower delta plain, as well as understory vegetation in coal swamps or on lake margins (Fig. 4.11).

Fern macrofloras are relatively poorly preserved in Permian Gondwanan sections and do not reflect the abundance and diversity present in the microfloral assemblages (Shi et al., 2010). Macrofloral remains recovered from eastern Australia are mainly associated with Osmundales (Neomariopteris, Dichotomopteris, Palaeosmunda) (McLoughlin, 1992; McLoughlin et al., 2019). This is well reflected in the palynological dataset (Figs. 4.9 & 4.10) as these are the dominant fern group in all sections apart from Taringa 7, where it appears that botryopterid ferns may be more dominant. Botryopterid ferns are rare in Permian Gondwana, but macrofloral remains have been examined from Brazil (Rößler & Galtier, 2003). These appear to grow as epiphytes and has been associated with the marattialean fern Psaronius (DiMichele & Phillips, 2002). Marattialean ferns have been recorded in the Permian of (Lundgren et al., 2001), but presently no macrofloral remains have been recovered from Permian strata in Australia. Marattiales does become established in Australia in the Triassic (Webb, 2001), which fits with the data presented herein, with the spore taxon Thymospora, first appearing in the TAM-9 of Tambo 1-1A directly underlying the P-T Boundary. Thus, the putative epiphytic botryopterids may be using as a physical support, either arborescent Osmundales, or Glossopteris.

Horsetails and lycopsids Lycopsid spores are a rare but consistent component of the sections studied herein. The highest abundances of these spores are usually found in mudstone partings in coal sequences of Montani 1/1A, Glue Pot Creek 1 and Taringa 7 but they are also a small component of the deltaic assemblages. Though arborescent forms are present, lycopsid macrofloras in eastern Gondwana are generally herbaceous (Isoetales, Selaginellales, Lycopodiales) (McLoughlin et al., 2015). Permian lycopsids represent a hygrophilous community presumably occupying a lacustrine margin/paludal environment but they might also represent a component of highly resistant pioneer flora in disturbed and flooded areas and were an understory component of peat-forming floras (Beeston, 1990; Retallack, 1997; McLoughlin et al., 2015; McLoughlin et al., 2019).

73

Horsetails in Permian Gondwana are generally quite diverse and macrofloral remains have been found across the continent. In Australia, horsetails occur in both broad-leafed morphotypes (Trizygia, Schizoneura, Raniganjia) and reed-like forms (Paracalamites, Phyllotheca) (McLoughlin, 1992; McLoughlin, 1993). Macrofloral remains of reed-like sphenophytes can be well preserved in the late Permian coal measures of Australia and extensive mats of sphenophyte axes can be found in a variety of environments, particularly lacustrine and paludal facies where they form stands in fens and marshes (Diessel, 1982; McLoughlin, 1993; van de Wetering et al., 2013b). Remains of Phyllotheca and Equisetum have been recovered alongside Glossopteris in Antarctica, with the palaeoecological interpretation being that the horsetails grew as reed-like stands on the margins of ponds or lakes (Gee, 1989). This is in agreement with macrofloral remains of horsetails recovered in Australia (McLoughlin, 1993).

Microfloral remains (like those presented herein) suggest horsetails are a consistent but minor component of the total assemblage but this is likely an underestimation of their parent plant’s abundance. Horsetail spores are likely overcrowded by the abundance and variety of glossopterid pollen and fern spores. Spores commonly produced by horsetails (i.e. Calamospora and Laevigatosporites) are also produced by other plant groups (Balme, 1995), this further exacerbates the underrepresentation of horsetails in the palynological record. Their also may have a role in this. Horsetails are able to propagate vegetatively and both modern and ancient forms exhibit a regenerative capability which makes them resistant to environmental disturbance (i.e. flooding, tidal influence, wave action) (Gastaldo, 1992).

Non-glossopterid gymnosperms (conifers, cordaitaleans, peltasperms and cycads)

The non-Glossopterid gymnosperms that occupied eastern Australia during the late Permian are a mix of conifers, cordaitaleans, cycads and potentially peltasperms and Ginkgoales. These groups are generally xerophytic and are interpreted to represent a woodland flora that occupies an upland setting, better drained areas on the floodplain and the margins of coal swamps (Fig. 4.11). Upland coniferous forests are known from the margins of the Sydney Basin (Retallack, 1980; Rigby, 1993). Pollen from these groups has high transport potential, which can explain their presence in more distal areas, though conifers like Walkomiella have an association with a forest swamp setting, which might suggest a wider environmental tolerance (Diessel, 1982). This might explain the consistent occurrence of conifer pollen in the Black Alley Shale in Tambo 1-1A, where a mix of conifers and Glossopteris might occupy more well-drained areas on the coastal plain. This environmental preference appears to be recorded In Tambo 1-1A and Taringa 7, though it must be considered that

74 pollen transport from the hinterland may create this signiature. The anomalously high proportion of conifer pollen in SPR-21 would be suggestive of this.

Cordaitaleans (Cordaites and Noeggerathiopsis) are well-recorded components of the Permian assemblage across Gondwana (McLoughlin et al., 2019). Cordaitaleans of the northern hemisphere during Carboniferous were well adapted for different environments, occupying upland regions, as well as being part of the coal-forming flora, suggesting that some cordaite species may have been hydrophilic (Falcon-lang & Bashforth, 2005; Raymond et al., 2009). Pollen transport would explain the distribution of cordaite pollen in the alluvial plain but their rare but consistent occurrence in Tambo 1-1A may be due to the presence of cordaitalean trees growing on the coastal plain.

Definitively establishing the presence of peltasperms either in the basin or the upland is not possible as macrofloral remains of peltasperms have as yet not been recovered from the Permian strata of eastern Australia. Thus far, the only recorded macrofloral peltasperm remains recovered from Gondwana were found in the Cisuralian strata of India and were thought to represent a temporarily warm and dry climatic period (Srivastava et al., 2011). Prevec et al., (2009) notes that pollen grains like Alisporites and Lunatisporites, may have been produced by peltasperms in extrabasinal or non- depositional areas, where macrofossils are unlikely to be preserved. The consistent occurrence of pollen designated a peltasperm affinity in the upper part of the Tambo 1-1A section and Montani 1 may reflect this. The pollen taxa Vitreisporites and Vittatina are common elements of palynological assemblages in Gondwana, and thus is must also be considered that these taxa may have an affinity with more common Permian plants that is yet to be discovered.

Cycad remains have been recorded in late Permian basins of eastern Australia (McLoughlin, 1992). While Cycopites pollen can be associated with multiple plant groups, it is highly likely that in the Bowen Basin it can primarily be attributed to the cycads. It is, however, quite rare within the data presented herein, suggesting that cycads occupy are rare within the studied sections, the pollen was transported, or cycads are underrepresented in palynological datasets due to the high abundance of other pollen and spore taxa (i.e., from Glossopteris or ferns).

75

Figure 4.11: Block diagram of the distribution of different components of the late Permian flora in the Bowen and Galilee basins. Glossopteris is relatively ubiquitous forming dense boreal forests on the floodplain and cohabitating with a fern, lycopsid and sphenophyte understory in peat-forming mires. The more proximal alluvial and upper delta plains feature xerophytic woodlands dominated by gymnosperms and a a hydrophytic lacustrine/paludal/coal-forming flora. Horsetails form reed-like stands in fens and marshes. The coastal plain features a coastal forest in more well-drained areas, whereas areas prone to flooding and incursion feature a coastal pioneer flora. Interdistributary bays feature a mix of prasinophytes and brackish-tolerant algae, while the open water features acanthomorph acritarchs.

4.5.3 Algae and Acritarchs

The occurrence of the Micrhystridium evansii acme event in the southern Galilee Basin is consistent with Norvick (1981) who used the event to define the extent of marine incursion within the basin. The basal section of the Black Alley Shale in both Tambo 1-1A and Taringa 7 features common to abundant Micrhystridium below a medium to coarse-grained sandstone. Above this sandstone

76

Michrystridium is absent and Botryococcus becomes the dominant aquatic palynomorph. Botryococcus has a wide range of environmental tolerances but is generally found in freshwater water bodies as well as brackish settings like lagoons (Guy-Ohlson & Lindström, 1994). This suggests a switch in the Black Alley Shale salinity from marine to brackish (perhaps freshwater) conditions (McLoughlin, 1990). In Springsure 19, Botryococcus is common to abundant throughout suggesting consistent freshwater-brackish conditions. Algal assemblages in Glue Pot Creek 1 and Montani 1 are consistent with those from the Cooper Basin in lacustrine mudstones occurring on the periphery of mires (Lindsay, 2000). The Zygnematacean (particularly Brazilea) and other algae appear to have a wide range of environmental and salinity tolerances, appearing in freshwater and in brackish to marine settings alongside prasinophytes and acritarchs.

4.6 Conclusions

By plotting floral communities inferred from palynology against the lithofacies in which the samples were collected, we have been able to gain insights into the palaeoecological controls affecting said communities. Furthermore, by examining these samples in a broad proximal-distal context, we are able to create a much higher resolution model of the distribution of certain plants and what controls this. In the distal deltaic-marine facies, there is a clear shift from a fern spore dominated flora to a Glossopteris-dominated flora linked to the progradation of the delta. Base level, tidal influence and disturbance are strong controls at a higher resolution. Many of the Permian plants are adapted for a variety of environments while others, like the parent plant of Dulhuntyispora, appear to be limited to a specific setting. The algal composition also shifts rapidly from marine acritarchs to freshwater- brackish algae in the Black Alley Shale with acanthomorph acritarchs only reappearing in the marker mudstone in the Galilee Basin (Fig. 4.11).

This study still represents a relatively broad examination of palaeoecological patterns in the Permian. Higher-resolution sampling using these techniques may yield a more detailed understanding particularly of plant assemblages and their potential relationship to the depositional environment in which they have grown. More studies integrating micro and macrofloral datasets should also prove useful in identifying the effect of biological, transport and preservational biases on the palynological assemblages.

77

Chapter 5 was published in 2020 in Palaeoworld (29(2): 439-452) as:

5. Palaeoenvironmental changes recorded in the palynology and palynofacies of a late Permian marker mudstone (Galilee Basin, Australia)

Alexander Wheeler a *, Nikola Van de Wetering a, Joan Esterle a, Annette E. Götz b, c

a School of Earth and Environmental Sciences, The University of Queensland, St. Lucia QLD 4072, Australia b School of Earth and Environmental Sciences, University of Portsmouth, Portsmouth, PO1 3QL, United Kingdom c Kazan Federal University, 18 Kremlyovskaya Street, Kazan 420008, Republic of Tatarstan, Russian Federation

* Corresponding author. E-mail address: [email protected]

5.1 Abstract

Reconstructing the terrestrial palaeoenvironment during the end-Permian is made challenging by widespread erosion and ecosystem destruction. High-resolution sampling for palynofacies and palynology in sections that preserve the boundary interval allows for detailed examination of the drastic environmental changes that characterize the Permian–Triassic mass extinction. In the Bowen and Galilee basins in eastern Australia, this environmental perturbation is recorded within a marker mudstone that occurs above the uppermost Permian coal seams. The marker mudstone is used as a stratigraphic reference level at many localities, but has previously only been studied at a single locality in the Bowen Basin. In the present study, borehole Tambo 1-1A drilled in the Galilee Basin was selected to clarify whether this black, organic-rich mudstone marks a marine transgression, and to examine potential indicators of the end-Permian mass extinction. A total of 22 samples were taken from the mudstone unit, and from the over- and underlying strata and processed for palynology, palynofacies, and carbon isotope analysis.

Biostratigraphic data indicate that the marker mudstone itself covers the uppermost part of unit APP5, with the first index taxa of unit APP6 floras occurring in samples less than 80 cm above this interval. This can be correlated with several other localities in the Bowen and Sydney basins where this shift occurs just above the uppermost Permian coal seam. Palynofacies data agree with previous interpretations of a southwards prograding delta that subsides as base level rises to form an extensive waterbody in which the marker mudstone was deposited. A change from translucent phytoclast- dominated to opaque phytoclast-dominated palynofacies within the marker mudstone suggests a shift

78 to more oxic conditions in the water column, while base level begins to fluctuate, or increased terrestrial input from fluvial systems as the hinterland rises. Algal bodies resembling Botryococcus are found in the strata above the marker mudstone, but differ in morphology from the algal bodies found in the deltaic facies below. The presence of acanthomorph acritarchs in the marker mudstone and in the overlying Rewan Formation may indicate marine influence. Forms resembling fungal spores are present, but they do not show a spike as seen in other P-T boundary localities.

The relative position of unit APP6 to the P-T boundary itself remains unclear. APP6 assemblages are dominated by simple acavate trilete and cavate trilete spores, which suggests stressed environment dominated by ferns and lycopods. The presence of degraded phytoclasts towards the top of the marker mudstone may also be used to suggest a mass-extinction interval. They may also be indicative of shifting local palaeoenvironmental changes, an interpretation that is supported by the low magnitude negative excursion of the δ13C isotope values within the marker mudstone. More datasets from the Bowen and Galilee basins will be essential to decoupling these signals.

Keywords: Palynology; Palynofacies; Carbon isotopes; Palaeoenvironment; Permian–Triassic boundary; Galilee Basin

5.2 Introduction

Deciphering patterns and processes of environmental change across the Permian–Triassic boundary in the terrestrial basins of Gondwana has long been a challenge (e.g., de Wit et al., 2002; Gastaldo et al., 2009; Smith & Botha-Brink, 2014). In eastern Australia, integrated palynological and geochronological studies have shown great utility for dating and correlating the late Permian deposits (Smith & Mantle, 2013; Laurie et al., 2016). Potential sections in the Galilee Basin in particular have remained understudied when compared to those of other eastern Australian basins such as the Bowen, Gunnedah and Sydney basins (Fig. 5.1; 5.2). The Galilee Basin is only recently beginning to receive attention for its economic resource potential (Hansen & Uroda, 2018; I’Anson et al., 2018), but it may also contain valuable climatic and environmental records of the late Palaeozoic that have yet to be exploited. Of special importance is that the Galilee Basin was situated relatively far away from the tectonic activity in the New England Orogen and that it captures a terrestrial-marine transition from the north to the south and across the Springsure Shelf, which allows for detailed study of various depositional environments. In many parts of the Bowen and Galilee basins, a regional disconformity to low-angle unconformity at the base of the Rewan Formation represents an erosional contact that probably removed the Permian–Triassic boundary itself within the basins (Brakel et al., 2009; Sliwa

79 et al., 2017). Recent systematic correlation, however, revealed that where this unconformity is not observed, there remains a laterally continuous mudstone interval that is deposited above the last coal deposits of the Permian (Sliwa et al., 2017). This mudstone is recognised as a prominent gamma spike in wireline logs, and has been termed the marker mudstone due to its consistency and utilization for basin-wide correlation. Michaelsen et al. (2000) and Michaelsen (2002) interpreted the marker mudstone as a lake deposit that developed above the peat deposits immediately preceding the P-T boundary; however, the precise nature and origin of this mudstone remain unknown. The aims of this study are: 1) to place the marker mudstone within a biostratigraphic context, 2) to develop a palaeoenvironmental interpretation based on palynology and palynofacies, and 3) to determine if palynological markers (e.g., fungal spike, lycopsid spike) reported from other localities exposing the Permian–Triassic boundary are present.

Figure 5.1: Map of the Galilee and Bowen basins with location of stratigraphically significant wells. GSQ Tambo 1-1A (marked with a star) occupies a position at the edge of the Springsure Shelf, which connects the two basins.

80

Figure 5.2: W-E cross section of the Permo-Triassic sediments infilling the Cooper, Galilee and Bowen basins overlain by the younger sediments of the Eromanga Basin (after Hobday, 1987).

5.3 Geological Setting

The Galilee Basin is an intracratonic basin that covers an area of around 247,000 km2 in central Queensland (Fig. 5.1) (Allen & Fielding, 2007). The sediment infill occurred during two major phases of deposition from the Pennsylvanian (late Carboniferous) to the Cisuralian (early Permian) and from the Lopingian (late Permian) to the Middle Triassic, separated by a mid-Permian depositional hiatus. This hiatus is attributed to reduced rates of deposition rather than to erosion, as there is little evidence of an erosional unconformity (Van Heeswijck, 2010; Phillips et al., 2018a). The basement comprises the Devonian–Carboniferous Drummond Basin in the east, the Thomson Orogen in the centre, and cratonic rocks in the west. Large areas of the basin are overlain by the Jurassic– Cretaceous deposits of the Eromanga Basin (Fig. 5.2). The Galilee and Bowen basins are separated by the Nebine Ridge on a structural high called the Springsure Shelf. The Galilee Basin is separated from the Cooper Basin by the Canaway Ridge in the south-west of the basin.

81

The complex nature of the late Permian deposits has long made lithostratigraphic and sequence stratigraphic correlation a challenge. Allen & Fielding (2007) worked on a sequence stratigraphic correlation of the low-accommodation setting of the Betts Creek Beds relative to the high accommodation setting of the Denison Trough in the Bowen Basin, and identified second (10–100 Ma) and third-order (1–10 Ma) sequences in outcrops and well logs. The lithostratigraphic correlation has been reinterpreted by different authors in different regions within the basin (Fig. 5.3) as summarised by Phillips et al. (2017b). Large-scale correlations across the Galilee Basin recently allowed the identification of the marker mudstone along the eastern margin of the basin and in the borehole GSQ Tambo 1-1A (Phillips et al., 2017a).

Figure 5.3: Lithostratigraphic scheme of the Galilee Basin and Denison Trough (modified from Phillips et al., 2017b). The relative positions of the major coal seams (A-F) are also displayed. Though there is evidence of a regional unconformity in many locations, the interface of the Bandanna and Rewan formations features the marker mudstone, and no unconformity is apparent. The interval being investigated in this study is marked between the two stars.

5.4 Permian–Triassic Boundary

The Permian–Triassic boundary is formally defined at the Global Boundary Stratotype Section and Point (GSSP) in the marine Meishan Beds in China by the first appearance of the conodont Hindeodus parvus. This horizon is located several centimetres above the mass extinction interval (Yin et al., 2001). The extinction is estimated to have affected between 80% and 95% of marine species, including trilobites, fusulinid foraminifera and various groups of echinoderms and brachiopods (Benton & Twitchett, 2003). In terrestrial basins, a number of proxies are useful in placing the

82 boundary. One of the most widely recognised indicators of the boundary is a negative carbon isotopic excursion (Retallack & Krull, 2006). This excursion is thought to be related to an increase in global

CO2 from increased volcanic activity at the end of the Permian (Svensen et al., 2009). In terms of palynological indicators, several authors (Eshet et al., 1995; Visscher et al., 1996; Steiner et al., 2003; Bercovici & Vajda, 2016) have recognised a spike in the abundance of fungal remains such as Reduviasporonites chalastus, though it has been suggested that this species may have an algal affinity (Foster et al., 2002) or even be a result of recent contamination (Hochuli, 2016). Spina et al. (2015) suggest that R. chalastus is a chlorophycean algae with a strong affinity for brackish to hypersaline water, and cannot be used as a consistent temporal marker.

In eastern Australia, the boundary has historically been placed at the top of the last Permian coal deposits (Laurie et al., 2016). In the Bowen Basin, this would correspond to the interface of the Bandanna Formation, Rangal and Baralaba coal measures with the overlying Rewan Group, Within a palynological framework, defining the P-T boundary in Australia has been particularly challenging due to several factors, including the apparently multiple floral turnovers during the transition from the late Permian Glossopterid flora to the Early Triassic flora, widespread erosional surfaces related to the rapidly aridifying climate, and a disconnect between the extinction event, negative carbon isotopic excursion (Retallack and Krull, 2006), and the GSSP-defined boundary. The boundary was initially thought to lie at or near the base of the Kraeuselisporites saeptatus and Lunatisporites pellucidus zones of western and eastern Australia, respectively (Dolby and Balme, 1976), correlative with the base of the informal eastern Australian unit APT1 of Price et al. (1985; see Price, 1997). Recent high-resolution radiometric analyses, however, indicate that the boundary may be situated stratigraphically lower (Laurie et al., 2016), i.e., at or near the base of the Protohaploxypinus microcorpus zone (including the Playfordiaspora crenulata (sub)zone; see Helby et al., 1987) equivalent to unit APP6 of Price et al. (1985; see Price, 1997). The marker mudstone was selected for further scrutiny of the palynostratigraphic data and age assignments as it appears to represent continuous deposition at the interface of the Bandanna and Rewan formations, and is fine-grained and organic-rich, which is ideal for palynological sampling and organic carbon isotope analysis.

5.5 Materials and Methods

5.5.1 Study Area

The borehole GSQ Tambo 1-1A is located on the Springsure Shelf in the south-east of the Galilee Basin (Fig. 5.1) and features a lithostratigraphic succession easily correlative with that of the Bowen

83

Basin and a 1 m thick expression of the marker mudstone with no discernible unconformity at the base of the Rewan Formation.

5.5.2 Palynological Processing and Analysis

Twenty-two (22) samples were taken throughout the 45 m thick sequence for palynological and palynofacies analysis. Intervals sampled included the siltstones above the marker mudstone (TAMP1–TAMP6), the marker mudstone itself (TAMP7–TAMP11) and siltstones and mudstones in the underlying Bandanna Formation and ‘Burngrove Formation’ equivalent (TAMP12–TAMP22). A 20 m thick sandstone interval separates samples TAMP1–TAMP12 and TAMP13–TAMP22.

Acid-processing of palynological samples was conducted by MGPalaeo laboratory in Perth. Standard acid processing techniques were used (Wood et al., 1996). This involved digesting 15–20 g of sample in HCl to remove carbonates. Excess HCl was decanted and HF was added to remove . After

48 hours the excess acid was decanted and neutralised by repeated washings in deionised H2O and centrifuging. The residue then underwent density separation using heavy liquid (density 2.0) and was mounted on a slide. Two slides per sample were prepared using this method, and two more were prepared after oxidation with HNO3 to remove palynodebris.

Analysis and counting of palynological slides was done using a Zeiss Photomicroscope III equipped with a Leica MC190HD camera. For biostratigraphy, counts of 200 palynomorphs per slide were done as per the recommendation of Traverse (2007). Classification of species was based on the nomenclature of Price (1997), Foster (1979, 1982), Backhouse (1991), de Jersey (1979) and Rigby and Hekel (1977). If taxonomic names differ between the authors, the most recently published names are used. Biostratigraphic assessment was based on the scheme of Price et al. (1985) and Price (1997). For palynofacies, counts of a minimum of 300 particles per sample were done. Particles were classified according to a palynofacies scheme based on that of Tyson (1995). Biostratigraphic and palynofacies data were plotted using the TILIA software package and statistically analysed using the CONISS cluster analysis to differentiate and palynofacies assemblages. Count results were also plotted using ternary diagrams based on major palynofacies components (opaque phytoclasts, translucent phytoclasts, and terrestrial palynomorphs) to distinguish between different palynofacies assemblages.

84

5.5.3 Carbon Isotope Processing and Analysis

Seventeen (17) samples of the total twenty-two (22) were crushed to powder for organic carbon isotope analysis at the University of Queensland, Australia. Isotope samples were compared to palynological samples taken at the equivalent depths to identify any co-occurring trends. The δ13C isotope values were determined with a stable-isotope-ratio mass spectrometer (Isoprime), coupled in continuous flow mode with an elemental analyser (Elementar Cube) (EA-CF-IRMS). Calibration was performed by use of two standards, USGS24 (-16.1‰ δ13CPDB) and NAT76H (-29.26‰ δ13CPDB), which were interspersed throughout analytical runs. Each sample was analysed in duplicate, using 50–200 μg of concentrate combusted at 1020°C in 3.5 mm × 5 mm tin capsules. Any sample with a beam size outside the working range of 1 × 10–9 to 9 × 10–9 Å, or with a δ13C result variation between duplicates of > 0.4‰, was re-analysed, in accordance with laboratory quality control practices.

5.6 Results

5.6.1 Palynostratigraphy

Of the twenty-two samples processed for palynology, thirteen yielded assemblages suitable for palynomorph counting (Fig. 5.4). These samples were TAMP1, TAMP3, TAMP4, TAMP9, TAMP10, TAMP11, TAMP16, TAMP17, TAMP18, TAMP19, TAMP20, TAMP21 and TAMP22. Samples TAMP5–TAMP8 yielded a few poorly preserved palynomorphs, though these were generally covered in degraded phytoclasts (Fig. 5.8A) and counts were not possible; the few species that were identifiable in these samples were noted.

Samples TAMP16–TAMP22 all yielded assemblages typical of unit APP5 (Price et al., 1985) characterized by a high abundance and diversity of striate bisaccate pollen grains (Protohaploxypinus spp., Striatopodocarpites spp.) and non-striate bisaccate pollen grains (Alisporites spp., Scheuringipollenites spp., Vitreisporites pallidus). Spinose and ornamented trilete spores (Microbaculispora spp., Horriditriletes spp.) are also common, as are monosulcate pollen grains (Marsupipollenites spp.). Dulhuntyispora parvithola, the index taxon of unit APP5, was not identified within these samples, but Microreticulatisporites bitriangularis (Fig. 5.8E), the key taxon of the APP5005 subunit, is a rare component in many of the samples. Lycopodiumsporites ‘crassus’, the index taxon for the APP5006 subunit was neither detected in this study; hence, the samples were placed within the APP5005 subunit. The marker mudstone samples (samples TAMP9–TAMP11) yielded assemblages closely resembling those of the samples stratigraphically lower in the section, with some notable differences. The abundances of the smooth-walled trilete spore Leiotriletes

85 directus and the ornamented trilete spore Brevitriletes spp. increase, whereas the abundance of the previously dominant pollen species Protohaploxypinus limpidus decreases.

Samples TAMP1, TAMP3 and TAMP4 feature an assemblage distinctly different to those of the marker mudstone. Samples TAMP3 and TAMP4 both contain rare specimens of Playfordiaspora crenulata (Fig. 5.8B), the index taxon for the lower P. microcorpus of Foster (1979). Sample TAMP1 further contains several specimens of Triquitrites proratus and Triplexisporites playfordii, both of which are also indicative of unit APP6. Other significant species featured in samples TAMP3 and TAMP4 include small ornamented trilete spores (Thymospora ipsviciensis, Brevitriletes spp.) and spinose trilete spores (Horriditriletes spp.). The abundance and diversity of bisaccate pollen grains has decreased. Smooth-walled trilete spores (Leiotriletes directus, Calamospora sp., Punctatisporites spp.) are particularly abundant in sample TAMP1.

Figure 5.4 (next page): Palynostratigraphic data showing selected species from Tambo 1-1A based on counts of 200 palynomorphs. Samples in which this count could not be reached are not included. The data indicates unit APP5 (?APP5006 subunit) continues into the marker mudstone. The transition to unit APP6 occurs above the mudstone (marked by the first sparse occurrence of Protohaploxypinus microcorpus). Cluster analysis suggests the assemblages in each unit are relatively distinct from one another even though some elements of unit APP5 are still present above the marker mudstone. For raw counts see Appendix B4.3.

86

87

5.6.2 Palynofacies

The samples taken from below the marker mudstone (TAMP12–TAMP22) feature varied palynofacies assemblages, but distinct trends can be recognized (Fig. 5.5). The proportion of opaque phytoclasts tends to increase up-section, though there is some variance (between 69.2% and 31.3% of the total assemblage). Equidimensional phytoclasts are the most abundant component of the assemblage, but structured and unstructured translucent phytoclasts are also common. Sporomorphs (particularly bisaccate pollen grains) have high relative abundance in TAMP21, TAMP20 and TAMP19, with palynomorphs outnumbering even the phytoclasts. The palynomorph abundance decreases up-section as the sandstone intervals become thicker until the assemblage features little to no palynomorphs preserved. Acanthomorph acritarchs and fungal spores are absent from this interval, but Botryococcus and Cymatiosphaera are present in rare abundances (Fig. 5.9). TAMP12, located slightly below the marker mudstone, features an assemblage dominated by opaque phytoclasts, with very few palynomorphs preserved but the highest proportion of amorphous organic matter (AOM) found in any of the studied samples.

The samples taken from the marker mudstone (TAMP7–TAMP11) show a striking trend with regard to oxidation of the phytoclasts. Sample TAMP11 features an assemblage dominated by translucent phytoclasts. The abundance of translucent phytoclasts decreases up-section whereas the opaque phytoclasts become dominant. AOM is present in low abundance in all samples within the marker mudstone, and degraded phytoclasts have a spike in sample TAMP8. The degraded phytoclasts are an unusual component of the assemblage as they appear to occur in the form of clusters of fragments of opaque and translucent phytoclast debris surrounded by a pseudoamorphous groundmass. Pollen and spores are rare, and indeterminate palynomorphs tend to be more abundant in TAMP7 and TAMP8. The marker mudstone contains rare Botryococcus, as well as acanthomorph acritarchs (Micrhystridium) and fungal remains (Reduviasporonites chalastus). Above the marker mudstone, samples TAMP1, TAMP2, TAMP5 and TAMP6 all have very high abundances of opaque phytoclasts and low abundances of translucent phytoclasts and palynomorphs. Samples TAMP5 and TAMP6 also feature low abundances of degraded phytoclasts. Samples TAMP3 and TAMP4 represent a distinct shift to an assemblage with even abundances of opaque and translucent phytoclasts and with higher abundances of palynomorphs. Botryococcus, acanthomorph acritarchs, and fungal remains are all present in low abundances in these samples as well (Fig. 5.9).

Results of our cluster analysis reveal three distinct palynofacies assemblages (Fig. 5.10). Samples TAMP19–TAMP21 encompass Assemblage A, which is mainly characterised by the high abundance of terrestrial palynomorphs (> 30%) relative to the proportion of phytoclasts. Samples TAMP3, TAMP4, TAMP9, TAMP10, TAMP11, TAMP13, TAMP15, TAMP16 and TAMP17 are classified

88 as part of Assemblage B, which is characterised by a lower proportion of opaque phytoclasts (< 60%) relative to the translucent phytoclasts and lower proportions of terrestrial palynomorphs (< 30%). Assemblage C (samples TAMP1, TAMP2, TAMP5, TAMP6, TAMP7, TAMP8, TAMP12, TAMP14, TAMP18, and TAMP22) is characterised by very high proportions of opaque phytoclasts (> 60%) and relatively low proportions of terrestrial palynomorphs (< 30%).

Figure 5.5: Line chart depicting palynofacies data from Tambo 1-1A. Lines measuring relative abundance (%) also mark sample locations on the lithology. Presence of key palaeoenvironmental indicators (algae/prasinophytes, acritarchs, fungal spores) are also marked. Cluster analysis of the samples suggests three main palynofacies patterns. Ternary diagrams plotting the data based on the major palynofacies components show three distinct palynofacies assemblages (A-C). The major palynofacies components are opaque phytoclasts (OP), translucent phytoclasts (TP) and terrestrial palynomorphs (PAL). For raw palynofacies counts see Appendix B5.1.

89

5.6.3 Carbon Isotopes

The values of organic δ13C within the Tambo 1-1A samples range from -26.2‰ to -23.2‰ (Table 1), typical for Permian-age coals (Korte and Kozur, 2010). At core depths shallower than 723.73 m, the δ13C values deplete significantly to an absolute minimum of -26.2‰ at 700.39 m. From the data it is possible to distinguish two clear data ranges: the upper data range occurring in TAMI1–TAMI7 with an average δ13C of -25.8‰ (σ = 0.3‰), and the lower data range within TAMI8–TAMI17 with an average δ13C of -23.6‰ (σ = 0.3‰). These two data ranges are separated by a maximum ∆ 3.0‰ isotopic excursion. This relatively low magnitude excursion is comparable to other organic carbon isotopic values in Australia during the preliminary stages of CO2 release, prior to the Permian– Triassic boundary (Retallack and Krull, 2006). Carbon isotopic values indicate no observable trends with palynofacies data, indicating that δ13C is not dependent on phytoclast type or assemblage.

Sample Name Depth (m) δ13C (VPDB ‰) Equivalent Palynological Sample TAMI1 699,13 -25,4 TAMP5 TAMI2 699,33 -25,5 TAMP6 TAMI3 700,14 -26,0 TAMP7 TAMI4 700,24 -26,1 TAMP8 TAMI5 700,39 -26,2 TAMP9 TAMI6 700,49 -25,9 TAMP10 TAMI7 700,64 -25,4 TAMP11 TAMI8 723,73 -23,2 TAMP13 TAMI9 729,01 -23,2 TAMP14 TAMI10 729,21 -24,1 TAMP15 TAMI11 733,91 -23,8 TAMP16 TAMI12 735,04 -23,4 TAMP17 TAMI13 735,4 -23,6 TAMP18 TAMI14 738,33 -23,5 TAMP19 TAMI15 738,81 -23,3 TAMP20 TAMI16 740,54 -23,7 TAMP21 TAMI17 741,66 -23,7 TAMP22

Table 2: Organic carbon isotope samples from GSQ Tambo 1-1A and their equivalent palynological samples.

5.7 Discussion

5.7.1 Age Determination – Unit APP6

The onset of unit APP6 or equivalent biostratigraphic units shows some degree of variability across the eastern Australian basins. Michaelsen (2002) mark its appearance at the base of the marker mudstone, while Rigby and Hekel (1977) and de Jersey (1979) mark the stratigraphic position of the Permian–Triassic boundary just above the uppermost coal seam. Difficulties arise due to the

90 distinction of the Playfordiaspora crenulata and Protohaploxypinus microcorpus zones (Foster, 1979). This is due to the appearance of forms resembling P. microcorpus below the level of the first appearance of P. crenulata (this work; de Jersey, 1979; McLoughlin, 1988; Price, 1997). Foster (1979, 1982) proposed a subzone P. microcorpus Zone as defined by Helby (1973) using the first appearance of P. crenulata to define a lower unit. Initially this was described as the Lower P. microcorpus Zone, but later it was amended to become the P. crenulata Oppel zone based on its biostratigraphic and geographic significance (Foster, 1982). Unit APP6 was initially defined by Price (1997), based on the first appearance of Triplexisporites playfordii. However, the subunits of APP601 and APP602 were equated with the P. crenulata and P. microcorpus associations, respectively. Many of the elements present within APP5005 subunit continue to appear within the lower APP6 unit though in much lower abundances (Protohaploxypinus, Scheuringipollenites), along with the appearance of several new key species (e.g., Triquitrites proratus, Triplexisporites playfordii). However, the samples examined in this work as well as the datasets of de Jersey (1979) show that the appearance of key index taxa marking the onset of unit APP6 do not necessarily appear in the same horizon, which suggests strong environmental or sedimentological controls on the base and extent of the APP6 unit.

5.7.2 Permian–Triassic Boundary

Placing the unit APP6 relative to the Permian–Triassic boundary in eastern Australia has challenged workers in the area for many decades. The presence of key P-T indicators such as macrofloral turnover and fungal spikes is inconsistent and so workers have turned to chronostratigraphic techniques. Metcalfe et al. (2015) used high precision U-Pb zircon dates to place the GSSP-defined P-T boundary in the Scarborough Sandstone uncoupling it from the extinction interval, which is located at the base of the Coalcliff Sandstone. This would place the extinction interval at the base of unit APP6 and the GSSP-defined P-T boundary into the lower part of unit APT1. Laurie et al. (2016) place the unit APP6 in the earliest Triassic on the basis of palynological and macrofloral data from the borehole Santos Yebna 1 (Powis, 2009). The base of the unit APP6 was defined at this locality just above the uppermost Permian coal seam by the first appearance of Triplexisporites playfordii (Murdoch, 2012). Several metres above this interval, Powis (2009) recorded non-glossopterid macroflora, which may suggest a post-extinction plant assemblage. U-Pb zircon dates were also used to calibrate the biostratigraphic scheme, placing the P-T boundary at the base of unit APP6, however no ages have so far been obtained from unit APP6 itself to support this interpretation (Laurie et al., 2016). In the Sydney Basin, Retallack (1995) associates the P. microcorpus Zone with the Dicroidium

91 callipteroides megafloral biozone in the Early Triassic and identifies a distinct change from a high diversity glossopterid flora to a low diversity post-extinction flora just above the Bulli Coal.

In Antarctica, the P-T boundary sections are well exposed. Collinson et al. (2006) identify the P. microcorpus Zone from a sample at the base of a carbonaceous mudstone parting in the uppermost Permian coals in the Buckley Formation at Graphite Peak. Vertebraria (glossopterid roots) were recovered several centimetres above the top of the mudstone parting at the base of the uppermost coal seam in the Buckley Formation (Retallack & Krull, 1999), which suggests the extinction interval lies above their sampled interval. The P. microcorpus flora is regarded as transitional between the Glossopteris vegetation and the final extinction event, which eliminates the last elements of the Glossopteris flora. In a section studied in the Prince Charles Mountains, the P-T boundary was placed at the interface of the McKinnon Member and Ritchie beds, based on the cessation of coal deposition. This interval also marks the first appearance of Lunatisporites pellucidus, the index taxon of unit APT1 (Lindström & McLoughlin, 2007). In the central Transantarctic Mountains, Glossopteris remains have been found 37 m above detrital zircon samples indicating an Early Triassic age (Elliot et al., 2017), while in South Africa, Glossopteris remains were recovered from the Assemblage Zone, which is regarded as Triassic in age (Gastaldo et al., 2017). Both these occurrences raise questions about the accuracy of proxies such as a macrofloral turnover when defining the terrestrial expression of the P-T boundary. Even in the marine basins of western Australia, placing the P-T boundary with confidence is a challenge due to unconformities in different sections in each basin and a lack of ashfall tuffs that can be age dated. However, in all basins the carbon isotopes tend to correlate well with the biostratigraphy (Gorter et al., 2009). The carbon isotope excursion begins to shift more negative at the base of the P. microcorpus Zone and reaches its most negative values at the base of the L. pellucidus Zone (Morante, 1996; Gorter et al., 2009), which agrees with the aforementioned Antarctic studies.

Using organic carbon isotopes, the age-proximity of the marker mudstone’s deposition to the P-T boundary may be speculated. From the relatively low magnitude excursion observed in these samples, it is assumed that the marker mudstone was not deposited during a time period to which the full extent of contemporaneous CO2 production occurred (Korte & Kozur, 2010). This interpretation is supported by the lack of observable unit APP6 taxa in the marker mudstone samples, which suggest the P-T boundary itself, and the apogee of the P-T carbon isotopic excursion occurred well after the deposition of the marker mudstone (Fig. 5.6). The use of carbon isotope trends as tool for relative age dating across relatively low-resolution sequences is problematic, given a number of palaeoenvironmental factors that may influence the uptake of carbon isotopes in the primordial plant material (Van de Wetering et al., 2013a). It is these palaeoenvironmental factors that are attributed to

92 the low magnitude (σ = 0.3‰) variation in carbon isotopic values within the upper (TAMI1–TAMI7) and lower (TAMI8–TAMI17) range of isotopic values for the marker mudstone, though these are not causally related to palynofacies changes observed within the equivalent samples. Furthermore, the absolute range and timing of carbon isotopic variation is variable between sample location and stratigraphy, dependent on both depositional setting and geographical proximity to the Siberian Traps and late Permian volcanic systems associated with massive CO2 expulsion (Svensen et al., 2009).

Figure 5.6: Carbon isotope trends with depth in Tambo 1-1A showing a low magnitude excursion within the marker mudstone, plotted alongside the pollen/spore ratio, bioevents, biostratigraphy and palynofacies cluster analysis. The current placement of the P-T boundary in Australia (Laurie et al., 2016) is marked, though evidence from this study is not sufficient to define it conclusively in this locality.

93

5.7.3 Palaeoenvironment

The presence of a variety of freshwater to brackish algae (Botryococcus, Brazilea) and prasinophytes (Cymatiosphaera) in the Bandanna Formation fits well with the previous interpretation of a deltaic environment (Fig. 5.7) (Phillips et al., 2017a). The mudstones and siltstones at around 740 m in core Tambo 1-1A are the distal equivalent of the B seam (Fig. 5,3). This interval falls into palynofacies Assemblage A, an assemblage dominated by palynomorphs, particularly bisaccate pollen grains, which are concentrated in distal areas due to the Neves Effect (Chaloner and Muir, 1968). The section coarsens upwards into a more sandstone-dominated facies as the delta prograded southwards. The palynofacies also appear consistent with this interpretation with a transition from Assemblage A to Assemblage B featuring an increased abundance of phytoclasts, suggesting increased proximity to the terrestrial source with a switch to a fluvial-dominated environment. The phytoclast assemblage is a mixture of opaque and translucent phytoclasts, which reflects varying lengths of transport either from vegetation occupying the delta and adjacent swamps or from fluvial systems carrying phytoclasts over a long distance.

The occurrence of the black organic-rich marker mudstone along with the presence of acritarchs, albeit in low abundances, suggests that this bed may represent a maximum flooding surface. This coincides with a spike in acritarch abundances in the Sydney Basin immediately above the uppermost Permian coals (Bulli Coal), which Retallack (1995) attributed to a short-term marine incursion. The low abundances of freshwater/brackish algae and AOM could reflect either poor preservation or conditions in the water column unsuitable for the proliferation of freshwater algal colonies. The palynofacies assemblage shifts from Assemblage B to Assemblage C, suggesting a change in the oxidation conditions in the water column or a change in the transport dynamics affecting input into the lake. A similar trend can be seen in the transition from palynofacies CPFII to CPFIII in late Permian coal deposits (Van de Wetering et al., 2013b), where an assemblage featuring mixed opaque and translucent phytoclasts and very low AOM abundances is replaced by an assemblage almost completely dominated by opaque phytoclasts. This is related to a base-level change from high to fluctuating, with redox conditions changing from stable to oxic. A high base level with anoxic bottom waters would preserve wood fragments and organic material in the sediments to form the organic- rich mudstone. The source of the degraded phytoclasts remains uncertain. They might derive from eroded peat further inland, or might indicate extreme conditions (i.e., high acidity) within the water column, both of which are phenomena typically associated with end-Permian environmental perturbation (Sephton et al., 2005; 2015).

Detrital-zircon data from Phillips et al. (2018b) indicate that the Anakie Inlier was submerged at the time the marker mudstone was deposited, and that the waterbody was a large-scale feature that spread

94 across both basins (Fig. 5.7). Accommodation for this waterbody would have been created by subsidence related to foreland loading on the eastern margin of the Bowen Basin. Thin sandstone beds sourced from the east occur within coarsening-upwards sequences above the marker mudstone and its regional equivalents (Grech and Dyson, 1997). These sandstones make up the basal unit of the Sagittarius Sandstone and may represent proximal deltaic facies, which progrades westwards (Grech, 2001; Sliwa et al., 2017). The full extent of the marker mudstone is still not well constrained and it is yet to be determined if it represents a single extensive lake or smaller separate bodies of water which infilled the topography during the transgression. This might explain the apparent diachroneity in the palynogical correlation between basins (Sliwa et al., 2017).

The palynofacies above the marker mudstone is generally opaque phytoclast dominated and has a very low abundance of palynomorphs (Assemblage C), which may reflect long-distance transport in fluvial channels associated with the regressive Rewan Formation. An alternative interpretation is that this assemblage could potentially also represent a P-T extinction interval as fungal remains (Reduviasporinites chalastus) are present above and below, though in rare abundances. However, the switch back to Assemblage B only a few centimetres above this interval in samples TAMP3 and TAMP4, which contain a diverse APP6 flora, raises questions about the extent and position of the extinction event and about the P-T boundary placement in eastern Australia.

95

Figure 5.7: Reconstruction of the potential palaeoenvironmental conditions of the Bandanna Formation (A) and the marker mudstone (B). Data from Tambo 1-1A supports the interpretation of a southwards-prograding delta which subsides as base level rises leading to the formation of a large waterbody (lake) across the Bowen and Galilee basins (Phillips et al., 2017a, 2018a). Blue arrows suggest areas where a potential marine transgression could have come in. The presence of both acanthomorph acritarchs and Botryococcus suggests increased salinity but not fully open marine conditions, leading to the interpretation of a marine influenced lacustrine environment.

5.8 Conclusions and Outlook

Palynological and palynofacies data from GSQ Tambo 1-1A provide new insight into the late Permian palaeoenvironment in the Galilee Basin. The data generally agree with previous interpretations of a prograding deltaic system, which then subsided and was flooded with a large waterbody (lake) that occupied the potentially combined centre of the Bowen and Galilee basins due to a submerged Anakie Inlier (Phillips et al., 2017a, 2018a). The change in algal components with rare occurrences of acanthomorph acritarchs suggests an increased salinity level, potentially related to a short-term marine transgression.

Based on palynological data alone, we cannot place the P-T boundary in GSQ Tambo 1-1A with any certainty. Whereas the transition from unit APP5 to unit APP6 correlates well with some coeval sections in the Bowen Basin (Smith & Mantle, 2013) and Sydney Basin (Retallack, 1995), it differs

96 from that of the marker mudstone section at the Newlands Coal Mine (Michaelsen et al., 2000), where the transition to unit APP6 lies at the base of the marker mudstone package. This suggests some degree of diachroneity, with regard to either the base of unit APP6 or the onset of the marker mudstone deposition. Another potential interpretation is that the marker mudstone featured in the Galilee Basin is not a true correlative of the marker mudstone mapped in the Bowen Basin. Carbon isotopic evidence from the marker mudstone concurs with palynological results, indicating the onset of contemporaneous CO2 release at the onset of the P-T carbon isotopic excursion, but not coincident with the apogee of this event.

More detailed studies of the marker mudstone in both the Bowen and Galilee basins are needed to fully understand its relationship to the P-T boundary. Mapping of the marker mudstone in the Galilee Basin is needed to understand its nature and spatial distribution, for comparison with the Bowen Basin. Future studies should also focus not just on palynological data, but on high-resolution carbon isotope and trace element trends to allow more detailed interpretations of the palaeoenvironmental and palaeoclimatic changes occurring in Australia at the end of the Permian.

5.9 Acknowledgements

The authors would like to thank the Vale UQ Coal Geoscience Program for funding this study, and The Palynological Society – AASP for providing additional funding. Many thanks also to MGPalaeo and Dr. Kim Baublys for assistance with sample preparation. We would also like to extend our thanks to the reviewers Benjamin Bomfleur and Cortland Eble whose comments greatly improved the manuscript.

97

Figure 5.8: Phytoclasts, palynomorphs and fungal hyphae, borehole Tambo 1-1A (Galilee Basin). Taxon name is followed by sample number, slide number (brackets) and stage coordinates for a ZEISS Photomicroscope III. (A) Degraded phytoclasts, TAMP8 (b), 97.1/8.2. (B) Playfordiaspora crenulata, TAMP4 (a), 91.8/0.2. (C) Protohaploxypinus microcorpus, TAMP6 (a), 114.3/17.8. (D) Fungal hyphae?, TAMP10 (b), 97.2/2.4. (E) Microreticulatisporites bitriangularis, TAMP18 (a), 100.3/12.2.

98

Figure 5.9: Palynomorphs of borehole Tambo 1-1A (Galilee Basin). Taxon name is followed by sample number, slide number (brackets) and stage coordinates for a ZEISS Photomicroscope III. (A) Micrhystridium sp., TAMP10 (b), 99.7/1.2. (B) Cymatiosphaera gondwanensis, TAMP17 (a), 87.2/3.3. (C) Botryococcus sp., TAMP22 (a), 80.8/13.9. (D) Botryococcus?, TAMP3 (a), 104.2/16.7. (E) Reduviasporonites chalastus, TAMP3 (a), 109.3/17.2.

99

Figure 5.10: Palynofacies of borehole Tambo 1-1A (Galilee Basin). (A) Palynofacies Assemblage A featuring a high proportion of palynomorphs, particularly bisaccate pollen. (B) Palynofacies Assemblage B featuring a moderate to high proportion of translucent phytoclasts (structured and unstructured) as well as a moderate proportion of opaque phytoclasts and some well-preserved palynomorphs. (C) Palynofacies Assemblage C featuring a high proportion of opaque phytoclasts and a low proportion of translucent phytoclasts; palynomorphs are rarely well preserved.

100

6. Palynological Investigation of Variation in the Palaeoflora and Palaeoenvironment following the end-Permian Extinction (Bowen Basin, Australia)

6.1 Abstract

The global biotic crises at the end of the Permian led to severe disruption of palaeofloral communities following the extinction of the Glossopteris-flora in Gondwana. In this study, we examine the composition of the post-extinction flora immediately above the extinction interval at four different localities within the Bowen Basin. We also use palynological indicators of palaeoenvironment (algae, acritarchs, fungi) to reconstruct palaeoenvironmental changes occurring in the basin. The extinction interval itself appears to be synchronous and can be placed in the immediate roof strata that overlies the uppermost Permian coal seams. The succeeding paleoflora appears to have consisted of hardy survivors and pioneers such as ferns and lycophytes that occupied flooded lowland areas. Peltasperms and voltzialean conifers are likely colonisers previously confined to upland areas. The “fungal spike” elsewhere associated with the P-T Boundary was only observed in the Springsure Shelf. Pyritic pockmarking of palynomorphs and organic material was observed in the eastern Taroom Trough, suggesting potentially anoxic conditions. Barren samples, reworked palynomorphs and fluctuating algal abundances suggest increased erosion in the hinterland that is related to the extinction of Glossopteris and subsidence from tectonic loading occurring on the eastern margin of the basin.

6.2 Introduction and Aims

Terrestrial exposures of the Permian-Triassic boundary are often poorly preserved and difficult to correlate with more complete marine sections due to uncertainty regarding the synchronicity of the terrestrial and marine extinction events (Twitchett et al., 2001; Gastaldo et al., 2009; Shen et al., 2011). The Permian-Triassic boundary is usually detected by micro- and turnover as evidence of the end-Permian extinction event (Stemmerik et al., 2001; Steiner et al., 2003; Benton et al., 2004; Jianxin et al., 2007; Smith & Botha-Brink, 2014) geochemical proxies and rapid sedimentological changes (Holser et al., 1989; Maruoka et al., 2003; Wang & Visscher, 2007).

The Permian-Triassic boundary in the Bowen-Gunnedah-Sydney Basin System has been studied for many years (Balme & Helby, 1973; Foster, 1982; Retallack, 1995; Michaelsen, 2002). The Sydney Basin features extensive outcrops exposing the uppermost Permian coals and Lower Triassic sequences (Connolly & Ferm, 1971; Liu et al., 1996). In the Bowen Basin, core is easily accessible,

101 but outcrops are generally limited to highwalls in coal mines. Palynology has been applied in conjunction with high-resolution zircon dating to calibrate the biostratigraphic scheme and to date the Permian-Triassic boundary as accurately as possible (Metcalfe et al., 2015; Laurie et al., 2016; Fielding et al., 2019). Palynology has also been a crucial tool in identifying the mass extinction interval and, more recently, in interpreting the recovery of the Early Triassic post-extinction flora in the Sydney Basin (Fielding et al., 2019; Mishra et al., 2019; et alMays et al., 2020a; Vajda et al., 2020). These studies focus on high-resolution sampling of one or two sections in the Sydney Basin. Thus, to corroborate these results on an intra- and interbasinal scale, it is necessary to examine multiple sections in different parts of a basin to gain insight into local and regional environmental and floral patterns.

The aims of this study are to: (1) Identify the end-Permian floral turnover as marked by the transition from the APP5 to the APP6 zones in different localities across the Bowen Basin. This is to determine if the boundary is synchronous. (2) Examine differences in the earliest Triassic floras in each studied locality around the Bowen Basin using palynology as a proxy. (3) Identify palynomorphs that might be useful indicators of environmental perturbation (e.g., tri- and tetrasaccate pollen, algae, acritarchs, fungi). These indicators may provide insight into the palaeoenvironmental conditions that the post- turnover flora inhabits.

6.3 Background

6.3.1 Tectonics and Geological Setting

The Bowen Basin forms part of the Bowen-Gunnedah-Sydney foreland basin system in eastern Australia. The basin has a complex tectonic history beginning with an extensional back-arc phase in the early Permian, followed by thermal senescence in the mid-Permian, and shifting to a foreland loading phase in the late Permian and Early Triassic (Murray, 1983; Fielding et al., 2000; Draper, 2013). The basin fill of the Bowen basin contains a mix of mainly clastic sedimentary rocks with some volcanic units, and can reach up to 10 km in thickness in the major depocentres (Baker et al., 1993). Sedimentary deposition occurred in a variety of depositional environments from terrestrial to paralic and marine, and includes significant coal deposits (Brakel et al., 2009). The Bowen Basin is separated into structural units (Fig. 6.1) based on the basement topography. The Nebo Synclinorium and Collinsville Shelf are located in the north, the Springsure Shelf and Denison Trough are located in the west, the Roma Shelf in the south-west, the Comet Ridge in the centre and the Taroom Trough in the east (Fig. 6.1) (Dickins & Malone, 1973).

102

The most extensive and thickest coal measures formed in the late Permian and are architecturally very complex as their distribution and splitting patterns are controlled by the complex basement topography and structural history of the basin (Ayaz, 2015; Sliwa et al., 2017). The formations hosting the uppermost Permian coal measures in the Bowen Basin are split into three geographical tiles (Fig. 6.1): The Bandanna Formation in the west, the Rangal Coal Measures in the north and the Baralaba Coal Measures in the east (Sliwa et al., 2017). Coal formation ceases at the top of these formations presumably related to the end-Permian extinction event initiating a global coal gap; as such, the top of the coal measures has traditionally been used to mark the Permian-Triassic Boundary in eastern Australia (Retallack et al., 1996)

A mudstone unit, termed the marker mudstone, that occurs directly above the uppermost Permian coals was initially examined as a marker for the Permian-Triassic boundary using palynology, sedimentology, and carbon isotope geochemistry (Michaelsen et al., 2000; Michaelsen, 2002). It was later mapped by Sliwa et al. (2017) along the western margin of the Bowen Basin. In the Nebo Synclinorium, the uppermost Permian coal seam is usually the to the Leichhardt Seam, though in many cases a rider seam termed the Phillips Rider splits above this seam and underlies the marker mudstone (Matheson, 1990; Sliwa et al., 2017). The coal seam terminology is complex due to significant seam splitting and the stratigraphic equivalents of these seams have different names in the Bandanna and Baralaba tiles or informal names used by the coal mines (Fig. 6.2) (Staines, 1972; Matheson, 1990; Sliwa et al., 2017). This has made correlation of the marker mudstone to other parts of the Bowen Basin and the Galilee Basin challenging. For example, Wheeler et al. (2020) examined a carbonaceous shale in the Galilee Basin that was hypothesized to be a correlative of the marker mudstone in the Bowen Basin (Chapter 5). However, the palynological analysis showed a transition to APP6 floras occurring above this unit whereas, in the Bowen Basin, the APP6 zone was observed to begin at the base of the marker mudstone. This raises questions about the correlation of these two units.

The basal part of the Rewan Group (also referred to as the Rewan Formation in the Galilee Basin) consists mainly of fine- to coarse-grained sandstones and green ferric clay-rich mudstones, though there is local variation in the thickness of each lithology, grain size and sedimentary cycles (Baker, 1997; Grech, 2001; Wilson, 2017). This basal unit of the Rewan Group referred to as the Sagittarius Sandstone and represents an extensive alluvial plain that formed during the Early Triassic in response to foreland loading and basin overfilling (Fielding et al., 1993; Sliwa et al., 2017). At the localities chosen for this study, the base of the Rewan Group is conformable above the late Permian coal measures. The transition above the uppermost Permian coal seam is usually marked by a gradual shift in colour from a grey to green (Matheson, 1990; Sliwa et al., 2017). Where present, the marker

103 mudstone is carbonaceous and dark grey with the transition occurring above it, raising the question about whether it should be classified as part of the Rewan Group or the underlying Permian sequence (Michaelsen et al., 2000).

Figure 6.1: Location of studied boreholes in the Bowen Basin. Borehole Tambo 1-1A previously studied in the Galilee Basin is also indicated (Wheeler et al., 2020).

104

Figure 6.2: Broad seam terminology used in each tile of the Bowen Basin and their stratigraphic equivalents, Local mines may apply their own terminology (from Sliwa et al., 2017).

6.3.2 The P-T Boundary

The Permian-Triassic boundary represents a period of significant change in Earth’s climate and biota (Erwin, 1993). The end-Permian extinction affected marine and terrestrial fauna and flora and led to the extinction of previously dominant groups such as Glossopteris, which was a major component of the floral ecosystem all over Gondwana (Benton & Twitchett, 2003). The boundary itself is formally defined by the first appearance datum of the conodont Hindeodus parvus in the Meishan Beds in China (Yin et al., 2001). Proxies that are used to mark the P-T Boundary in other basins include a negative carbon isotope excursion, turnover in the macro- and microflora, vertebrate extinction, fungal, algal and acritarch spikes, and elemental spikes in nickel and mercury. Currently the most widely cited U/Pb zircon age dates of the boundary interval have placed the FAD of H. parvus at 251.902 ± 0.024 Ma (Burgess et al., 2014).

In the basins of eastern Australia, the boundary has generally been placed using palynology. The APP5 zone (or Dulhuntyispora parvithola biozone) is regarded to represent the climax Glossopteris

105 flora during the late Permian (Fig. 6.3), due to the high abundance of striate bisaccate pollen grains associated with Glossopteris (Foster, 1982; Price, 1997). Foster (1982) erected a lower Playfordiaspora crenulata Zone based on a significant assemblage change from the APP5 zone and the appearance of Playfordiaspora crenulata. Above this, he proposed the Protohaploxypinus microcorpus Zone based on further modifications to the assemblage and the appearance of Protohaploxypinus microcorpus. Price (1997) later erected the APP6 zone based on the first appearance of Triplexisporites playfordii with the APP601 and APP602 subzones being correlative with the P. crenulata and P. microcorpus Zones respectively. Above this stands the APT1 zone based on the appearance of Lunatisporites pellucidus (Price, 1997).

Recent palynological studies have focused on calibrating the palynostratigraphic scheme using high- resolution zircon age dates (Smith & Mantle, 2013; Metcalfe et al., 2015; Laurie et al., 2016). In the Sydney Basin, these age dates suggest that the floral turnover representing the terrestrial expression of the end-Permian extinction occurs slightly earlier than the GSSP-defined boundary age (Metcalfe et al., 2015; Fielding et al., 2019). The floral turnover occurs at the interface of the APP5 and APP6 zones (Mays et al., 2020a; Vajda et al., 2020) but the P-T boundary itself has been placed in the Lunatisporites pellucidus Zone (Metcalfe et al., 2015) and more recently in the Playfordiaspora crenulata Zone (Fielding et al., 2019). This work examines the timing and aftermath of the floral turnover, the composition of the surviving flora and the palaeoenvironment in which this flora survived.

Figure 6.3: Late Permian stratigraphy of the Bowen Basin (modified from Ayaz et al., 2015).

106

6.4 Materials and Methods

6.4.1 Sampling

Four localities in different parts of the Bowen Basin were selected for sampling: Springsure 19 on the Springsure Shelf, Taringa 7 on the Roma Shelf, CGIN0067 and CGIE0144 from the Isaac Plains Coal mine on the Collinsville Shelf, and A41859 from the Dawson Coal Mine in the eastern part of the Taroom Trough (Fig. 6.1). Fine-grained, organic-rich siltstones and mudstones in the immediate roof above the coal seam were targeted for sampling. A total of 46 samples were collected. The presence of fine- to coarse-grained sandstones hindered sampling and the green, clay-rich mudstones in the Rewan Group tended to yield poorly or not at all.

6.4.2 Palynology

Samples from Isaac Plains, Dawson and Taringa 7 were sent to MGPalaeo, a commercial palynological lab to undergo standard palynological processing. 15 – 23 g of sample had 32% HCl and 48% HF added to remove carbonate and silicate minerals, respectively. After digestion the acid was decanted and the residue washed and centrifuged.10 – 15 ml of lithium heteropolytungstate (s.g. 2.1) and the residues were centrifuged to further concentrate organic matter and remove minerals before sieving using a 10 µm polycarbonate filter to remove fine organic matter.

Samples from Springsure 19 were processed at The University of Queensland using an acid-free processing technique (chapter 3). 10 – 20 g of sample were crushed and submerged in 10% Sodium Hexametaphosphate to deflocculate the clays. The residue was then washed through an 8 μm mesh to remove the clay minerals. Density separation was then carried out using Sodium polytungstate (s.g. 1.9) to remove any remaining minerals. The residue was then mounted on slides using EUKITT, a resin-based mounting medium.

Samples were examined using a ZEISS Photomicroscope III and a Leica MC190HD camera. Counts were done up to 200 specimens where possible and plotted using the Stratabugs software package. Specimens that exhibited potential pyritic pockmarking were examined using a Hitachi TM3030PLUS scanning electron . The mix function (secondary and backscatter electrons) allowed for differentiating the organic materials and small clay particles as well as clearly displaying the topography to identify pockmarks.

107

6.5 Results

A41859 (Dawson Coal Mine - eastern Taroom Trough)

Two sets of samples were collected from A41859: UQ1 to UQ6 and UD1 to UD7. These samples had a relatively high thermal maturity and samples taken from the Rewan Group were poorly preserved due to significant pyrite degradation. However, the palynomorph yield was high and species assignments were possible. Taphonomic bias may favour the more resilient taxa. Samples UD1, UD2 and UD7 were barren of palynomorphs.

From the base of the section, samples UQ6 to UD6 feature a typical late Permian APP5 assemblage (Fig. 6.3). Trilete spores such as Leiotriletes and Microbaculispora are dominant components of the assemblage in samples UQ2, UQ3 and UQ6. UD6 and UQ5 feature higher proportions of bisaccate pollen grains (mainly Protohaploxypinus, Striatopodocarpites and Alisporites). Other consistent pollen components of the Permian assemblage include Marsupipollenites, Triadispora, Vitreisporites and Striatoabieites. Other rare but consistent spore components include Brevitriltes, Horriditriletes, Osmundacidites, and Calamospora.

Samples from UD5 up to UD3 feature a characteristically APP6 flora containing a moderate abundance of the index taxa Triplexisporites playfordii (6 – 11%) as well as minor components of other index taxa such as Playfordiaspora crenulata and Protohaploxypinus microcorpus. Major components of the assemblage are a variety of cavate trilete spores (Lundbladispora, Indotriradites) as well as Brevitriletes bulliensis and Calamospora. Striate bisaccate pollen grains are rare but non- striate bisaccate pollen grains (Alisporites australis) are common. UQ1 features all components of the APP6 assemblage, but also several components of the APT1 assemblage (Lunatisporites pellucidus, Lunatisporites noviaulensis, Limatulasporites spp.) in rare abundances (less than 3%). This sample also includes other trilete spores common in APP6 and APT1 assemblages such as Densoisporites playfordii and Thymospora ipsviciensis. Unseparated spore tetrads were also observed. The pyrite pockmarking that affects the palynomorphs and palynodebris of the Rewan Group samples (UD5 to UQ1) tends to concentrate in the bisaccate pollen grains and the outer cingulum of cavate trilete spores like Playfordiaspora crenulata.

108

Figure 6.4: Palynostratigraphy of A41859 (eastern Taroom Trough). Palynological data displayed as a relative abundane (%). AC: acritarchs; AL: freshwater algae; ALPR: prasinophytes; FU: fungi; PO: pollen; SP: spores. For raw counts see Appendix B4.9.

109

CGIN0067 and CGIE0144 (Isaac Plains Coal Mine - Collinsville Shelf)

Of 9 samples collected from CGIN0067 and CGIE0144, only 4 yielded enough palynomorphs to count. These samples have a relatively high thermal maturity, but organic particles are still preserved well enough to be identifiable.

Two samples from CGIE0144 (IPA3 and IPA4) capture the palynological assemblage from the immediate floor and roof of the Leichhardt Seam (Fig. 6.5A), which is conformable with the overlying Rewan Group. The immediate floor of the Leichhardt Seam (IPA4) features a typical late Permian assemblage dominated by glossopterid pollen (Protohaploxypinus and Striatopodocarpites) and fern spores (Cyclogranisporites, Microbaculispora). A single specimen of the APP5005 index taxon Microreticulatisporites bitriangularis was also identified though no specimens of APP5006 index taxon Lycopodiumsporites “crassus” were observed. APP6 floras are present in the immediate roof above the Leichhardt Seam (IPA2). Index taxa for APP6 zone are present in rare to moderate abundances (Protohaploxypinus microcorpus, Triplexisporites playfordii and Playfordiaspora crenulata). This sample is dominated by trilete spores, particularly Brevitriletes bulliensis (14%) and Leiotriletes directus (15.5%), along with moderate abundances of Calamospora (6.5%) and Microbaculispora (8.5%), Alisporites sp. (8%) and Alisporites australis (7%). However, IPA2 also features rare specimens of Lunatisporites pellucidus (1%) and Lunatisporites noviaulensis (2.5%), which are more characteristic of an APT1 assignment.

Two samples from CGIN0067 (IPB1 and IPB5) record the assemblage from the marker mudstone and the overlying siltstones and sandstones (Fig. 6.5B). Sample IPB5 from the marker mudstone features a diverse spore assemblage that includes a moderately high abundance of Brevitriletes (particularly Brevitriletes bulliensis) (12%) and Lundbladispora (13%) as well as relatively common abundances of Densoisporites (9%) and Microbaculispora (8%). Rewanispora foveolata (6%) and APP6 index taxon Triplexisporites playfordii (6%) are slightly less common but still present. Pollen grains are rare in this sample. Sample IPB1 overlying the marker mudstone features a much higher abundance of the pollen taxon Alisporites (13%). Microbaculispora is the most abundant spore taxon (19%) along with Brevitriletes (13%), Densoisporites (8%), Lundbladispora (9%) and Leiotriletes (7%). Both samples feature rare Lunatisporites pellucidus, Lunatisporites noviaulensis, as well as spores like Limatulasporites and Rewanispora which are more characteristic of an APT1 assemblage. A single specimen of Krauselisporites saeptatus was also observed in IPB1. IPB5 features rare Protohaploxypinus microcorpus (0.5%), while IPB1 features rare Playfordiaspora crenulata (1%).

110

A.

B.

Figure 6.5: Palynostratigraphy of A) CGIE0144 and B) CGIN0067 (Collinsville Shelf). Palynological data displayed as a relative abundane (%). AL: freshwater algae; FU: fungi; PO: pollen; SP: spores. For raw counts see Appendices B4.7 and B4.8.

111

Springsure 19 (Springsure Shelf)

Of the 14 samples taken from the Rewan Group and Bandanna Formation 10 yielded countable abundances of palynomorphs. SPR1, SPR6 and SPR10 were barren of palynomorphs. The samples that did yield, were well preserved and contained a diverse palynological assemblage.

Samples taken from the Bandanna Formation exhibit a typically APP5 late Permian assemblage featuring high abundances of Protohaploxypinus and Striatopodocarpites (Fig. 6.6). The index taxon Dulhuntyspora parvithola is rare but present from SPR12 upwards. APP5005 index taxa Microreticulatisporites bitriangularis and APP5006 index taxa Lycopodiumsporites “crassus” are both present with L. “crassus” reaching up to 11% in SPR11. Algal taxa are rare but relatively diverse with common to rare Botryococcus and several types of Zygnematacean algae.

Three samples taken in the immediate roof less than 1 m above the uppermost Permian coal seam (SPR5, SPR4 and SPR3) show distinct and rapid changes from the underlying assemblage. The first sample which sits just above the coal seam features a particularly high abundance of Botryococcus and other freshwater to brackish algae (Quadrisporites horridus) appearing further upsection. SPR5 features a distinct spike (7%) in Reduviasporonites chalastus and the index taxa for APP6 Playfordiaspora crenulata and Triplexisporites playfordii are both present. An increase in the abundance of Brevitriletes bulliensis (11%) occurs in SPR5 10 to 12 cm above the uppermost coal seam along with Alisporites (14%) and Calamospora (7%). Above this, the abundance of Brevitriletes (particularly Brevitriletes bulliensis) and Verrucosisporites increases further. Triquitrites proratus is also present but in relatively lower abundances compared to the Dawson and Isaac Plains localities. A single specimen of Dulhuntyispora parvithola was observed in SPR5. Alisporites is particularly abundant in SPR5 (14%) and SPR3 (10%), while several non-striate pollen grains display a tri- and tetra-saccate morphology.

112

Figure 6.6: Palynostratigraphy of Springsure 19 (Springsure Shelf). Palynological data displayed as a relative abundane (%). AL: freshwater algae; ALBO: chlorophycean algae; ALPR: prasinophytes; FU: fungi; PO: pollen; SP: spores. For raw counts see Appendix B4.5.

113

Taringa 7 (Roma Shelf)

Of three samples taken above the uppermost Permian coal seam, only the sample taken immediately above the roof (TAR3) yielded adequate palynomorph content (Fig. 6.7).

Samples TAR6 and TAR7 in the Bandanna Formation were also barren of palynomorphs. Samples from the Bandanna Formation and Black Alley Shale (TAR4 to TAR10) generally feature high abundances of Protohaploxypinus (up to 50% in TAR4). These samples also feature varying abundances of Leiotriletes, Microbaculispora, Didecitriletes and Striatopodocarpites. APP5005 index taxon Microreticulatisporites bitriangularis was also observed in both the Black Alley Shale and Bandanna Formation.

TAR3, which was sampled 0 – 3 cm above the uppermost Permian coal seam (Fig. 6.6) features a very low diversity of pollen taxa, dominated by a high proportion of Alisporites australis (29%). The highest proportions of spores were Microbaculispora (10%), Leiotriletes (9%), Horriditriletes (9%), Brevitriletes (8%) and Calamospora (7%). The index taxa Triplexisporites playfordii, Playfordiaspora crenulata and Protohaploxypinus microcorpus were present but rare. Specimens of P. microcorpus tend to be damaged and tear transversely along the corpus. The only cavate trilete taxon present was Lundbladispora. Botryococcus is also present (7%) but these specimens are relatively small in size compared to specimens observed in other sections. A single specimen of the algal taxon Quadrisporites horridus was observed in TAR3.

114

Figure 6.7: Palynostratigraphy of Taringa 7 (Roma Shelf). Palynological data displayed as a relative abundane (%). AL: freshwater algae; ALBO: chlorophycean algae; ALPR: prasinophytes; FU: fungi; PO: pollen; SP: spores. For raw counts see Appendix B4.6.

115

6.6 Discussion

6.6.1 Preservation

Vitrinite reflectance in the Bowen Basin increases to the north in the Nebo Synclynorium and to the east in the north-eastern Taroom Trough (Beeston, 1986; Uysal et al., 2000; McKillop, 2016; Sliwa et al., 2017). Thus, palynological samples acquired from the Collinsville Shelf sections and eastern Taroom Trough have a relatively high thermal maturity (i.e., reflectance greater than 1.0%) and tend to be poorly preserved. However, the palynomorph yield from some of the samples in these sections was quite high and the quality of preservation sufficient for species assignment. Certainly, the taphonomic record may bias the results to more resilient taxa. The high number of samples barren of palynomorphs may reflect the “dead zone” observed in the Sydney Basin, though more samples would be necessary to confirm if it is a regional event (Mays et al., 2020a; Vajda et al., 2020).

6.6.2 Biostratigraphy of the end-Permian Extinction

The overall assemblage shift from the APP5 zone to the APP6 and potentially the APT1 zones appear to occur in the immediate roof above the uppermost coals in all studied sections (Springsure 19, the Collinsville Shelf sections, A41859 and Taringa 7). This also marks a significant decrease in the abundance of pollen associated with Glossopteris (Protohaploxypinus, Striatopodocarpites). At the same time there are relative increases in spores, particularly Brevitriletes and cavate trilete forms like Lundbladispora and Indotriradites. The APP6 zone is also defined by the appearance of certain index taxa. Price (1997) used Triplexisporites playfordii, which appears consistently in all studied sections. Playfordiaspora crenulata appears in lower abundances but was still observed in all studied sections. The occurrence of Protohaploxypinus microcorpus is sporadic, but it does appear in the sample immediately above the uppermost Permian coal seam in CGIE0144. These specimens are relatively poorly preserved and tend to split transversely along the corpus. Some of these index taxa (Playfordiaspora crenulata, Triplexisporites playfordii) are observed in Permian sections in other parts of Gondwana and appear to radiate from lower latitudes to higher latitudes perhaps as a function of changing climate (Lindström & McLoughlin, 2007). This makes utilising them for continental- scale palynological correlations of the Permian-Triassic Boundary difficult in Gondwana. The presence of taxa associated with APT1, most notably in sample UQ1 in A41859, and sample IPB1 in CGIN0067 appears to be somewhat consistent with Michaelsen (2002), who assigned an APT1 age to the Sagittarius Sandstone immediately above the marker mudstone. This would indicate the APP6 zone has a highly variable thickness, as it can be up to 22 m in the Newlands Coal Mine, while in the

116 sections studied herein, it is less than 10 m in thickness. Comparitevely, the combined thickness of the Playfordiaspora Crenulata and Protohaploxypinus microcorpus Palynozones (equivalent to the lower and upper APP6 subzones respectively) in the Sydney Basin can be close to 45 m in thickness (Fielding et al., 2019). This could be due to higher rates of sedimentation in the Sydney Basin compared to the Bowen Basin, however, this explanation is complicated the presence of taxa such as Lunatisporites and Limitulasporites within and below the marker mudstone in the Collinsville Shelf sections. The parent plants of these taxa may have already been present on the northern margin of the Bowen Basin albeit as a still minor component of the basinal flora. Only once the fluvial-dominated palaeoenvironment of the Sagittarius Sandstone became established across the Bowen Basin, did these plants become the dominant basinal floras. Similarly, in the Sydney Basin, Mays et al., (2020a) describes the shift from the Protohaploxypinus microcorpus Palynozone to the Lunatisporites pellucidus Palynozone to represent a shift from a supressed recovery flora, to a steady recovery flora following the EPE.

6.6.3 Correlating the marker mudstone

Wheeler et al. (2020) examined a marker mudstone candidate in the borehole Tambo 1-1A in the Galilee Basin. Results showed the transition to APP6 occurred above the mudstone based on the appearance of Protohaploxypinus microcorpus, Playfordiaspora crenulata and Triplexisporites playfordii, but elements of the APP6 flora (Brevitriletes bulliensis, Thymospora ipsviciensis) and aquatic elements (Micrhystridium, Botryococcus) are present albeit within and above the marker mudstone in very rare abundances. Results from CGIN0067 and CGIE0144 agree with the placement of the Bowen Basin marker mudstone in the APP6 zone (Michaelsen, 2002). The lithological sequence in Springsure 19 bears a striking resemblance to that of Tambo 1-1A but coal is present in place of a carbonaceous mudstone at the top of the Bandanna Formation. This coal is 1 meter thick and appears in a similar stratigraphic position to the marker mudstone in the Galilee Basin. Based on the palynology we can infer that the marker mudstone identified in Tambo 1-1A is actually the lacustrine equivalent to the uppermost Permian coal seam but the presence of elements of the APP6 flora and acanthomorph acritarchs raise questions about environmental perturbation immediately preceding the cessation of coal deposition.

117

6.6.4 Composition and Distribution of the post-Extinction Flora

The late Permian flora was dominated by glossopterids and an undergrowth of ferns, horsetails and lycophytes (Shi et al., 2010). The Glossopterids appear to be significantly affected during the end- Permian extinction event and quickly disappear from the fossil record immediately above the boundary. Initially after the extinction, ferns remained a dominant part of the assemblage as inferred from a relative increase in trilete fern spores. The putative marattialean fern spore, Thymospora ipsviciences (Balme, 1995), shows a distinct increase in abundance in the earliest Triassic, suggesting localised increase in the abundance of certain fern groups as a part of the recovery flora. Isoetalean lycophytes also increase in abundance coinciding with the environmental shift following the EPE. Macrofloral remains of Isoetes beestonii are known from both the Rewan Group in the Bowen Basin and the Coal Cliff Sandstone in the Sydney Basin (Retallack, 1997). Plueromeian lycophytes are absent in the lowermost Triassic strata with macro- and microfossil remains only appearing in the Protohaploxypinus samoilovichii Palynozone (Retallack, 1997; Mays et al., 2020). Macrofloral remains of peltasperms like Lepidopteris callipteroides (which is botanically affiliated to Alisporites australis, as are other seed fern groups) have been observed in the Dooralong Shale immediately above the Vales Point coal seam in the Sydney Basin (Retallack, 2002). In the Bowen Basin, we find Alisporites australis in samples taken in the roof of the uppermost Permian coal seam. Voltzialean conifers also appear in the roof shales less than 2 metres above the coal in the Sydney Basin (Vajda et al., 2020). Dispersed pollen found alongside Voltziopsis specimens distinctly resemble Protohaploxypinus microcorpus (Townrow, 1967). It is plausible that these gymnosperms were previously confined to the upland and basin margin during the late Permian and expanded their distribution colonising the lowlands as the climate further ameliorated and fluvial processes became dominant and peat accumulation had ceased. Floral diversity is relatively low in the earliest Triassic, and likely represents a supressed recovery flora (Mays, et al., 2020). It was only in the late Early Triassic, that floral diversity exploded, heralding the dominance of the Dicroidium flora (Escapa et al., 2011).

In the APP6 assemblages, herbaceous floral elements, which were previously supressed by the dominant glossopterids, became the major component of the floral assemblage. Highly stress- resistant forms and pioneer groups such as the isoetalean lycophytes were apparently able to withstand the highly unstable conditions in the immediate aftermath of the extinction (Looy et al., 2001). They were more likely to colonise flooded areas and the margins of the marker mudstone lacustrine system along with some of the more resilient ferns. The disappearance of glossopterids and warmer temperatures also opened niches for other gymnosperm floras like Voltzialean conifers and

118 peltasperms, previously restricted to the uplands and basin margins (McLoughlin, 1993), forming open woodlands on the broad floodplains and broadening alluvial plain (Fig. 6.8).

A.

B.

Figure 6.8: (A) Late Permian “Climax” flora; (B) Post-extinction flora consisting of resistant pioneer plants and colonisers from the upland areas.

119

6.6.5 P-T Boundary Palaeoenvironment in Eastern Australia

While there is no indication of an unconformity in the sections presented herein, other authors have identified lithological unconformities in parts of the Bowen and Galilee basins that appear to have removed the interval containing the P-T boundary (Grech & Dyson, 1997; Korsch et al., 2009). The presence of palynomorphs, previously limited to the Permian, in Early Triassic sections (i.e., reworked Dulhuntyispora in Springsure 19) does suggest localised weathering and reworking in the Bowen Basin. High levels of weathering have been described based on sedimentological and geochemical studies at other Permian-Triassic sections (Retallack, 2005; Sephton et al., 2005). The loss of the extensive Glossopteris forest cover that made up the majority of biomass in the late Permian floras almost certainly affected the erosional dynamics and water retention within the basin (Vajda et al., 2020). This has been observed in clay geochemical proxies showing a strong shift to kaolinite, suggesting increased weathering. High levels of weathering and soil erosion have significant knock-on effects. Increased sedimentary input into lacustrine and shallow marine systems can lead to eutrophication, bottom water hypoxia and even anoxia (Diaz & Rosenberg, 1995; Rabalais et al., 2002; Sephton et al., 2005). The presence of the marker mudstone and associated sporadic increases in algae and acritarchs suggest that limited flooding did occur in the Bowen Basin. Reasons for this flooding could be related to foreland loading (Wheeler et al., 2020), the removal of the Glossopteris flora in a ‘fire and algal event” (Vajda et al., 2020) or differential compaction of the peat creating accommodation for subsequent lacustrine development (Michaelsen et al., 2000). In the Bowen Basin, this potentially creates two environments for plants to colonise: a well-drained fluvial- floodplain affected by wildfire, followed by increased erosion; and a lacustrine system affected by base-level rise, salinity fluctuations and anoxia. Acritarch spikes indicative of a short-term Early Triassic base-level rise were observed in both western and eastern Australia (Retallack, 1995; Gorter et al., 2009).

The presence of unseparated spore tetrads (Fig. 6.9) is a notable occurrence in many terrestrial P-T localities (Visscher et al., 2004; Looy et al., 2005; Hochuli et al., 2010). Generally, these are limitied to the lycopod spores (e.g., Lundbladispora) but can occasionally be seen in fern spores (e.g., Brevitriletes). Another unusual occurrence is increased abundances of tri- and tetrasaccate pollen grains. In the samples examined herein, this type of pollen remains rare, but that may be an effect of the generally low abundances of pollen in the samples and the poor preservation of pollen that has remained. Both of these phenomena have been attributed to increased levels of UV-B radiation caused by a damaged ozone layer (Vissher et al., 2004; Foster & Afonin, 2005). Ozone damage is speculated to be either a direct result of sulphate emission from the Siberian LIP or from hydrogen sulphides released from the ocean (Kump et al., 2005; Beerling et al., 2007), though there is some disagreement

120 whether these emissions would have significantly affected the ozone layer (Kaiho & Koga, 2013). Benca et al. (2018) irradiated modern conifers and examined the effects on the parent plant and their pollen, speculating that radiation-induced sterilisation may have contributed to the extinction of the glossopterids.

Figure 6.9: Key palynomorphs that act as palaeoenvironmental indicators: a) Unseparated spore tetrad; b) Tetrasaccate pollen grain; c) Quadrisporites horridus; d) Rediviasporonites chalastus; e) Botryococcus; f) Cymatiosphaera gondwanensis.

121

Pyrite scarring on palynomorphs can occur as a result of biogenic processes in anoxic, stagnant water during the early stages of diagenesis (Neves & Sullivan, 1964; Tiwari et al., 1990; Tiwari et al., 1994). Occasionally, framboidal pyrite can be found within these pits, but even when it is absent the morphology of the pits has been used to infer its former presence (Srivastava et al., 1999). Framboidal pyrite found in Permian-Triassic boundary sections likely indicates bacterial sulphate reduction in anoxic, potentially euxinic conditions (Shen et al., 2007; Bond & Wignall, 2010). Herein, examination of the pits using standard microscopy and scanning electron microscopy (Fig. 6.10) did not provide conclusive evidence of framboidal pyrite. However, pyrite pockmarks have also been noted within algal microfossils immediately above the extinction horizon in the Sydney Basin and are suggested to be related to bacterial reduction of iron in an anoxic lacustrine environment (Vajda et al., 2020). Pyritic pockmarking textures identical to those observed in A41859 can be seen in specimens of Reduviasporonites chalastus from borehole N.S.30 in the eastern Taroom Trough (Foster et al. (2002): Plate 1), suggesting that this phenomenon is not limited to a single isolated locality, but may be characteristic of this part of the Bowen Basin. Brakel et al. (2009) cited high sulphur and boron content as possible evidence for a marine incursion entering the Bowen Basin from the northeast. Boron has previously been cited as indicative of marine incursion in the Newcastle Coal Measures in the Sydney Basin (Creech, 2002). However, Foster (1982) notes the occurrence of the taxon “Micrhystridium?” sp. B. in the basal Rewan Group in the eastern part of the Taroom Trough. This specimen closely resembles Mehlisphaeridium, which is likely of zygnematacean origin and is not associated with marine conditions in the way that acanthomorph acritarchs are (Zavattieri et al., 2017).

122

Figure 6.10: SEM and light microphotographs of pyritic pockmarking in organic matter from A41859: a, d) pyritic pockmarking in a bisaccate pollen grain; b, c) pyritic pockmarking in woody debris (small, light particles covering organic matter are clay minerals); e) pyritic pockmarking in Playfordiaspora crenulata; f) pyritic pockmarking in a structured phytoclast.

123

6.7 Conclusions and Outlook

The floral turnover marked by the interface of the APP5 and APP6 zones appears consistently at the top of the uppermost Permian coal seams. The lack of any evidence of erosion at this horizon suggests that the terrestrial expression of the end-Permian extinction is synchronous across the Bowen and Galilee basins. Palaeoenvironmental indicators suggest a high base level which led to the formation of the marker mudstone in the Bowen Basin, fluctuating salinity, algal blooms and localised flooding. Ozone damage is also suggested by the presence of the abnormal palynomorphs. The putative marker mudstone in the Galilee Basin is likely not a direct correlative of the marker mudstone in the Bowen Basin, but is instead a distal correlative of the uppermost Permian coal seam. The extinction of the Glossopteris-flora and the extreme palaeoenvironmental changes led to a post-extinction flora dominated by ferns and isoetalean lycophytes occupying flooded lowland areas, with peltasperms and voltzialean conifers occupying the more well-drained alluvial plains. Future work should focus on examining the palynology, organic petrology of polished sections and geochemistry of the uppermost Permian coal seams for environmental perturbation immediately below the floral turnover event. Further investigations will also focus on examining the floral recovery within the Rewan Group for regional comparison.

124

7. Synthesis

7.1 Synthesis of the Findings

This thesis targeted the relationship between palaeofloral communities and depositional environments, against the backdrop of global biotic change through the late Permian to Early Triassic, using palynological assemblages as a proxy. For higher-resolution correlations, improved palaeoenvironmental reconstructions, and a better understanding of the Permian-Triassic boundary events, the distribution, composition and environmental characteristics of the palynological assemblages needed to be examined. Linking palynological assemblages to a sedimentological context, has helped define the spatial and temporal limits of key taxa (such as Dulhuntyispora parvithola), which in turn are controlled by their environmental preferences. Palynological assemblages studied from multiple localities in the Bowen and Galilee basins record both changes in the palaeoenvironment and a stable “climax” flora in the late Permian followed by extreme environmental perturbation in the Early Triassic.

Proximal-distal variation in Permian plant communities was shown to be related to spatiotemporal facies changes allowing for the description of more detailed plant communities and their potential environmental preferences. The investigation of a laterally continuous carbonaceous mudstone in the southern Galilee Basin as a potential correlative of the marker mudstone in the Bowen Basin found distinct environmental perturbation immediately preceding the Permian-Triassic floral turnover. This led to an evaluation of EPE synchronicity, and showed local variation in the composition of the surviving floras and the disrupted palaeoenvironment. More detailed discussion on each of these points is provided below. An additional output was a test of the utility of an acid-free palynological processing technique when compared to the conventional acid processing. This was necessary to confirm the validity of comparing palynological yields and compositions of samples from different localities that were processed by different techniques. Different techniques are often used by different labs, and this comparison addresses the uncertainties that might arise in current samples or in comparing against results obtained by others in different labs and reported in the literature.

In the Bowen and Galilee basins, these major findings of this thesis have implications for the regional biostratigraphy (particularly in relation to the distribution of Dulhuntyispora parvithola and Lycopodiumsporites “crassus”) and for palaeoenvironmental models that in turn could potentially assist in exploration and mapping of new energy resources. The observed effects of the end-Permian extinction on local and regional-scale vegetation communities in the Bowen and Galilee basins contribute to the discussion in other areas of eastern Australia and greater Gondwana. The findings of this thesis are described in greater detail below:

125

7.1.1 Acid-free Processing vs conventional Processing

The acid-free processing was demonstrated to be as effective at liberating palynomorphs from Permian-aged samples, and was comparable in quality to the conventional acid processing technique. There were some quantitative differences between the techniques, particularly in the palynofacies assemblage in that the acid-free samples featured a much higher proportion of opaque phytoclasts while several of the conventional samples featured degraded phytoclasts. The acid-free process also appeared to produce higher yields of algae, particularly Botryococcus; further work is required to determine if this is a true reflection of the environment or a result of the processing as it has implications for interpretations. The success of the acid-free technique at liberating a well-preserved assemblage of palynomorphs from samples representing several depositional environments suggests the technique is viable, though further testing and refinement is required. This would entail more samples, higher palynological counts and the use of a Lycopodium spike to measure palynomorph concentrations.

7.1.2 Palynology as a Tool for Palaeoecological Reconstruction in the late Permian and Early Triassic

Palynology has long been used as a proxy for palaeoecology, palaeoenvironment and palaeoclimate. It has been particularly effective in the marine realm as many dinoflagellate cyst species have strong ecological preferences and make good proxies (De Vernal et al., 1997; Brinkhuis et al., 1998; Vellekoop et al., 2015). Terrestrial palynomorphs such as pollen and spores tend to be less accurate proxies of the Palaeozoic climate and environments compared to geochemical proxies, plant proxies such as stomatal density and other organic palaeothermometers (Royer, 2001; Littler et al., 2011; O’Brien et al., 2017). This is due to the inherent variability and often unknown compositions of the parent plant assemblages, transport, and preservation bias (though this can also affect dispersed cuticles used for stomatal density proxies). With these limitations in mind though, palynology can be highly effective at reconstructing floral successions, and assist in palaeoenvironmental and palaeoclimatic interpretations, particularly in environments like mires, which preserve the local plant material and allows for direct comparison of the micro- and macrofloras (Beeston 1987; Hagemann & Wolf, 1987; Nichols, 1995; Nichols & Warwick, 2005; Van de Wetering et al., 2013b; Korasidis et al., 2016; Wagner et al., 2019). Palynological research on Carboniferous-age coal balls has been particulary insightful into the local coal-forming floras and how to connect micro- and macrofloral

126 datasets (Phillips et al., 1974; Phillips & Peppers, 1984; Scott, 1991; Eble, 2003; Dai et al., 2020). Palynological studies in the Mesozoic applying the Sporomorph Ecogroup (SEG) model have allowed for detailed palaeoenvironmental reconstructions based on changes in microfloral elements with strong environmental associations (Abbink et al., 2004; Ruckwied et al., 2008). The use of palynology in younger systems, where there are extant cognates, or in older systems where you have coal balls with well preserved in-situ macro- and microfloras, allows for a more straightforward examination of the connection between environments, plants and palynomorphs. Permineralised forests and peat floras have been essential for the reconstruction of the Permian vegetation (Taylor et al., 1989; Taylor et al., 1992; Slater et al., 2012; Slater et al., 2015; Nishida et al., 2018; McLoughlin et al., 2019). Studies of permineralised sporangia and their contained spores are incredibly useful for identifying botanical affinities (Lindström et al., 1997; Ryberg et al., 2012), but more studies examining dispersed spores could also offer insight into the relationship between sporomorph assemblages and the composition of the local vegetation. This thesis demonstrates that the use of palynological assembages can offer insight into the composition and spatiotemporal variation of the basinal floras, providing one consider all factors that may affect the dataset (i.e., botanical affinities, transport, preservation). Etching of intact polished blocks to reveal plant anatomy coupled with palynology may provide a similarly robust record for the Permian, as that of coal balls or permineralised peat deposits (McLoughlin et al., 2019).

This thesis combines palynological datasets with sedimentological data and depositional models to show that the composition of local assemblages is controlled in part by the depositional environment. This can be seen in a stable late Permian flora where proximal to distal facies shifts (i.e., from a fluvio-lacustrine environment to a deltaic-coastal environment) feature discernable differences even with a relatively low diversity of floral components (mainly Glossopteris, ferns and horsetails with accessory lycophytes, cordaitaleans, conifers and potentially peltasperms). Glossopteris has a broad environmental preference dominating fluvial-floodplain systems as well as forest bogs. Ferns, however, are the dominant flora of the lower delta plain along with accessory lycophytes and horsetails, and are a significant part of the understory vegetation in mire and lacustrine settings. Osmundalean ferns appear to be the dominant fern group in both deltaic and paludal/lacustrine settings, based on both the high abundance of spores and macrofloral records (McLoughlin, 1992; McLoughlin et al., 2019). Spores putatively assigned to Botryopteridales are locally abundant, potentially acting as climbers and scramblers utilising tree ferns or glossopterids as support structures in coastal marshes more prone to flooding and high salinity, similar to their use of Psaronius in the northern hemisphere (Rothwell & Stockey, 2008). Lycophyte spores are rare but consistent in their occurrence, and are mainly concentrated in coal seam partings. This suggests that the lycophytes were probably a minor component of peat forming floras, occupied lake margins, and were present in

127 coastal marsh settings. In contrast to horsetail macrofossils, spores assigned to horsetails are low in diversity and relatively rare. This is partly due to some of their spores (Calamospora, Laevigatosporites) sharing affinities with other plant groups or their spores having a lower preservation potential compared to other, more robust spores (Balme, 1995). However, horsetails are also able to reproduce vegetatively and may produce fewer spores compared to the ferns, leading to an underrepresentation of these plants in the microfossil record (Gastaldo, 1992). Cordaitaleans, conifers and cycads make up the main non-glossopterid gymnosperms, likely occupying well-drained settings such as basinal margins. Some conifers, cordataleans and cycads are associated with a forest swamp setting as an accessory to Glossopteris (Diessel, 1992; McLoughlin, 1992; McLoughlin, 1993), and results from chapter 4 suggest they may have been components of forests along the coastal plain. Pollen grains putatively assigned to peltasperms (Vitreisporites, Vitattina) are present in a number of the studied sections. As there are no macrofossil records of peltasperms in Australia during the late Permian, it is difficult to assess if they were actually a part of any basinal floras. They may have occupied the upland with their pollen being transported into the basins, or the botanical affinities of these pollen taxa are not fully known. The appearance of the peltasperm Lepidopteris in basinal floras during the earliest Triassic does seem to suggest peltasperms spread quite quickly across Gondwana, or were present in non-depositional and upland areas (Retallack, 2002).

Another significant finding of chapter 4 was that Dulhuntyispora parvithola, the index taxon for the APP5 zone, appears to be mainly limited to paralic facies. Facies control of Dulhuntyispora was previously suggested based on the taxon generally exhibiting a large size and thick spore wall, which would potentially limit the transport potential of this taxon compared to smaller spores and bisaccate pollen grains (Archbold & Dickins, 1991; McMinn, 1987; Price 1983 Mory et al., 2017). This study corroborates this suggestion, and the implications are that Dulhuntyispora might be unsuitable as a biostratigraphic index taxon for proximal-distal intrabasinal as well as for regional interbasinal correlations. Furthermore, the appearance of the APP5005 subzone taxon Lycopodiumsporites “crassus” in a single section in the Bowen Basin also suggests a highly limited dristribution, which should prompt a rethink of the use of this taxon for biostratigraphic correlation.

7.1.3 The marker mudstone

One task of this thesis was to examine a carbonaceous mudstone horizon in the southern Galilee Basin that is positioned at the interface of the Bandanna and Rewan formations and that was previously interpreted to be a correlative of the marker mudstone in the Bowen Basin (Michaelsen et al., 2000; Sliwa et al., 2017; Phillips et al., 2017a). Palynological data, however, demonstrates that the marker

128 mudstone interval in the Galilee Basin falls into the APP5 zone with the APP6 zone beginning several centimeters above in the overlying Rewan Group. This suggests that the marker mudstone in the Galilee Basinis a distal equivalent of the uppermost Permian coals, rather than a correlative of the marker mudstone in the Bowen Basin, which has an APP6 age assignment.

More significant though might be the distinct changes in palynofacies, as well as the appearance of acritarchs and fungal spores within or just above this unit. The palynofacies shifting from translucent phytoclast-dominated to opaque phytoclast-dominated within the 1 m interval appears to indicate increased input from the hinterland likely due to increased wildfire and subsequent erosion (Sephton et al., 2005). The appearance of algae and acritarchs suggests increased salinity and base level rise, a phenomenon previously also observed in the Sydney Basin (Retallack, 1995). Formation of the marker mudstone was initially proposed as a feature of flooding and compaction of a localised peat deposit (Michaelsen et al., 2000), but the widespread distribution of the feature points to a partly tectonic origin. Foreland loading created a downwarp of the Bowen Basin during the latest Permian, driving base level rise and forming the marker mudstone prior to a huge influx of weathered sediments in the Rewan Group (Baker et al., 1993; Korsch & Totterdall, 2009). The erosion and flooding is likely also strongly influenced by the disappearance of Glossopteris, exposing large areas to weathering and affecting the hydrodynamic cycle significantly (Michaelsen, 2002; Vajda et al., 2020).

7.1.4 Post-EPE Palaeoflora and Palaeoenvironment

This thesis also documents localised ecological and environmental variability post-EPE. The marker mudstone in the Bowen Basin appears to be a lacustrine unit that featured variable anoxia, algal blooms and salinity shifts based on the appearance of algae, acritarchs and pyritic pockmarking damage affecting the palynomorphs. The margins of this lake/pond setting were likely populated by surviving ferns and lycopods as suggested by the high abundances of spores, while the alluvial plains were colonised by peltasperms and voltzialean conifers from the upland and extrabasinal areas (DiMichele et al., 2005; Grauvogel-Stamm & Ash, 2005; Prevec et al., 2010). This post-extinction flora appears to become established very soon after the EPE but there is stronglocalised variability among the palaeoenvironmental indicators and microfloras suggesting variable and unstable conditions in the aftermath of the EPE. The “fungal spike” was only observed in a single locality, Springsure 19, and the magnitude of the spike was relatively low, though other localities in the Bowen Basin have previously demonstrated a pronounced fungal spike (Foster et al., 2002). The appearance of tri- and tetrasaccate pollen grains along with unseparated spore tetrads is rare but consistent and

129 interpreted to represent the onset of higher levels of UV-B radiation which may have acted as one of the kill mechanisms for the Glossopteris flora (Foster & Afonin, 2005; Looy et al., 2005).

The significantly increased degree of weathering, both physical and chemical, can be seen in several pieces of evidence. Retallack & Krull (1999) and Veevers (2004) point to a ‘tuff gap’ in the Early Triassic in eastern Australia and Antarctica as evidence for chemical weathering. Volcanism would likely have been an ongoing process during the Hunter-Bowen Orogeny (265-230 Ma) and detrital zircons have been found in Early Triassic sediments in the intra-arc Gympie Terrane, east of the Bowen Basin (Rosenbaum et al., 2020). High base level may also be a reason for the “tuff gap” as the the ash is falling into standing water of some depth would be quickly dispersed and unrecognisable as a volcanic event. An increase of phytoclasts and charcoal may not just be related to in-situ events like wildfires but may represent recycled debris related to the increased weathering of sediments and coal seams (Eshet et al., 1995; Rampino & Eshet, 2018). The high degree of weathering can also be noted by the sporadic appearance of local low-angle unconformities, disconformities and incisions (Malone et al., 1969; Mollan et al., 1972; Dickins & Malone, 1973; Foster, 1982; Brakel et al., 2009), though no erosional unconformity could be identfied in the sections studied in this work apart from a potential erosional contact in Montani 1.

Palaeosol evidence from the Sydney Basin has been used to suggest a warm, humid climate during the Early Triassic (Retallack et al., 1996; Retallack, 1999). This is consistent with the high occurrence of aquatic lycopsids such as Isoetes beestonii and the survival of ferns, but does not explain the cessation of coal formation. Likely explanations for the coal gap include the extinction of the coal- forming flora, with the surviving species not adapted for the acidic environment of a bog (Retallack et al., 1996) but any attempt at peat formation would be hindered by increased erosion and sedimentary input. Instead, the remnants of the meso- and hydrophytic flora would survive around lakes and ponds as well as rivers/streams flowing into them, with the upland plants becoming established in upriver areas and on the floodplain.

7.1.5 Timing of the Extinction

Chapter 2 of this work describes the nature of the Permian-Triassic Boundary as it is formally defined at the GSSP. When examining the boundary in terrestrial sections there are a number of caveats and challenges one must expect. Without an ashfall tuff that would allow for absolute age dating of the boundary, one relies solely on proxies and correlations. In the case of the Galilee and Bowen basins, the boundary is defined by the cessation of coal deposition and floral turnover that represents the extinction of the Glossopteris flora. Results presented in this thesis point to the floral turnover

130 occurring at the interface of the APP5 and APP6 zones. In all of the studied sections, this transition appears to coincide with the termination of coal occurence. This is in agreement with results from the Sydney Basin (Retallack et al., 1995; Mays et al., 2020a; Vajda et al., 2020), though high-resolution age dates from Fielding et al. (2019) suggest the terrestrial extinction event begins at between 252.60 ± 0.04 and 252.31 ± 0.07, several hundred thousand years before the currently accepted GSSP-defined boundary age of 251.902 ± 0.0241 (Burgess et al., 2014). This agrees with the timing of the terrestrial extinction in low latitudes (Chu et al., 2019). The floral turnover appears to occur rapidly and was observed to be near-synchronous in all studied sections. The initial recovery flora represented by the APP6 zone appears very soon after and appears to comprise of hardy survivors of the lowland and colonisers from upland floras. These colonisers are comprised of Voltzialean conifers and peltaspersms such as Lepidopteris, groups that are rare in Permian Gondwana, but go on to dominate the Triassic floras in the Bowen and Galilee basins along with the corytosperm Dicroidium.

7.1.6 Comparison with wider Gondwana

The Permian-Triassic Boundary in South Africa is generally recognised in the faunal fossil record by the last appearance of therapsid species like Dicynodon and large predators like Rubidgea giving way to much smaller species such as the burrowing dicynodont Lystrosaurus (Retallack et al., 2003; Gastaldo et al., 2005; Botha-Brink, 2017). As significant as the EPE was, in at least some areas, the ecosystem must have been stable enough to allow these faunas to survive. There are also sparse reports of Glossopteris remains in the Early Triassic of India and Antarctica (Pant & Pant, 1987; Holmes, 1992; Lindström & McLoughlin, 2007) though currently no Triassic remains of Glossopteris have been found in eastern Australia. This suggests that certain high latitude localities acted as a potential refuge for the last remnants of Glossopteris, a phenomenon which has also been observed in other plant groups in the Mesozoic and as well (Bomfleur et al., 2018). Palynologically, the Early Triassic in South Africa shows similar trends to eastern Gondwana, with the surviving vegetation also being made up of lycophytes, conifers and peltasperms and the climate also being interpreted as moving from arid in the Permian to more humid in the Triassic (Retallack et al., 2003). The ecology and environment of the Permian-Triassic boundary interval is rarely seen in South America. The hot arid climate of the late Permian and Early Triassic and the oxidising conditions would hinder micro- and macrofossil preservation (Araujo et al., 2016) and may be a contributing factor to the dearth of Permian-Triassic boundary intervals identified in South America.

During the Guadalupian and Lopingian, parts of northern Gondwana featured a diverse mix of elements of the Glossopteris, Cathaysian and Euramerican flora reflected in both palynological and

131 macrofloral datasets (Broutin et al., 1995; Berthelin et al., 2003). This part of Gondwana was situated in the equatorial and subequatorial latitudes and was considered to feature a tropical, warm and humid climate (Berthelin et al., 2003). The corytosperm Dicroidium, a characteristic flora of the Triassic, was identified in Permian sections from the Dead Sea region (Hamad et al., 2008; Blomenkemper et al., 2020). After the extinction of Glossopteris, the new climatic regime and opening of ecological niches allowed for corytosperms and peltasperms previously limited to tropical regions and extrabasinal settings (Pfefferkorn, 1980; DiMichele et al., 2005), to radiate out and occupy the rest of Gondwana (Kerp et al., 2006). Cuticles associated with Dicroidium have been recorded in the upper Permian Chhidru Formation in the Amb Section, , and the fructification was reported in the Raniganj Formation in India (Chandra et al., 2008; Schneebeli-Hermann et al., 2015). The presence of Dicroidium in the mid-latitudes of Permian Gondwana suggests that corytosperms were already radiating out from the tropics to the higher latitudes across the Tethyan margin.

7.2 Future Research

7.2.1 Processing Techniques

While the acid-free processing technique has shown to be effective on the studied samples, future research should focus on refining and improving the technique. The acid-free technique should be tested on a variety of samples such as marine sediments and clay-rich samples from the Early Triassic. For comparison with the conventional technique, the processes should be modified to remove extra variables. Mesh size and type should be consistent between the two processes. Ideally, slides would be made and studied after each step of the process, to determine how each step has affected the assemblage.

7.2.2 Palaeoflora and Palaeoenvironment

Generation and comparison of new palynological and macrofloral datasets is essential to validate the results produced in this work, especially with regards to the relationship between ecology and depositional environment. Palynological assemblages from Permian-aged coal seams in Australia can suffer from high rank and low palynomorph yields but can offer the best chance to preserve these kinds of remains, (Beeston, 1987). Chemical etching of the coals, which enhances cellular detail in polished sections can be done to identify and compare phyteral anatomy to the microflora (e.g. Lapo

132

& Drozdova, 1989; Moore & Swanson, 1993; Van de Wetering et al., 2013b). Special attention should also be paid to marginal environments that feature unique interactions between environmental stress and ecological competition that can lead to significantly different floral communities than those found in the coal-forming mires. This palynological data, when combined with detailed mapping of palaeoshorelines, can be used to assist in further exploration for energy resources in the Bowen and Galilee basins. Of particular use would be the study of core from the margins of the basins, to determine if there are extrabasinal elements entering the assemblage and to assess their transport. A palynological study of algae and acritarchs in the marine sediments of the Bowen Basin would also prove useful in determining their palaeoecology and for comparison to marginal and lacustrine assemblages.

7.2.3 The Permian-Triassic Boundary

The next step in examining the palaeoenvironment of the end-Permian would be to obtain palynological data from the uppermost Permian coal seams, especially in areas where there is no obvious unconformity. Generally, the coals in the Bowen Basin have fairly high rank and low liptinite content, making extracting palynomorphs a difficult endeavour. However, the Leichhardt Seam and equivalents from the western part of the basin where ranks are lower and liptinite content is higher may be an ideal candidate for further testing. Other potential sites for future testing include the Bandanna seams in the Denison Trough or in the south-east of the Taroom Trough. Palynological samples from these localities may provide evidence of environmental instability immediately preceding the end-Permian extinction. On the other hand, they may show that, in the coal-forming mires of eastern Australia, there were no obvious changes to the palaeoflora until the sudden turnover. High resolution sampling for stable carbon isotope geochemistry may assist in examining the relationship between the floral turnover and the boundary. It is imperative to sample multiple localities to establish if the magnitude and position of the excursion is consistent in different parts of the Bowen and Galilee basins, and to consider factors that may affect these values. Well correlated borehole data and absolute age dates from ashfall tuffs, where present, would also assist in providing a spatio-temporal context for the palynological and geochemical data allowing for the development of more accurate models.

133

7.2.4 Biostratigraphy in Eastern Australia

The finding that Dulhuntyispora parvithola may be primarily limited to paralic settings, has significant ramifications for Permian biostratigraphy in Australia. Future studies must focus on further examining the ecological range of D. parvithola and other spores within the genus Dulhuntyispora. This would be done using detailed sampling in a wide variety of localities combined with forming a detailed understanding of the local using sedimentology to form large-scale correlations of its distribution. Factors that would affect its distribution would be the ecological preference of the parent plant, transport, and preservation potential. While D. parvithola will remain an important index taxon, areas where this species would potentially not be found (e.g., intracratonic basins, alluvial plains, basin margins) will require other solutions for biostratigraphic correlation at an intra- and interbasinal scale.

7.3 Summary

• The palynomorph assemblages of the late Permian show observable changes in composition relative to changes in depositional environment and facies, which can influence palynomorph species used as biostratigraphic index taxa. This has implications for interbasinal correlations and correlations on a continental scale with regards to the presence or absence of index taxa and diachroneity in their distribution.

• The marker mudstone is likely not representative of a single large waterbody. Instead, it represents a localised to regional system of lakes. This unit is relatively continuous along the western margin of the Bowen Basin, but varies in thickness. To the south and east it thins out or cannot be observed. It was formed from base level rise as a function of foreland loading during the late Permian and Early Triassic coinciding with the rapid changes in vegetation which influenced weathering and the hydrological cycle that infilled the lowland topography.

• The flora of the earliest Triassic consisted of hardy survivors and pioneers, some of which were extrabasinal colonisers taking advantage of new ecological niches created by the extinction of Glossopteris. These extrabasinal peltasperms and conifers appear to have established themselves in the Bowen Basin very quickly after the EPE. Future studies of the post-extinction ecosystem need to take this into account and should examine multiple sections to avoid extrapolating data from a single local assemblage.

134

References

Abbink, O. A. (1998). Palynological Investigations in the Jurassic of the North Sea Region. LPP Contribution Series, 8. 192 pp.

Abbink, O., Targarona, J., Brinkhuis, H., & Visscher, H. (2001). Late Jurassic to earliest Cretaceous palaeoclimatic evolution of the southern North Sea. Global and Planetary Change, 30(3-4), 231-256.

Abbink, O., Van Konijnenburg-Van Cittert, J., & Visscher, H. (2004). A sporomorph ecogroup model for the Northwest European Jurassic-Lower Cretaceous: concepts and framework. Netherlands Journal of Geosciences, 83(1), 17-38.

Afonin, S. A., Barinova, S. S., & Krassilov, V. A. (2001). A bloom of Tympanicysta Balme ( of zygnematalean affinities) at the Permian-Triassic boundary. Geodiversitas, 23(4), 481-487.

Algeo, T. J., Kuwahara, K., Sano, H., Bates, S., Lyons, T., Elswick, E., Hinnov, L., Ellwood, B., Moser, J., & Maynard, J. B. (2011). Spatial variation in sediment fluxes, redox conditions, and productivity in the Permian–Triassic Panthalassic Ocean. Palaeogeography, Palaeoclimatology, Palaeoecology, 308(1-2), 65-83.

Allen, J. P., & Fielding, C. R. (2007). Sequence architecture within a low-accommodation setting: An example from the Permian of the Galilee and Bowen basins, Queensland, Australia. AAPG Bulletin, 91(11), 1503-1539.

Anderson, J. M. (1977). The biostratigraphy of the Permian and Triassic. Part 3. A review of Gondwana Permian palynology with particular reference to the northern Karoo Basin, South Africa. Memoirs of the Botanical Survey of South Africa, 41, 1-188.

Anderson, J. M., Anderson, H., Archangelsky, S., Bamford, M., Chandra, S., Dettmann, M., Hill, R., McLoughlin, S., & Rösler, O. (1999). Patterns of Gondwana plant colonisation and diversification. Journal of African Earth Sciences, 28(1), 145-167.

Angiolini, L., Checconi, A., Gaetani, M., & Rettori, R. (2010). The latest Permian mass extinction in the Alborz Mountains (north Iran). Geological Journal, 45(2‐3), 216-229.

Araújo, R. N., Nogueira, A. C. R., Bandeira, J., & Angélica, R. S. (2016). Shallow lacustrine system of the Permian Pedra de Fogo Formation, Western Gondwana, Parnaíba Basin, Brazil. Journal of South American Earth Sciences, 67, 57-70.

135

Archangelsky, S., & Cúneo, R. (1990). Polyspermophyllum, a new Permian gymnosperm from Argentina, with considerations about the Dicranophyllales. Review of Palaeobotany and Palynology, 63(1-2), 117-135.

Archbold, N., & Dickins, J. (1989). Australian Phanerozoic Timescales 6. Permian. A standard for the Permian System in Australia. Bureau Mineral Resources, Geology Geophysics, Record, 1989/36, 1-17.

Ayaz, S., Esterle, J., & Martin, M. (2015). Spatial variation in the stratigraphic architecture of the Fort Cooper and equivalent coal measures, Bowen Basin, Queensland. Australian Journal of Earth Sciences, 62(5), 547-562.

Ayaz, S., Martin, M., Esterle, J., Amelin, Y., & Nicoll, R. (2016a). Age of the Yarrabee and accessory tuffs: implications for the upper Permian sediment-accumulation rates across the Bowen Basin. Australian Journal of Earth Sciences, 63(7), 843-856.

Ayaz, S., Rodrigues, S., Golding, S., & Esterle, J. (2016b). Compositional variation and palaeoenvironment of the volcanolithic Fort Cooper Coal Measures, Bowen Basin, Australia. International Journal of Coal Geology, 166, 36-46.

Backhouse, J. (1990). Permian palynostratigraphic correlations in south-western Australia and their geological implications. Review of Palaeobotany and Palynology, 65(1-4), 229-237.

Backhouse, J. (1993). Palynology and correlation of Permian sediments in the Perth, Collie and Officer basins. Geological Survey of Western Australia Report, 34, 111-128.

Baker, J. (1997). Green ferric clay in non-marine sandstones of the Rewan Group, southern Bowen Basin, eastern Australia. Clay Minerals, 32(4), 499-506.

Baker, J. C., Fielding, C. R., De Caritat, P., & Wilkinson, M. M. (1993). Permian evolution of sandstone composition in a complex back-arc extensional to foreland basin; the Bowen Basin, eastern Australia. Journal of Sedimentary Research, 63(5), 881-893.

Balme, B. E. (1995). Fossil in situ spores and pollen grains: an annotated catalogue. Review of Palaeobotany and Palynology, 87(2-4), 81-323.

Balme, B. E., & Helby, R. J. (1973). Floral Modifications at the Permian-Triassic Boundary in Australia. In A. Logan & L. V. Hills (Eds.), The Permian and Triassic Systems and their Mutual Boundary. (Vol. 2, pp. 433-443): Memoirs - Canadian Society of .

136

Barbolini, N., Bamford, M. K., & Rubidge, B. (2016). demonstrates that Permian spore‐pollen zones of Australia and South Africa are diachronous. Gondwana Research, 37, 241-251.

Barbolini, N., Rubidge, B., & Bamford, M. K. (2018). A new approach to biostratigraphy in the Karoo retroarc foreland system: Utilising restricted-range palynomorphs and their first appearance datums for correlation. Journal of African Earth Sciences, 140, 114-133.

Baresel, B., Bucher, H., Brosse, M., Cordey, F., Guodun, K., & Schaltegger, U. (2017). Precise age for the Permian–Triassic boundary in South China from high-precision U-Pb and Bayesian age–depth modeling. Solid Earth, 8(2), 361-378.

Becker, L., Poreda, R., Basu, A., Pope, K., Harrison, T., Nicholson, C., & Iasky, R. (2004). Bedout: a possible end-Permian impact crater offshore of northwestern Australia. Science, 304(5676), 1469-1476.

Beerling, D. J., Harfoot, M., Lomax, B., & Pyle, J. A. (2007). The stability of the stratospheric ozone layer during the end-Permian eruption of the Siberian Traps. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 365(1856), 1843-1866.

Beeston, J. (1987). Aspects of inertinite formation and deposition in the Denison Trough, Queensland. Australian Coal Geology, 7, 33-45.

Beeston, J. W. (1986). Coal rank variation in the Bowen Basin, Queensland. International Journal of Coal Geology, 6(2), 163-179.

Beeston, J. W. (1990). Cyclodendron leslii (Seward) Kräusel 1928 and associated palynomorphs in the Early Permian Reids Dome beds, Queensland, Australia. , 14(4), 325-330.

Benca, J. P., Duijnstee, I. A., & Looy, C. V. (2018). UV-B–induced forest sterility: Implications of ozone shield failure in Earth’s largest extinction. Science Advances, 4(2), e1700618.

Benton, M. J., & Twitchett, R. J. (2003). How to kill (almost) all life: the end-Permian extinction event. Trends in Ecology & Evolution, 18(7), 358-365.

Benton, M. J., Tverdokhlebov, V. P., & Surkov, M. (2004). Ecosystem remodelling among vertebrates at the Permian–Triassic boundary in Russia. Nature, 432(7013), 97-100.

Bercovici, A., & Vajda, V. (2016). Terrestrial Permian–Triassic boundary sections in South China. Global and Planetary Change, 143, 31-33.

137

Bercovici, A., Cui, Y., Forel, M.-B., Yu, J., & Vajda, V. (2015). Terrestrial paleoenvironment characterization across the Permian–Triassic boundary in South China. Journal of Asian Earth Sciences, 98, 225-246.

Bernardi, M., Petti, F. M., Kustatscher, E., Franz, M., Hartkopf-Fröder, C., Labandeira, C. C., Wappler, T., van Konijnenburg-van Cittert, J. H., Peecook, B. R., & Angielczyk, K. D. (2017). Late Permian (Lopingian) terrestrial ecosystems: a global comparison with new data from the low-latitude Bletterbach Biota. Earth-Science Reviews, 175, 18-43.

Berner, R. A. (2002). Examination of hypotheses for the Permo–Triassic boundary extinction by carbon cycle modeling. Proceedings of the National Academy of Sciences, 99(7), 4172- 4177.

Berthelin, M., Broutin, J., Kerp, H., Crasquin-Soleau, S., Platel, J.-P., & Roger, J. (2003). The Oman Gharif mixed paleoflora: a useful tool for testing Permian Pangea reconstructions. Palaeogeography, Palaeoclimatology, Palaeoecology, 196(1-2), 85-98.

Blomenkemper, P., Kerp, H., Hamad, A. A., & Bomfleur, B. (2020). Contributions towards whole- plant reconstructions of Dicroidium plants (Umkomasiaceae) from the Permian of Jordan. Review of Palaeobotany and Palynology, 104210.

Bolch, C. (1997). The use of sodium polytungstate for the separation and concentration of living dinoflagellate cysts from marine sediments. Phycologia, 36(6), 472-478.

Bomfleur, B., Blomenkemper, P., Kerp, H., & McLoughlin, S. (2018). Polar regions of the Mesozoic–Paleogene greenhouse world as refugia for relict plant groups. In Transformative (pp. 593-611). Academic Press.

Bomfleur, B., Mörs, T., Unverfärth, J., Liu, F., Läufer, A., Castillo, P., Changhwan, O., Tae-Yoon, S. P., Woo, W., & Crispini, L. (2020). Uncharted Permian to Jurassic continental deposits in the far north of Victoria Land, East Antarctica. Journal of the Geological Society, 178(1), jgs2020-062.

Bond, D. P., & Wignall, P. B. (2010). Pyrite framboid study of marine Permian–Triassic boundary sections: a complex anoxic event and its relationship to contemporaneous mass extinction. GSA Bulletin, 122(7-8), 1265-1279.

Botha-Brink, J. (2017). Burrowing in Lystrosaurus: preadaptation to a postextinction environment? Journal of Vertebrate , 37(5), e1365080.

138

Bowring, S. A., Erwin, D., Jin, Y., Martin, M., Davidek, K., & Wang, W. (1998). U/Pb zircon geochronology and tempo of the end-Permian mass extinction. Science, 280(5366), 1039- 1045.

Brain, T. J., McKellar, J. L., & Carmichael, D. C. (1991). Stratigraphic drilling report, GSQ Muttaburra 1. Geological Survey of Queensland, Record 1991/25.

Brakel, A., Totterdell, J., Wells, A., & Nicoll, M. (2009). Sequence stratigraphy and fill history of the Bowen Basin, Queensland. Australian Journal of Earth Sciences, 56(3), 401-432.

Brinkhuis, H., Bujak, J., Smit, J., Versteegh, G., & Visscher, H. (1998). Dinoflagellate-based sea surface temperature reconstructions across the Cretaceous–Tertiary boundary. Palaeogeography, Palaeoclimatology, Palaeoecology, 141(1-2), 67-83.

Broutin, J., Roger, J., Platel, J.-P., Angiolini, L., Baud, A., Bucher, H., Marcoux, J., & Hasmi, H. A. (1995). The Permian Pangea. Phytogeographic implications of new discoveries in Oman (). Comptes Rendus de l'Académie des Sciences de Paris, Série II a, 321, 1069-1086.

Brown, C. (2008). Palynological techniques. In J. B. Riding & S. Warny (Eds.), AASP Special Publications (2nd Ed. ed., pp. 137). Dallas, Texas, U.S.A: American Association of Stratigraphic Palynologists Foundation.

Burgess, S. D., Bowring, S., & Shen, S.. (2014). High-precision timeline for Earth’s most severe extinction. Proceedings of the National Academy of Sciences, 111(9), 3316-3321.

Campbell, I., Czamanske, G., Fedorenko, V., Hill, R., & Stepanov, V. (1992). Synchronism of the Siberian Traps and the Permian-Triassic boundary. Science, 258(5089), 1760-1763.

Chaloner, W., & Muir, M. (1968). Spores and floras. In: Murchison, DG and Westall, TS (eds.) Coal and coal-bearing strata. (pp. 127-146). Oliver & Boyd. Edinburgh.

Chandra, S., Singh, K. J., & Jha, N. (2008). First report of the fertile plant genus Umkomasia from Late Permian beds in India and its biostratigraphic significance. Palaeontology, 51(4), 817- 826.

Chu, D., Tong, J., Benton, M. J., Yu, J., & Huang, Y. (2019). Mixed continental-marine biotas following the Permian-Triassic mass extinction in South and North China. Palaeogeography, Palaeoclimatology, Palaeoecology, 519, 95-107.

Claoué-Long, J., Zichao, Z., Guogan, M., & Shaohua, D. (1991). The age of the Permian-Triassic boundary. Earth and Planetary Science Letters, 105(1-3), 182-190.

139

Colbath, G. K., & Grenfell, H. R. (1995). Review of biological affinities of Paleozoic acid- resistant, organic-walled eukaryotic algal microfossils (including “acritarchs”). Review of Palaeobotany and Palynology, 86(3-4), 287-314.

Collinson, J. W., Hammer, W. R., Askin, R. A., & Elliot, D. H. (2006). Permian-Triassic boundary in the central Transantarctic Mountains, Antarctica. Geological Society of America Bulletin, 118(5-6), 747-763.

Conolly, J. R., & Ferm, J. C. (1971). Permo-Triassic sedimentation patterns, Sydney Basin, Australia. AAPG bulletin, 55(11), 2018-2032.

Creech, M. (2002). Tuffaceous deposition in the Newcastle Coal Measures: challenging existing concepts of peat formation in the Sydney Basin, New South Wales, Australia. International Journal of Coal Geology, 51(3), 185-214.

Cúneo, N. R. (1996). Permian phytogeography in Gondwana. Palaeogeography, Palaeoclimatology, Palaeoecology, 125(1-4), 75-104.

Dai, S., Bechtel, A., Eble, C. F., Flores, R. M., French, D., Graham, I. T., Hood, M. M., Hower, J. C., Korasidis, V. A., & Moore, T. A. (2020). Recognition of peat depositional environments in coal: A review. International Journal of Coal Geology, 219, 103383.

De Jersey, N. J. (1946). Microspore types in some Queensland Permian coals. Department of Geology, University of Queensland Papers, New Series 3(5), 1-12.

De Jersey, N. J., & McKellar, J. L. (1979). Composition of palynofloras and taxonomic notes on selected species from GSQ Hughenden 1-2R, 5 and 6 (northern Galilee Basin). Geological Survey of Queensland, Record 1979/28.

De Jersey, N. J., & McKellar, J. L. (1981). Triassic palynology of the Warang Sandstone (northern Galilee Basin) and its phytogeographic implications. In M. M. Cresswell & P. Vella (Eds.), Gondwana Five - Selected papers and abstracts of papers presented at the Fifth International Gondwana Symposium. Proceedings of the Fifth International Gondwana Symposium (pp. 31-37), Wellington, New Zealand, A.A. Balkema, Rotterdam.

De Vernal, A., Rochon, A., Turon, J.-L., & Matthiessen, J. (1997). Organic-walled dinoflagellate cysts: palynological tracers of sea-surface conditions in middle to high latitude marine environments. Geobios, 30(7), 905-920. de Wit, M. J., Ghosh, J. G., de Villiers, S., Rakotosolofo, N., Alexander, J., Tripathi, A., & Looy, C. (2002). Multiple organic carbon isotope reversals across the Permo-Triassic boundary of

140

terrestrial Gondwana sequences: clues to extinction patterns and delayed ecosystem recovery. The Journal of Geology, 110(2), 227-240.

Devernrscu, E. (1998). The chlorococcalean alga Botryococcus and its significance in hydrocarbon exploration. Paper presented at the Workshop on" Modern and Ancient Sedimentary Environments and Processes" in Moeciu.

Diaz, R. J., & Rosenberg, R. (1995). Marine benthic hypoxia: a review of its ecological effects and the behavioural responses of benthic macrofauna. Oceanography and Marine Biology. An Annual Review, 33, 245-203.

Dickins, J. M. (1964). PEC Warrong 1. Well Completion Report: Appendix 3. 9 pp.

Dickins, J. M., & Malone, E. J. (1973). Geology of the Bowen Basin, Queensland Bureau of Mineral Resources, Geology and Geophysics, Australia, Bulletin 130, 1-154

Diessel, C. (1982). An appraisal of coal facies based on maceral characteristics. Australian Coal Geology, 4(2), 474-484.

Diessel, C. F. (2012). Coal-bearing depositional systems: Springer Science & Business Media.

DiMichele, W. A., & Phillips, T. L. (2002). The ecology of Paleozoic ferns. Review of Palaeobotany and Palynology, 119(1-2), 143-159.

DiMichele, W. A., Kerp, H., Krings, M., & Chaney, D. (2005). The Permian peltasperm radiation: evidence from the southwestern United States. New Mexico Museum of Natural History and Science Bulletin, 30, 67-79.

Dolby, J. H., & Balme, B. (1976). Triassic palynology of the Carnarvon Basin, western Australia. Review of Palaeobotany and Palynology, 22(2), 105-168.

Draper, J. (2013). Bowen basin. In Geology of Queensland (Vol. 5, pp. 371-384): Geological Survey of Queensland Brisbane, QLD.

Dulhunty, J. A. (1945). Principal microspore types in the Permian coals of New South Wales. Proceedings of th e Linnean Society of New South Wales, 70, 147-157.

Dulhunty, J. A. (1946). Distribution of microspore types in New South Wales Permian coalfields. Proceedings of the Linnean Society of New South Wales, 71, 239-251.

Dulhunty, J. A., & Dulhunty, R. (1949). Notes on microspore types in Tasmanian Permian coals. Proceedings of the Linnean Society of New South Wales, 74, 132-139.

141

Eble, C. F., Pierce, B. S., & Grady, W. C. (2003). Palynology, petrography and geochemistry of the Sewickley coal bed (Monongahela Group, Late Pennsylvanian), northern Appalachian Basin, USA. International Journal of Coal Geology, 55(2-4), 187-204.

Elliot, D. H., Fanning, C. M., Isbell, J. L., & Hulett, S. R. (2017). The Permo-Triassic Gondwana sequence, central Transantarctic Mountains, Antarctica: zircon geochronology, , and basin evolution. Geosphere, 13(1), 155-178.

Elsik, W. C. (1999). Reduviasporonites Wilson 1962: Synonymy of the fungal organism involved in the late Permian crisis. Palynology, 23(1), 37-41.

Erwin, D. (1990). The end-Permian mass extinction. Annual Review of Ecology and Systematics, 21(1), 69-91.

Erwin, D. (1994). The Permo–Triassic extinction. Nature, 367(6460), 231-236..

Erwin, D. H. (1993). The great Paleozoic crisis: life and death in the Permian: Columbia University Press.

Erwin, D. H. (1998). After the end: recovery from extinction. Science, 279(5355), 1324-1325.

Erwin, D. H., Bowring, S. A., & Yugan, J. (2002). End-Permian mass extinctions: a review. Special Papers-Geological Society of America, 363-384.

Escapa, I. H., Taylor, E. L., Cuneo, N.R., Bomfleur, B., Bergene, J., Serbet, R., & Taylor, T. N. (2011). Triassic floras of Antarctica: plant diversity and distribution in high paleolatitude communities. Palaios, 26(9), 522-544.

Eshet, Y., Rampino, M. R., & Visscher, H. (1995). Fungal event and palynological record of ecological crisis and recovery across the Permian-Triassic boundary. Geology, 23(11), 967- 970.

Evans, P. R. (1969). Upper Carboniferous and Permian palynological stages and their distribution in eastern Austalia. In A. J. Amos (Ed.), Gondwana Stratigraphy (pp. 41-54). Paris: UNESCO.

Exon, N. F. (Cartographer). (1970). 1:250 000 geological series – explanatory notes Tambo, Queensland.

Falcon-Lang, H. J., & Bashforth, A. R. (2005). Morphology, anatomy, and upland ecology of large cordaitalean trees from the Middle Pennsylvanian of Newfoundland. Review of Palaeobotany and Palynology, 135(3-4), 223-243.

142

Falkner, A., & Fielding, C. (1993). Quantitative facies analysis of coal-bearing sequences in the Bowen Basin, Australia: applications to reservoir description. The geologic modelling of hydrocarbon reservoirs and outcrop analogues, 15, 81-98.

Feist-Burkhardt, S., Götz, A. E., Ruckwied, K., & Russell, J. W. (2008). Palynofacies patterns, acritarch diversity and stable isotope signatures in the Lower Muschelkalk (Middle Triassic) of N Switzerland: Evidence of third-order cyclicity. Swiss Journal of Geosciences, 101(1), 1-15.

Fielding, C. (1984). A coal depositional model for the Durham Coal Measures of NE England. Journal of the Geological Society, 141(5), 919-931.

Fielding, C. R., Frank, T. D., McLoughlin, S., Vajda, V., Mays, C., Tevyaw, A. P., Winguth, A., Winguth, C., Nicoll, R. S., & Bocking, M. (2019). Age and pattern of the southern high- latitude continental end-Permian extinction constrained by multiproxy analysis. Nature Communications, 10(1), 1-12.

Fielding, C. R., Sliwa, R., Holcombe, R. J., & Kassan, J. (2000). A new palaeogeographic synthesis of the Bowen Basin of central Queensland. In Bowen Basin Symposium (pp. 287-302). Geological Society of Australia Brisbane.Filatoff, J. (1985). AAR Yellowbank 2. Well Completion Report: Appendix 4. 8 pp.

Fielding, C., Falkner, A., & Scott, S. (1993). Fluvial response to foreland basin overfilling; the late Permian Rangal Coal Measures in the Bowen basin, Queensland, Australia. Sedimentary Geology, 85(1-4), 475-497.

Fielding, C., Sliwa, R., Holcombe, R., & Jones, A. (2001). A new palaeogeographic synthesis for the Bowen, Gunnedah and Sydney Basins of eastern Australia. In: Hill, K. C. & Bernecker, T. (eds) Eastern Australasian Basins Symposium. Petroleum Exploration Society of Australia Special Publication, 269–278.

Filatoff, J., & Price, P. L. (1991a). ATP375P "P3c" Palynological Study on BON Spring Grove 1, UOD Bainbilla 1, HEP Bloodwood 1, OLA Beranga South 1 and OLA Bottletree 1. Report No. 264/5, 1-9.

Filatoff, J., & Price, P. L. (1991b). EPP 337, A Northern Denison Trough (Springsure) Palynostratigraphical Reference Section. Report No. 606/2, 1-7.

Foster, C. (1982). Spore-pollen assemblages of the Bowen Basin, Queensland (Australia): their relationship to the Permian/Triassic boundary. Review of Palaeobotany and Palynology, 36(1-2), 165-183.

143

Foster, C. B. (1979). Permian plant microfossils of the Blair Athol Coal Measures, Baralaba Coal Measures, and basal Rewan Formation of Queensland. Geological Survey of Queensland Publications, 372(Palaeontological Papers 45), 1-244.

Foster, C., & Afonin, S. (2005). Abnormal pollen grains: an outcome of deteriorating atmospheric conditions around the Permian–Triassic boundary. Journal of the Geological Society, 162(4), 653-659.

Foster, C., & Archbold, N. W. (2001). Chronologic anchor points for the Permian and Early Triassic of the Eastern Australian Basins. Contributions to Geology and Palaeontology of Gondwana in Honour of Helmut Wopfner. Geological Institute, University of Cologne, Germany, 175-197.

Foster, C., Stephenson, M., Marshall, C., Logan, G., & Greenwood, P. (2002). A revision of Reduviasporonites Wilson 1962: description, illustration, comparison and biological affinities. Palynology, 26(1), 35-58.

Gastaldo, R. A. (1992). Regenerative growth in fossil horsetails following burial by alluvium. Historical Biology, 6(3), 203-219.

Gastaldo, R. A., Adendorff, R., Bamford, M., Labandeira, C. C., Neveling, J., & Sims, H. (2005). Taphonomic trends of macrofloral assemblages across the Permian–Triassic boundary, Karoo Basin, South Africa. Palaios, 20(5), 479-497.

Gastaldo, R. A., Kamo, S. L., Neveling, J., Geissman, J. W., Bamford, M., & Looy, C. V. (2015). Is the vertebrate-defined Permian-Triassic boundary in the Karoo Basin, South Africa, the terrestrial expression of the end-Permian marine event? Geology, 43(10), 939-942.

Gastaldo, R. A., Neveling, J., Clark, C. K., & Newbury, S. S. (2009). The terrestrial Permian- Triassic boundary event bed is a nonevent. Geology, 37(3), 199-202.

Gastaldo, R. A., Neveling, J., Looy, C. V., Bamford, M. K., Kamo, S. L., & Geissman, J. W. (2017). Paleontology of the Blaauwater 67 and 65 Farms, South Africa: Testing the Daptocephalus/Lystrosaurus Biozone Boundary in a stratigraphic framework. Palaios, 32(6), 349-366.

Gedl, P., & Ziaja, J. (2012). Palynofacies from Bathonian (Middle Jurassic) ore-bearing clays at Gnaszyn, Kraków-Silesia Homocline, Poland, with special emphasis on sporomorph eco- groups. Acta Geologica Polonica, 62(3), 325-349.

144

Gee, C. T. (1989). Permian Glossopteris and Elatocladus megafossil floras from the English coast, eastern Ellsworth Land, Antarctica. Antarctic Science, 1(1), 35-44.

Glikson, A. (2004). Bedout basement rise, offshore northwestern Australia: evidence of an unshocked mafic volcanic hyaloclastite volcanic breccia. AGUFM, 2004, B33C-0269.

Gorter, J. D. (1996). Speculation on the Origin of the Bedout High-A Large, Circular Structure of Pre Mesozoic Age in the Offshore Canning Basin, Western Australia.

Gorter, J., Nicoll, R. S., Metcalfe, I., Willink, R., & Ferdinando, D. (2009). The Permian–Triassic boundary in Western Australia: evidence from the Bonaparte and Northern Perth basins— exploration implications. The APPEA Journal, 49(1), 311-336.

Gould, R., & Delevoryas, T. (1977). The biology of Glossopteris: evidence from petrified seed- bearing and pollen-bearing organs. Alcheringa, 1(4), 387-399.

Gradstein, F. M., Ogg, J. G., Schmitz, M., & Ogg, G. (2012). The Geologic Time Scale 2012: Elsevier. 1176 pp.

Grauvogel-Stamm, L., & Ash, S. R. (2005). Recovery of the Triassic land flora from the end- Permian life crisis. Comptes Rendus Palevol, 4(6-7), 593-608.

Grech, P. V. J. W. (2001). Sedimentology and sequence stratigraphy of the Early Triassic Rewan Group, Bowen Basin. (Unpublished PhD Thesis). University of Adelaide.

Grech, P., & Dyson, I. (1997). An integrated approach to the study of the Early Triassic Rewan Group, Bowen Basin. The APPEA Journal, 37(1), 192-204.

Green, P. M., McKellar, J. L., Carmichael, D. C., & Smith, R. J. (1991). Stratigraphic drilling report, GSQ Longreach 1-1B. Geological Survey of Queensland, Record 1991/20.

Grenfell, H. R. (1995). Probable fossil zygnematacean algal spore genera. Review of Palaeobotany and Palynology, 84(3-4), 201-220.

Grice, K., Cao, C., Love, G. D., Böttcher, M. E., Twitchett, R. J., Grosjean, E., Summons, R. E., Turgeon, S. C., Dunning, W., & Jin, Y. (2005a). Photic zone euxinia during the Permian- Triassic superanoxic event. Science, 307(5710), 706-709.

Grice, K., Nabbefeld, B., & Maslen, E. (2007). Source and significance of selected polycyclic aromatic hydrocarbons in sediments (Hovea-3 well, Perth Basin, Western Australia) spanning the Permian–Triassic boundary. Organic Geochemistry, 38(11), 1795-1803.

145

Grice, K., Twitchett, R. J., Alexander, R., Foster, C., & Looy, C. (2005b). A potential biomarker for the Permian–Triassic ecological crisis. Earth and Planetary Science Letters, 236(1-2), 315-321.

Grime, J. P. (1979). Plant strategies and vegetation processes. Chichester, UK: John Wiley and Sons. 456 pp.

Grimm, E. C. (2004). TGView Version 2.0.2. Springfield: Illinois State Museum.

Gulbranson, E., Isbell, J., Taylor, E., Ryberg, P., Taylor, T., & Flaig, P. (2012). Permian polar forests: deciduousness and environmental variation. Geobiology, 10(6), 479-495.

Guy-Ohlson, D. (1992). Botryococcus as an aid in the interpretation of palaeoenvironment and depositional processes. Review of Palaeobotany and Palynology, 71(1-4), 1-15.

Guy-Ohlson, D., & Lindström, S. (1994). Palaeoecology of the Early Permian strata at Heimefrontfjella, Dronning Maud Land, Antarctica. Antarctic Science, 6, 507-507.

Hagemann, H. W., & Wolf, M. (1987). New interpretations of the facies of the Rhenish brown coal of West Germany. International Journal of Coal Geology, 7(4), 335-348.

Hallam, A., & Wignall, P. B. (1997). Mass extinctions and their aftermath: Oxford University Press, UK. 328 pp.

Hamad, A. A., Kerp, H., Vörding, B., & Bandel, K. (2008). A late Permian flora with Dicroidium from the Dead Sea region, Jordan. Review of Palaeobotany and Palynology, 149(3-4), 85- 130.

Hansen, J., & Uroda, A. (2018). Reinterpretation of Wire-line Log Data in the Eastern Galilee Basin, Queensland: Stratigraphical and Hydrogeological Implications. ASEG Extended Abstracts, 2018(1), 1-8.

Helby, R. J. (1970). A biostratigraphy of the Late Permian and Triassic of the Sydney Basin. (Unpublished PhD Thesis). University of Sydney.

Helby, R. J. (1973). Review of Late Permian and Triassic palynology of New South Wales. Geological Society of Australia Special Publication, 4, 141-155.

Helby, R., Morgan, R., & Partridge, A. (1987). A palynological zonation of the Australian Mesozoic. Association of Australasian Palaeontologists Memoir, 4, 1-94.

146

Hobday, D. (1987). Gondwana coal basins of Australia and South Africa: tectonic setting, depositional systems and resources. Geological Society, London, Special Publications, 32(1), 219-233.

Hochuli, P. A. (2016). Interpretation of “fungal spikes” in Permian-Triassic boundary sections. Global and Planetary Change, 144, 48-50.

Hochuli, P. A., Hermann, E., Vigran, J. O., Bucher, H., & Weissert, H. (2010). Rapid demise and recovery of plant ecosystems across the end-Permian extinction event. Global and Planetary Change, 74(3-4), 144-155.

Holdgate, G., McLoughlin, S., Drinnan, A., Finkelman, R., Willett, J., & Chiehowsky, L. (2005). Inorganic chemistry, petrography and palaeobotany of Permian coals in the Prince Charles Mountains, East Antarctica. International Journal of Coal Geology, 63(1-2), 156-177.

Holmes, W. K. (1992). Glossopteris-like leaves from the Triassic of eastern Australia. Geophytology, 22, 119-125.

Holser, W. T., Schönlaub, H.-P., Attrep, M., Boeckelmann, K., Klein, P., Magaritz, M., Orth, C. J., Fenninger, A., Jenny, C., & Kralik, M. (1989). A unique geochemical record at the Permian/Triassic boundary. Nature, 337(6202), 39-44.

Hopkins, J., & McCarthy, F. (2002). A laboratory experiment studying the effects of progressive oxidation on palynological assemblages in marine sediments. Palynology, 26, 167-184.

Huleatt, M. (1991). Handbook of Australian Black Coals: Geology, Resources, Seam Properties, and Product Specifications. Bureau of Mineral Resources, Australia, Resource Report, 7. Canberra, ACT: Australian Government Publishing Service, 116 pp.

I’Anson, A., Deighton, I., Müller, R. D., Dutkiewicz, A., & Heine, C. (2018). Burial and exhumation history of the Galilee Basin, Australia: Implications for unconventional hydrocarbon prospectivity. AAPG Bulletin, 102(3), 483-507.

Irmis, R. B., & Whiteside, J. H. (2012). Delayed recovery of non-marine tetrapods after the end- Permian mass extinction tracks global carbon cycle. Proceedings of the Royal Society B: Biological Sciences, 279(1732), 1310-1318.

Isbell, J. L., Askin, R. A., & Retallack, G. J. (1999). Search for evidence of impact at the Permian- Triassic boundary in Antarctica and Australia: Comment and reply. Geology, 27(9), 859- 860.

147

Jell, P. A. (Ed.). (2013). Geology of Queensland. Geological Survey of Queensland, Brisbane. 970 pp.

Jensen, A. R. (1975). Permo-Triassic stratigraphy and sedimentation in the Bowen Basin, Queensland. Bureau of Mineral Resources, Australia, Geology and Geophysics Bulletin, 154. Canberra, ACT: Australian Government Publishing Service, 187 pp.

Jin, Y., Wang, Y., Wang, W., Shang, Q., Cao, C., & Erwin, D. (2000). Pattern of marine mass extinction near the Permian-Triassic boundary in South China. Science, 289(5478), 432- 436.

Jones, M. J. (1986). Palynological Laboratory Report No. 291/6, GSQ Eddystone No. 5.

Jones, M., & Truswell, E. (1992). Late Carboniferous and Early Permian palynostratigraphy of the Joe Joe Group, southern Galilee Basin, Queensland, and implications for Gondwanan stratigraphy. BMR Journal of Australian Geology and Geophysics, 13(2), 143-185.

Kaiho, K., & Koga, S. (2013). Impacts of a massive release of methane and hydrogen sulfide on oxygen and ozone during the late Permian mass extinction. Global and Planetary Change, 107, 91-101.

Kaiho, K., Kajiwara, Y., Nakano, T., Miura, Y., Kawahata, H., Tazaki, K., Ueshima, M., Chen, Z., & Shi, G. R. (2001). End-Permian catastrophe by a bolide impact: evidence of a gigantic release of sulfur from the mantle. Geology, 29(9), 815-818.

Kemp, E., Balme, B., Helby, R., Kyle, R., Playford, G., & Price, P. (1977). Carboniferous and Permian palynostratigraphy in Australia and Antarctica: a review. BMR Journal of Australian Geology and Geophysics, 2(3), 177-208.

Kerp, H., Hamad, A. A., Vörding, B., & Bandel, K. (2006). Typical Triassic Gondwanan floral elements in the Upper Permian of the paleotropics. Geology, 34(4), 265-268.

Knepprath, N. E. (2006). Late Permian forests of the Buckley Formation, Beardmore Glacier area, Antarctica. (Unpublished MSc Thesis). Vanderbilt University.

Korasidis, V. A., Wallace, M. W., Wagstaff, B. E., Holdgate, G. R., Tosolini, A.-M. P., & Jansen, B. (2016). Cyclic floral succession and fire in a Cenozoic wetland/peatland system. Palaeogeography, Palaeoclimatology, Palaeoecology, 461, 237-252.

Korsch, R., & Totterdell, J. (2009). Subsidence history and basin phases of the Bowen, Gunnedah and Surat Basins, eastern Australia. Australian Journal of Earth Sciences, 56(3), 335-353.

148

Korte, C., & Kozur, H. W. (2010). Carbon-isotope stratigraphy across the Permian–Triassic boundary: a review. Journal of Asian Earth Sciences, 39(4), 215-235.

Kump, L. R., Pavlov, A., & Arthur, M. A. (2005). Massive release of hydrogen sulfide to the surface ocean and atmosphere during intervals of oceanic anoxia. Geology, 33(5), 397-400.

Kustatscher, E., van Konijnenburg-van Cittert, J. H., & Roghi, G. (2010). Macrofloras and palynomorphs as possible proxies for palaeoclimatic and palaeoecological studies: A case study from the Pelsonian (Middle Triassic) of Kühwiesenkopf/Monte Prà della Vacca (Olang Dolomites, N-Italy). Palaeogeography, Palaeoclimatology, Palaeoecology, 290(1- 4), 71-80.

Kutzbach, J., & Ziegler, A. (1993). Simulation of Late Permian climate and biomes with an atmosphere-ocean model: comparisons with observations. Philosophical Transactions of the Royal Society of London. Series B: Biological Sciences, 341(1297), 327-340.

Langenhorst, F., Kyte, F. T., & Retallack, G. J. (2005). Reexamination of quartz grains from the Permian-Triassic boundary section at Graphite Peak, Antarctica. Proceedings of the 36th Lunar and Planetary Science Conference, abstract 2358.

Lapo, A. V., & Drozdova, I. N. (1989). Phyterals of humic coals in the USSR. International Journal of Coal Geology, 12(1-4), 477-510.

Laurie, J., Bodorkos, S., Nicoll, R., Crowley, J., Mantle, D., Mory, A., Wood, G., Backhouse, J., Holmes, E., & Smith, T. (2016). Calibrating the middle and late Permian palynostratigraphy of Australia to the geologic time-scale via U–Pb zircon CA-IDTIMS dating. Australian Journal of Earth Sciences, 63(6), 701-730.

Lei, Y., Servais, T., & Feng, Q. (2013). The diversity of the Permian . Review of Palaeobotany and Palynology, 198, 145-161.

Leipe, C., Kobe, F., & Müller, S. (2019). Testing the performance of sodium polytungstate and lithium heteropolytungstate as non-toxic dense media for pollen extraction from lake and peat sediment samples. Quaternary International, 516, 207-214.

Lele, K. M., Maithy, P. K., & Mandal, J. (1981). In situ spores from Lower Gondwana ferns: Their morphology and variation. The Palaeobotanist, 28-29, 128-154.

Li, L., Wang, Y., Liu, Z., Zhou, N., & Wang, Y. (2016). Late Triassic palaeoclimate and palaeoecosystem variations inferred by palynological record in the northeastern Sichuan Basin, China. PalZ, 90(2), 327-348.

149

Lignum, J., Jarvis, I., & Pearce, M. A. (2008). A critical assessment of standard processing methods for the preparation of palynological samples. Review of Palaeobotany and Palynology, 149(3-4), 133-149.

Lindsay, J. (2000). Source rock potential and algal-matter abundance, Cooper Basin, South Australia. PIRSA Report Book, 32.

Lindström, S. (2005). Palynology of Permian shale, clay and sandstone clasts from the Basen till in northern Vestfjella, Dronning Maud Land. Antarctic Science, 17(1), 87-96.

Lindström, S., & McLoughlin, S. (2007). Synchronous palynofloristic extinction and recovery after the end-Permian event in the Prince Charles Mountains, Antarctica: implications for palynofloristic turnover across Gondwana. Review of Palaeobotany and Palynology, 145(1-2), 89-122.

Lindström, S., McLoughlin, S., & Drinnan, A. N. (1997). Intraspecific variation of taeniate bisaccate pollen within Permian glossopterid sporangia, from the Prince Charles Mountains, Antarctica. International Journal of Plant Sciences, 158(5), 673-684.

Littler, K., Robinson, S. A., Bown, P. R., Nederbragt, A. J., & Pancost, R. D. (2011). High sea- surface temperatures during the Early Cretaceous Epoch. Nature Geoscience, 4(3), 169- 172.

Liu, K., Boult, P., Painter, S., & Paterson, L. (1996). Outcrop analog for sandy braided stream reservoirs: permeability patterns in the Triassic Hawkesbury Sandstone, Sydney Basin, Australia. AAPG bulletin, 80(12), 1850-1866.

Looy, C. V. (2000). The Permian-Triassic biotic crisis: collapse and recovery of terrestrial ecosystems. LPP Contribution Series, 13. 114 pp.

Looy, C. V., Collinson, M. E., Cittert, J. H. V. K.-V., Visscher, H., & Brain, A. P. (2005). The ultrastructure and botanical affinity of end-Permian spore tetrads. International Journal of Plant Sciences, 166(5), 875-887.

Looy, C. V., Twitchett, R. J., Dilcher, D. L., Van Konijnenburg-Van Cittert, J. H., & Visscher, H. (2001). Life in the end-Permian dead zone. Proceedings of the National Academy of Sciences, 98(14), 7879-7883.

Looy, C., Brugman, W., Dilcher, D., & Visscher, H. (1999). The delayed resurgence of equatorial forests after the Permian–Triassic ecologic crisis. Proceedings of the National Academy of Sciences, 96(24), 13857-13862.

150

Lundgren, K. R., Cúneo, N. R., Escapa, I. H., & Tomescu, A. M. (2019). A new marattialean fern from the lower Permian of Patagonia (Argentina) with cautionary tales on synangial morphology and pinnule base characters. International Journal of Plant Sciences, 180(7), 667-680.

Luo, G., Wang, Y., Grice, K., Kershaw, S., Algeo, T. J., Ruan, X., Yang, H., Jia, C., & Xie, S. (2013). Microbial–algal community changes during the latest Permian ecological crisis: Evidence from lipid biomarkers at Cili, South China. Global and Planetary Change, 105, 36-51.

Malone, E. J., Olgers, F., & Kirkegaard, A. (1969). The geology of the Duaringa and Saint- Lawrence 1: 250,000 sheet areas, Queensland (Vol. 121): Bureau of Mineral Resources, Geology and Geophysics.

Maruoka, T., Koeberl, C., Hancox, P., & Reimold, W. (2003). Sulfur geochemistry across a terrestrial Permian–Triassic boundary section in the Karoo Basin, South Africa. Earth and Planetary Science Letters, 206(1-2), 101-117.

Matheson, S. (1990). Coal geology and exploration in the Rangal Coal Measures, north Bowen Basin. Brisbane: Department of Resource Industries, Queensland.

Maxwell, W. D. (1992). Permian and Early Triassic extinction of non-marine tetrapods. Palaeontology, 35(3), 571-583.

Mays, C., Vajda, V., & McLoughlin, S. (2021). Permian–Triassic non-marine algae of Gondwana—Distributions, natural affinities and ecological implications. Earth-Science Reviews, 212, 103382.

Mays, C., Vajda, V., Frank, T. D., Fielding, C. R., Nicoll, R. S., Tevyaw, A. P., & McLoughlin, S. (2020). Refined Permian− Triassic floristic timeline reveals early collapse and delayed recovery of south polar terrestrial ecosystems. Geological Society of America Bulletin, 132(7-8), 1489-1513.

McCartain, E., Backhouse, J., Haig, D., Balme, B., & Crawly, M. K. (2006). Gondwana-related Late Permian palynoflora, foraminifers and lithofacies from the Wailuli Valley, Timor Leste. Neues Jahrbuch für Geologie und Paläontologie - Abhandlungen 240(1), 53-80.

McIldowie, M., & Alexander, R. (2005). Identification of a novel C33 n-alkylcyclohexane biomarker in Permian–Triassic sediments. Organic Geochemistry, 36(10), 1454-1458.

151

McKellar, J. L. (1975). Palynology of samples from QDM Aramac 1 and Hexham 1. In C. F. J. Swarbrick & C. I. Wallin (Eds.), Well completion report–QDM Aramac 1 and Hexham 1 (Vol. Report 92, Appendix 5, pp. 31-42): Geological Survey of Queensland.

McKellar, J. L. (1977a). Miospore assemblages from GSQ Hughenden 2R, 3-4R, 5 and 6. Geological Survey of Queensland, Record 1977/15.

McKellar, J. L. (1977b). Palynostratigraphy of core samples from the Hughenden 1:250 000 Sheet area, northern Galilee and Eromanga Basins. Queensland Government Mining Journal, 78(910), 393-399.

McKellar, J. L. (1978). Analysis of organic residues from GSQ Hughenden 7. Geological Survey of Queensland, Record 1978/41.

McKellar, J. L. (1979). Palynostratigraphy of core samples from GSQ Hughenden 7. Queensland Government Mining Journal 80(932), 295-302.

McKellar, J. L., Gray, A., Glenn, K. C., & Nicoll, R. S. (1999). Cooper, Eromanga and Galilee Basins Biozonation and Stratigraphy, 1999, Chart 19. Canberra: Geoscience Australia.

McKillop, M. D. (2016). Coal Maturity trends within the Bowen Basin. An update of vitrinite reflectance stude by Beeston (1986). Queensland Geological Record, 016/04.

McLoughlin, S. (1988). Geology of the Inglis Dome, Denison Trough, central Queensland. Department of Geology, University of Queensland, 12, 229-263.

McLoughlin, S. (1990). Some Permian glossopterid fructifications and leaves from the Bowen Basin, Queensland, Australia. Review of Palaeobotany and Palynology, 62(1-2), 11-40.

McLoughlin, S. (1992). Late Permian megafossils from the Bowen Basin, Queensland, Australia: Part1. Palaeontographica Abteilung B, 228, 105-149.

McLoughlin, S. (1993). Plant fossil distributions in some Australian Permian non-marine sediments. Sedimentary Geology, 85(1-4), 601-619.

McLoughlin, S. (1994). Late Permian plant megafossils from the Bowen Basin, Queensland, Australia: Part 2. Palaeontographica Abteilung B, 231, 1-1.

McLoughlin, S. (2011a). Glossopteris–insights into the architecture and relationships of an iconic Permian Gondwanan plant. Journal of the Botanical Society of Bengal, 65(2), 93-106.

McLoughlin, S. (2011b). New records of leaf galls and oviposition scars in Permian– Triassic Gondwanan gymnosperms. Australian Journal of , 59(2), 156-169.

152

McLoughlin, S., Drinnan, A. N., Slater, B. J., & Hilton, J. (2015). Paurodendron stellatum: a new Permian permineralized herbaceous lycopsid from the Prince Charles Mountains, Antarctica. Review of Palaeobotany and Palynology, 220, 1-15.

McLoughlin, S., Maksimenko, A., & Mays, C. (2019). A new high-paleolatitude late Permian permineralized peat flora from the Sydney Basin, Australia. International Journal of Plant Sciences, 180(6), 513-539.

McMinn, A. (1987). Palynostratigraphy of the Stroud-gloucester trough, NSW. Alcheringa, 11(2), 151-164.

Mertens, K. N., Verhoeven, K., Verleye, T., Louwye, S., Amorim, A., Ribeiro, S., Deaf, A. S., Harding, I. C., De Schepper, S., & González, C. (2009). Determining the absolute abundance of dinoflagellate cysts in recent marine sediments: the Lycopodium marker- grain method put to the test. Review of palaeobotany and palynology, 157(3-4), 238-252.

Metcalfe, I., Crowley, J., Nicoll, R., & Schmitz, M. (2015). High-precision U-Pb CA-TIMS calibration of Middle Permian to Lower Triassic sequences, mass extinction and extreme climate-change in eastern Australian Gondwana. Gondwana Research, 28(1), 61-81.

Metcalfe, I., Nicoll, R., & Willink, R. (2008). Conodonts from the Permian–Triassic transition in Australia and position of the Permian–Triassic boundary. Australian Journal of Earth Sciences, 55(3), 365-377.

Metcalfe, I., Nicoll, R., Crowley, J., Ives, M., Mantle, D., Ruming, K., Huyskens, M., & Foster, C. (2011). New Middle Permian-Early Triassic U-Pb zircon CA-IDTIMS isotopic ages of tuffs in the Sydney Basin, Australia: International calibration of stratigraphy and biostratigraphy. Paper presented at the Programme & Abstracts: The XVII International Congress on the Carboniferous and Permian.

Meyer, K., Kump, L., & Ridgwell, A. (2008). Biogeochemical controls on photic-zone euxinia during the end-Permian mass extinction. Geology, 36(9), 747-750.

Michaelsen, P. (2002). Mass extinction of peat-forming plants and the effect on fluvial styles across the Permian–Triassic boundary, northern Bowen Basin, Australia. Palaeogeography, Palaeoclimatology, Palaeoecology, 179(3-4), 173-188.

Michaelsen, P., Henderson, R. A., Crosdale, P. J., & Mikkelsen, S. O. (2000). Facies architecture and depositional dynamics of the Upper Permian Rangal coal measures, Bowen Basin, Australia. Journal of Sedimentary Research, 70(4), 879-895.

153

Miller, M. F., Knepprath, N. E., Cantrill, D. J., Francis, J. E., & Isbell, J. L. (2016). Highly productive polar forests from the Permian of Antarctica. Palaeogeography, Palaeoclimatology, Palaeoecology, 441, 292-304.

Millsteed, B. D. (1997). Late Carboniferous and Permian palynology of the Galilee Basin, Central Queensland, Australia. (Unpublished PhD Thesis). The University of Queensland,

Mishra, S., Jha, N., Stebbins, A., Brookfield, M., & Hannigan, R. (2019). Palaeoenvironments, flora, and organic carbon and nitrogen isotope changes across the non-marine Permian- Triassic boundary at Wybung Head, Australia. Palaeogeography, Palaeoclimatology, Palaeoecology, 534, 109292.

Mollan, R., Forbes, V., Jensen, A., Exon, N., & Gregory, C. (1972). Geology of the Eddystone and Taroon 1: 250 000 Sheet areas and the western part of the Mundubbera 1: 250 000 Sheet area. Bur. Miner. Resour. Geol. Geophys. Aust., Rep., 142.

Moore, T., & Swanson, K. (1993). Application of etching and SEM in the identification of fossil plant tissues in coal. Organic Geochemistry, 20(6), 769-777.

Morante, R. (1996). Permian and early Triassic isotopic records of carbon and strontium in Australia and a scenario of events about the Permian‐Triassic boundary. Historical Biology, 11(1-4), 289-310.

Mory, A. J., & Backhouse, J. (1997). Permian stratigraphy and palynology of the Carnarvon Basin, Western Australia: Geological Survey of Western Australia.

Mory, A., Crowley, J., Backhouse, J., Nicoll, R., Bryan, S., López Martínez, M., & Mantle, D. (2017). Apparent conflicting Roadian–Wordian (middle Permian) CA-IDTIMS and palynology ages from the Canning Basin, Western Australia. Australian Journal of Earth Sciences, 64(7), 889-901.

Moss, P. T. (2013). Palynology and its application to geomorphology. In: Shroder, J.F., Switzer, A.D., Kennedy, D.M. (Eds.), Treatise on Geomorphology, 14. Academic Press, San Diego, CA, Methods in Geomorphology, pp. 315–325.

Moss, P. T., & Kershaw, A. P. (2007). A late Quaternary marine palynological record (oxygen isotope stages 1 to 7) for the humid tropics of northeastern Australia based on ODP Site 820. Palaeogeography, Palaeoclimatology, Palaeoecology, 251(1), 4-22.

154

Moss, P. T., Greenwood, D. R., & Archibald, S. B. (2005). Regional and local vegetation community dynamics of the Eocene Okanagan Highlands (British Columbia Washington State) from palynology. Canadian Journal of Earth Sciences, 42(2), 187-204.

Moss, P. T., Smith, R. Y., & Greenwood, D. R. (2016). A window into mid-latitudinal Early Eocene environmental variability: a high-resolution palynological analysis of the Falkland site, Okanagan Highlands, British Columbia, Canada. Canadian Journal of Earth Sciences, 53(6), 605-613.

Mundil, R., Ludwig, K. R., Metcalfe, I., & Renne, P. R. (2004). Age and timing of the Permian mass extinctions: U/Pb dating of closed-system zircons. Science, 305(5691), 1760-1763.

Munsterman, D., & Kerstholt, S. (1996). Sodium polytungstate, a new non-toxic alternative to bromoform in heavy liquid separation. Review of palaeobotany and palynology, 91(1-4), 417-422.

Murdoch, C. (2010). Palynostratigraphic analysis, Yebna 1, PL 232, Bowen Basin. Brisbane, Qld: Santos Stratigraphic Services, Palynology Report, 34.

Murray, C. (1983). Permian geology of Queensland. Paper presented at the Permian Geology of Queensland: Proceedings of the Symposium on the Permian Geology of Queensland, Geological Society of Australia Queensland Division, pp. 1-32.

Mutton, A. J. (2003). Queensland Coals (Vol. 1): Queensland Department of Natural Resources and Mines. 116 pp.

Neves, R., & Sullivan, H. (1964). Modification of fossil spore exines associated with the presence of pyrite crystals. , 10(4), 443-452.

Nichols, D. J. (1995). The role of palynology in paleoecological analyses of Tertiary coals. International Journal of Coal Geology, 28(2-4), 139-159.

Nichols, D. J., & Warwick, P. (2005). Palynology in coal systems analysis-The key to floras, climate, and stratigraphy of coal-forming environments. Special Papers-Geological Society of America, 387, 51.

Nicoll, R., McKellar, J., Ayaz, S. A., Laurie, J., Esterle, J., Crowley, J., Wood, G., & Bodorkos, S. (2015). CA-IDTIMS dating of tuffs, calibration of palynostratigraphy and stratigraphy of the Bowen and Galilee basins. Paper presented at the Bowen Basin Symposium.

155

Nishida, H., Pigg, K. B., Kudo, K., & Rigby, J. F. (2007). New evidence of reproductive organs of Glossopteris based on permineralized fossils from Queensland, Australia. I. Ovulate organ Homevaleia gen. nov. Journal of plant research, 120(4), 539-549.

Nishida, H., Pigg, K. B., Kudo, K., & Rigby, J. F. (2014). New evidence of the reproductive organs of Glossopteris based on permineralized fossils from Queensland, Australia. II: pollen-bearing organ Ediea gen. nov. Journal of plant research, 127(2), 233-240.

Norton, D. A. (1989). Floristics and structure of mire-forest ecotones, west coast South Island, New Zealand. Journal of the Royal Society of New Zealand, 19(1), 31-42.

Norton, D. A. (1991). Seedling and sapling distribution patterns in a coastal podocarp forest, Hokitika Ecological District, New Zealand. New Zealand Journal of Botany, 29(4), 463- 466.

Norvick, M. (1981). Permian and Late Carboniferous palynostratigraphy of the Galilee Basin, Queensland: Bureau of Mineral Resources, Geology and Geophysics. 156 pp.

Nowak, H., Schneebeli-Hermann, E., & Kustatscher, E. (2019). No mass extinction for land plants at the Permian–Triassic transition. Nature communications, 10(1), 1-8.

O’Keefe, J., & Eble, C. (2012). A comparison of HF-based and non-HF-based palynology processing techniques in clay-rich lignites from the Claiborne Group, upper Mississippi Embayment, United States. Palynology, 36(1), 116-130.

O'Brien, C. L., Robinson, S. A., Pancost, R. D., Damsté, J. S. S., Schouten, S., Lunt, D. J., Alsenz, H., Bornemann, A., Bottini, C., & Brassell, S. C. (2017). Cretaceous sea-surface temperature evolution: Constraints from TEX86 and planktonic foraminiferal oxygen isotopes. Earth-Science Reviews, 172, 224-247.

Ogg, J. G., Ogg, G. M., & Gradstein, F. M. (2016). A concise geologic time scale: 2016: Elsevier.

Pant, D. D., & Pant, R. (1987). Some Glossopteris leaves from Indian Triassic beds. Palaeontographica Abteilung B Paläophytologie, 205(1-6), 165-178.

Paten, R. (1969). Palynologic contributions to petroleum exploration in the Permian formations of the Cooper Basin, Australia. The APPEA Journal, 9(1), 79-87.

Peng, Y., Zhang, S., Yu, T., Yang, F., Gao, Y., & Shi, G. R. (2005). High-resolution terrestrial Permian–Triassic eventostratigraphic boundary in western Guizhou and eastern Yunnan, southwestern China. Palaeogeography, Palaeoclimatology, Palaeoecology, 215(3-4), 285- 295.

156

Pfefferkorn, H. W., & Resnik, D. A. (1980). The occurrence of Sphenopteris germanica (pteridosperm, Autunian, Lower Permian) in . Journal of Paleontology, 81- 84.

Phillips, L., Crowley, J., Mantle, D., Esterle, J., Nicoll, R., McKellar, J., & Wheeler, A. (2018a). U–Pb geochronology and palynology from Lopingian (upper Permian) coal measure strata of the Galilee Basin, Queensland, Australia. Australian Journal of Earth Sciences, 65(2), 153-173.

Phillips, L., Edwards, S., Bianchi, V., & Esterle, J. (2017a). Paleo-environmental reconstruction of Lopingian (upper Permian) sediments in the Galilee Basin, Queensland, Australia. Australian Journal of Earth Sciences, 64(5), 58a-609.

Phillips, L., Esterle, J., & Edwards, S. (2017b). Review of Lopingian (upper Permian) stratigraphy of the Galilee Basin, Queensland, Australia. Australian Journal of Earth Sciences, 64(3), 283-300.

Phillips, L., Verdel, C., Allen, C., & Esterle, J. (2018b). Detrital zircon U–Pb geochronology of Permian strata in the Galilee Basin, Queensland, Australia. Australian Journal of Earth Sciences, 65(4), 465-481.

Phillips, T. L., & Peppers, R. A. (1984). Changing patterns of Pennsylvanian coal-swamp vegetation and implications of climatic control on coal occurrence. International Journal of Coal Geology, 3(3), 205-255.

Phillips, T. L., Peppers, R. A., Avcin, M. J., & Laughnan, P. F. (1974). Fossil plants and coal: patterns of change in Pennsylvanian coal swamps of the Illinois Basin. Science, 184(4144), 1367-1369.

Pickering, S. A. (1985a). AAR Arcturus 5. Well Completion Report: Appendix 6. 13 pp.

Pickering, S. A. (1985b). Palynological Report on AAR Christmas Creek 1, ATP 337, (Report no. 246/1). 30 pp.

Pickering, S. A. (1985c). Palynological Report on AAR Merivale 1. ATP 337, (Report no. 270/1). 15 pp.

Pigg, K. B., & McLoughlin, S. (1997). Anatomically preserved Glossopteris leaves from the Bowen and Sydney basins, Australia. Review of Palaeobotany & Palynology, 97(3-4), 339- 359.

157

Powis, G. (2009). Sequence stratigraphy of the Yebna 1 core, 1110–1320 m. Noosa Heads, Qld: Geodelta Pty Ltd (unpublished report).

Prevec, R., Gastaldo, R. A., Neveling, J., Reid, S. B., & Looy, C. V. (2010). An autochthonous glossopterid flora with latest Permian palynomorphs and its depositional setting in the Dicynodon Assemblage Zone of the southern Karoo Basin, South Africa. Palaeogeography, Palaeoclimatology, Palaeoecology, 292(3-4), 391-408.

Prevec, R., Labandeira, C. C., Neveling, J., Gastaldo, R. A., Looy, C. V., & Bamford, M. (2009). Portrait of a Gondwanan ecosystem: a new late Permian fossil locality from KwaZulu- Natal, South Africa. Review of Palaeobotany and Palynology, 156(3-4), 454-493.

Price, P. (1976). Permian palynology of the Bowen Basin. A guide to the geology of the Bowen and Surat Basins in Queensland. Excursion Guide C, 3.

Price, P. (1983). A Permian palynostratigraphy for Queensland. Paper presented at the Permian Geology of Queensland: Proceedings of the Symposium on the Permian Geology of Queensland, Geological Society of Australia Queensland Division, pp. 155-211.

Price, P. (1997). Permian to Jurassic palynostratigraphic nomenclature of the Bowen and Surat Basins. The Surat and Bowen Basins, south-east Queensland. Queensland Minerals and Energy Review Series, Queensland Department of Mines and Energy, Brisbane, 137, 178.

Price, P. L. (1984). AAR Myrtleville 1. Well Completion Report: Appendix 5. 5 pp.

Price, P., & Filatoff, J. (1990). Application of morphological lineages in Australian palynostratigraphy. Review of palaeobotany and palynology, 65(1-4), 195-207.

Price, P., Filatoff, J., Williams, A., Pickering, S., & Wood, G. (1985). Late Palaeozoic and Mesozoic palynostratigraphic units. CSR Oil and Gas Division Palynology Facility Report 274/25. Unpublished report held by the Queensland Department of Natural Resources and Mines as CR14012.

Purcell, R. (2011). LRH Saltern 1-1A. Well Completion Report. 238 pp.

Rabalais, N. N., Turner, R. E., & Wiseman Jr, W. J. (2002). Gulf of Mexico hypoxia, aka “The dead zone”. Annual Review of Ecology and Systematics, 33(1), 235-263.

Rampino, M. R., & Eshet, Y. (2018). The fungal and acritarch events as time markers for the latest Permian mass extinction: an update. Geoscience Frontiers, 9(1), 147-154.

158

Rampino, M. R., Prokoph, A., & Adler, A. (2000). Tempo of the end-Permian event: High- resolution cyclostratigraphy at the Permian-Triassic boundary. Geology, 28(7), 643-646.

Raup, D. M. (1979). Size of the Permo-Triassic bottleneck and its evolutionary implications. Science, 206(4415), 217-218.

Raymond, A., Lambert, L., Costanza, S., Slone, E. J., & Cutlip, P. C. (2010). Cordaiteans in paleotropical wetlands: an ecological re-evaluation. international Journal of Coal geology, 83(2-3), 248-265.

Reichow, M. K., Pringle, M., Al'Mukhamedov, A., Allen, M., Andreichev, V., Buslov, M., Davies, C., Fedoseev, G., Fitton, J., & Inger, S. (2009). The timing and extent of the eruption of the Siberian Traps large igneous province: Implications for the end-Permian environmental crisis. Earth and Planetary Science Letters, 277(1-2), 9-20.

Reid, J. (1929). Geology of the Bowen River Coalfield. Queensland Geological Survey Publication, 276, 1-107.

Reid, J. H. (1930). Geological Survey Reports. Geology of the Springsure district, Parts 1 and 2. Queensland Government Mining Journal, 31, 87-157.

Renne, P. R., Black, M. T., Zichao, Z., Richards, M. A., & Basu, A. R. (1995). Synchrony and causal relations between Permian-Triassic boundary crises and Siberian flood volcanism. Science, 269(5229), 1413-1416.

Renne, P. R., Melosh, H. J., Farley, K. A., & Reimold, W. U. (2004). Is Bedout an Impact Crater? Take 2/Response. Science, 306(5696), 610-612.

Retallack, G. (1980). Late Carboniferous to Middle Triassic megafossil floras from the Sydney Basin. Geological Survey New South Wales, Bulletin, 26, 384-430.

Retallack, G. (1995). Permian-Triassic life crisis on land. Science, 267(5194), 77-80.

Retallack, G. (2005). Earliest Triassic claystone breccias and soil-erosion crisis. Journal of Sedimentary Research, 75(4), 679-695.

Retallack, G. J. (1997). Earliest Triassic origin of Isoetes and quillwort evolutionary radiation. Journal of Paleontology, 500-521.

Retallack, G. J. (1999). Postapocalyptic greenhouse paleoclimate revealed by earliest Triassic paleosols in the Sydney Basin, Australia. Geological Society of America Bulletin, 111(1), 52-70.

159

Retallack, G. J. (2002). Lepidopteris callipteroides, an earliest Triassic seed fern of the Sydney Basin, southeastern Australia. Alcheringa, 26(4), 475-500.

Retallack, G. J., & Krull, E. (1999). Landscape ecological shift at the Permian‐Triassic boundary in Antarctica. Australian Journal of Earth Sciences, 46(5), 785-812.

Retallack, G. J., & Krull, E. S. (2006). Carbon isotopic evidence for terminal-Permian methane outbursts and their role in extinctions of animals, plants, coral reefs, and peat swamps. Special Papers: Geological Society of America, 399, 249-268.

Retallack, G. J., Seyedolali, A., Krull, E. S., Holser, W. T., , C. P., & Kyte, F. T. (1998). Search for evidence of impact at the Permian-Triassic boundary in Antarctica and Australia. Geology, 26(11), 979-982.

Retallack, G. J., Sheldon, N. D., Carr, P. F., Fanning, M., Thompson, C. A., Williams, M. L., Jones, B. G., & Hutton, A. (2011). Multiple Early Triassic greenhouse crises impeded recovery from Late Permian mass extinction. Palaeogeography, Palaeoclimatology, Palaeoecology, 308(1-2), 233-251.

Retallack, G. J., Smith, R. M., & Ward, P. D. (2003). Vertebrate extinction across Permian– Triassic boundary in Karoo Basin, South Africa. GSA Bulletin, 115(9), 1133-1152.

Retallack, G. J., Veevers, J. J., & Morante, R. (1996). Global coal gap between Permian–Triassic extinction and Middle Triassic recovery of peat-forming plants. Geological Society of America Bulletin, 108(2), 195-207.

Riding, J. B., & Kyffin-Hughes, J. E. (2004). A review of the laboratory preparation of palynomorphs with a description of an effective non-acid technique. Revista Brasileira de Paleontologia, 7(1), 13-44.

Riding, J. B., & Kyffin‐Hughes, J. E. (2006). Further testing of a non‐acid palynological preparation procedure. Palynology, 30(1), 69-87.

Riding, J. B., & Kyffin-Hughes, J. E. (2011). A direct comparison of three palynological preparation techniques. Review of Palaeobotany and Palynology, 167(3-4), 212-221.

Riding, J. B., Kyffin-Hughes, J. E., & Owens, B. (2007). An effective palynological preparation procedure using hydrogen peroxide. Palynology, 31(1), 19-36.

Rigby, J. (1993). Review of the Early Permian flora of the Nychum Volcanics north of Chillagoe, north Queensland. In: R.H. Findlay, R. Unrug, M.R. Banks, J.J. Veevers (Eds.), Gondwana Eight: Assembly, Evolution and Dispersal, A.A. Balkema, Rotterdam (1993), pp. 241-247.

160

Rigby, J. F., & Hekel, H. (1977). Palynology of the Permian sequence in the Springsure anticline, central Queensland. Geological Survey of Queensland Publications, 363 (Palaeontological Papers 37), 1-76.

Roscher, M., Stordal, F., & Svensen, H. (2011). The effect of global warming and global cooling on the distribution of the latest Permian climate zones. Palaeogeography, Palaeoclimatology, Palaeoecology, 309(3-4), 186-200.

Rosenbaum, G., Slade, A., & Hoy, D. (2020). Sedimentological responses to the Hunter–Bowen Orogeny (eastern Australia): evidence from the northern Gympie Terrane. Australian Journal of Earth Sciences, 67(1), 59-73.

Rößler, R., & Galtier, J. (2003). The first evidence of the fern Botryopteris from the Permian of the Southern Hemisphere reflecting growth form diversity. Review of Palaeobotany and Palynology, 127(1-2), 99-124.

Rothwell, G. W., & Stockey, R. A. (2008). Phylogeny and evolution of ferns: a paleontological perspective. Biology and evolution of ferns and lycophytes, 332, 366.

Royer, D. (2001). Stomatal density and stomatal index as indicators of paleoatmospheric CO2 concentration. Review of Palaeobotany and Palynology, 114(1-2), 1-28.

Ruckwied, K., Götz, A. E., Pálfy, J., & Török, Á. (2008). Palynology of a terrestrial coal-bearing series across the Triassic/Jurassic boundary (Mecsek Mts, Hungary). Central European Geology, 51(1), 1-15.

Ryberg, P. E., Taylor, E. L., & Taylor, T. N. (2012). Permineralized lycopsid from the Permian of Antarctica. Review of Palaeobotany and Palynology, 169, 1-6.

Saunders, A., & Reichow, M. (2009). The Siberian Traps and the End-Permian mass extinction: a critical review. Chinese Science Bulletin, 54(1), 20-37.

Schneebeli-Hermann, E., Kürschner, W. M., Kerp, H., Bomfleur, B., Hochuli, P. A., Bucher, H., Ware, D., & Roohi, G. (2015). Vegetation history across the Permian–Triassic boundary in Pakistan (Amb section, Salt Range). Gondwana Research, 27(3), 911-924.

Scott, A. C. (1991). An introduction to the application of palaeobotany and palynology to coal geology. Bulletin de la Société Géologique de France, 162(2), 145-153.

Segroves, K. L. (1970). Permian spores and pollen grains from the Perth Basin, Western Australia. Grana, 10(1), 43-73.

161

Sephton, M. A., Jiao, D., Engel, M. H., Looy, C. V., & Visscher, H. (2015). Terrestrial acidification during the end-Permian biosphere crisis? Geology, 43(2), 159-162.

Sephton, M. A., Looy, C. V., Brinkhuis, H., Wignall, P. B., De Leeuw, J. W., & Visscher, H. (2005). Catastrophic soil erosion during the end-Permian biotic crisis. Geology, 33(12), 941-944.

Sephton, M. A., Looy, C. V., Veefkind, R. J., Brinkhuis, H., De Leeuw, J. W., & Visscher, H. (2002). Synchronous record of d13C shifts in the and atmosphere at the end of the Permian. Geological Society of America Special Papers, 356, 455-462.

Sephton, M. A., Visscher, H., Looy, C. V., Verchovsky, A. B., & Watson, J. S. (2009). Chemical constitution of a Permian-Triassic disaster species. Geology, 37(10), 875-878.

Shen, S. Z., Crowley, J. L., Wang, Y., Bowring, S. A., Erwin, D. H., Sadler, P. M., Cao, C.-q., Rothman, D. H., Henderson, C. M., & Ramezani, J. (2011). Calibrating the end-Permian mass extinction. Science, 334(6061), 1367-1372.

Shen, W., Lin, Y., Xu, L., Li, J., Wu, Y., & Sun, Y. (2007). Pyrite framboids in the Permian– Triassic boundary section at Meishan, China: evidence for dysoxic deposition. Palaeogeography, Palaeoclimatology, Palaeoecology, 253(3-4), 323-331.

Shi, G., Waterhouse, J., & McLoughlin, S. (2010). The Lopingian of Australasia: a review of biostratigraphy, correlations, palaeogeography and palaeobiogeography. Geological Journal, 45(2‐3), 230-263.

Six, J., Schultz, P., Jastrow, J., & Merckx, R. (1999). Recycling of sodium polytungstate used in soil organic matter studies. Soil Biology & Biochemistry, 31(8), 1193-1196.

Slater, B. J., McLoughlin, S., & Hilton, J. (2012). Animal–plant interactions in a middle Permian permineralised peat of the Bainmedart Coal Measures, Prince Charles Mountains, Antarctica. Palaeogeography, Palaeoclimatology, Palaeoecology, 363, 109-126.

Slater, B. J., McLoughlin, S., & Hilton, J. (2015). A high-latitude Gondwanan lagerstätte: the Permian permineralised peat biota of the Prince Charles Mountains, Antarctica. Gondwana Research, 27(4), 1446-1473.

Slater, S. M., & Wellman, C. H. (2015). A quantitative comparison of dispersed spore/pollen and plant megafossil assemblages from a Middle Jurassic plant bed from Yorkshire, UK. , 41(4), 640-660.

162

Sliwa, R., Esterle, J., Phillips, L., & Wilson, S. (2017). Rangal Supermodel 2015: The Rangal- Baralaba-Bandanna Coal Measures in the Bowen and Galilee Basins. Final Report ACARP Project C22028.

Smith, R. M., & Botha-Brink, J. (2014). Anatomy of a mass extinction: sedimentological and taphonomic evidence for drought-induced die-offs at the Permo-Triassic boundary in the main Karoo Basin, South Africa. Palaeogeography, Palaeoclimatology, Palaeoecology, 396, 99-118.

Smith, R. M., & Ward, P. D. (2001). Pattern of vertebrate extinctions across an event bed at the Permian-Triassic boundary in the Karoo Basin of South Africa. Geology, 29(12), 1147- 1150.

Smith, T. E., & Mantle, D. J. (2013). Late Permian palynozones and associated CA-IDTIMS dated tuffs from the Bowen Basin, Australia (1922201847).

Smith, T., Nicoll, R., Laurie, J., Crowley, J., McKellar, J., Campbell, H., Raine, I., Mantle, D., & Mory, A. (2018). Recalibrating Australian Triassic Palynostratigraphy to the International Geologic Timescale Using High Resolution CA-IDTIMS Dating. ASEG Extended Abstracts, 2018(1), 1-1.

Spina, A., Cirilli, S., Utting, J., & Jansonius, J. (2015). Palynology of the Permian and Triassic of the Tesero and Bulla sections (Western Dolomites, Italy) and consideration about the enigmatic species Reduviasporonites chalastus. Review of Palaeobotany and Palynology, 218, 3-14.

Srivastava, A. K., Krassilov, V. A., & Agnihotri, D. (2011). Peltasperms in the Permian of India and their bearing on Gondwanaland reconstruction and climatic interpretation. Palaeogeography, Palaeoclimatology, Palaeoecology, 310(3-4), 393-399.

Srivastava, S. C., Srivastaba, A. K., Bhattacharyya, A. P., & Tewari, R. (1999). Degraded Permian palynomorphs from North-East Himalaya, India. Permophiles, 33, 32-36.

Staines, H. R. E. (1972). Blackwater Coalfield; correlation of seams in the Rangal Coal Measures, Mackenzie River to Sirius Creek (Vol. 70). Brisbane, Queensland, Australia: Geological Survey of Queensland.

Steiner, M. B., Eshet, Y., Rampino, M. R., & Schwindt, D. M. (2003). Fungal abundance spike and the Permian–Triassic boundary in the Karoo Supergroup (South Africa). Palaeogeography, Palaeoclimatology, Palaeoecology, 194(4), 405-414.

163

Stemmerik, L., Bendix-Almgreen, S. E., & Piasecki, S. (2001). The Permian–Triassic boundary in central East Greenland: past and present views. Bulletin of the Geological Society of Denmark, 48(2), 159-167.

Svensen, H., Planke, S., Polozov, A. G., Schmidbauer, N., Corfu, F., Podladchikov, Y. Y., & Jamtveit, B. (2009). Siberian gas venting and the end-Permian environmental crisis. Earth and Planetary Science Letters, 277(3-4), 490-500.

Taylor, E. L., Taylor, T. N., & Collinson, J. W. (1989). Depositional setting and paleobotany of Permian and Triassic permineralized peat from the central Transantarctic Mountains, Antarctica. International Journal of Coal Geology, 12(1-4), 657-679.

Taylor, E. L., Taylor, T. N., & Collinson, J. W. (1989). Depositional setting and paleobotany of Permian and Triassic permineralized peat from the central Transantarctic Mountains, Antarctica. International Journal of Coal Geology, 12(1-4), 657-679.

Taylor, E. L., Taylor, T. N., & Cúneo, N. R. (1992). The present is not the key to the past: a polar forest from the Permian of Antarctica. Science, 257(5077), 1675-1677.

Thomas, B., Willink, R., Grice, K., Twitchett, R., Purcell, R., Archbold, N., George, A., Tye, S., Alexander, R., & Foster, C. (2004). Unique marine Permian‐Triassic boundary section from Western Australia. Australian Journal of Earth Sciences, 51(3), 423-430.

Tiwari, R.S, Misra, B.K, & Meena, K.L. (1990). Pyritic degradation of palynomorphs in Early triassic assemblage from Raniganj coalfield, West Bengal. Indian Journal of Geology, 62(2), 89-95.

Tiwari, R.S, Vijaya & Misra, B.K. (1994). Taphonomy of spores and pollen in Gondwana sequence of India. Palaeobotanist, 42(2) 108-119.

Totterdell, J., Moloney, J., Korsch, R., & Krassay, A. (2009). Sequence stratigraphy of the Bowen–Gunnedah and Surat Basins in New South Wales. Australian Journal of Earth Sciences, 56(3), 433-459.

Townrow, J. A. (1967). Voltziopsis, a southern conifer of lower Triassic age. Paper presented at the Papers and Proceedings of the Royal Society of Tasmania, 101, 173-190.

Traverse, A. (2007). Paleopalynology (Vol. 28): Springer, New York. 814 pp.

Twitchett, R. J., Looy, C. V., Morante, R., Visscher, H., & Wignall, P. B. (2001). Rapid and synchronous collapse of marine and terrestrial ecosystems during the end-Permian biotic crisis. Geology, 29(4), 351-354.

164

Tyson, R. V. (1995). Sedimentary Organic Matter: Organic Facies and Palynofacies. London: Chapman & Hall. 615 pp.

Uysal, I. T., Glikson, M., Golding, S. D., & Audsley, F. (2000). The thermal history of the Bowen Basin, Queensland, Australia: vitrinite reflectance and clay mineralogy of Late Permian coal measures. Tectonophysics, 323(1-2), 105-129.

Vajda, V., & McLoughlin, S. (2007). Extinction and recovery patterns of the vegetation across the Cretaceous–Palaeogene boundary—a tool for unravelling the causes of the end-Permian mass-extinction. Review of palaeobotany and palynology, 144(1-2), 99-112.

Vajda, V., McLoughlin, S., Mays, C., Frank, T. D., Fielding, C. R., Tevyaw, A., Lehsten, V., Bocking, M., & Nicoll, R. S. (2020). End-Permian (252 Mya) , wildfires and flooding—An ancient biotic crisis with lessons for the present. Earth and Planetary Science Letters, 529, 115875.

Van de Wetering, N., Esterle, J., & Baublys, K. (2013a). Decoupling δ13C response to palaeoflora cycles and climatic variation in coal: a case study from the Late Permian Bowen Basin, Queensland, Australia. Palaeogeography, Palaeoclimatology, Palaeoecology, 386, 165-179.

Van de Wetering, N., Mendonça Filho, J. G., & Esterle, J. (2013b). Palynofacies changes and their reflection on preservation of peat accumulation stages in the Late Permian coal measures of the Bowen Basin, Australia: A new system for coal palynofacies characterisation. International Journal of Coal Geology, 120, 57-70.

Van der Kaars, W. (1991). Palynology of eastern Indonesian marine piston-cores: a Late Quaternary vegetational and climatic record for Australasia. Palaeogeography, Palaeoclimatology, Palaeoecology, 85(3-4), 239-302.

Van Heeswijck, A. (2010). Late Paleozoic to early Mesozoic deformation in the northeastern Galilee Basin, Australia. Australian Journal of Earth Sciences, 57(4), 431-451.

Veevers, J. (2004). Gondwanaland from 650–500 Ma assembly through 320 Ma merger in Pangea to 185–100 Ma breakup: supercontinental tectonics via stratigraphy and radiometric dating. Earth-Science Reviews, 68(1-2), 1-132.

Vellekoop, J., Smit, J., van de Schootbrugge, B., Weijers, J. W., Galeotti, S., Damste, J. S. S., & Brinkhuis, H. (2015). Palynological evidence for prolonged cooling along the Tunisian continental shelf following the K–Pg boundary impact. Palaeogeography, Palaeoclimatology, Palaeoecology, 426, 216-228.

165

Venkatachala, B. S., & Kar, R. K. (1990). Reworked Permian Dulhuntyispora in tertiary sediments of northeastern India. Alcheringa, 14(3), 177-180.

Vine, R., Casey, D., & Johnson, N. (1964). Progress Report, 1963, on the Geology of part of the north-eastern Eromanga Basin, Queensland. 394 pp.

Visscher, H., Brinkhuis, H., Dilcher, D. L., Elsik, W. C., Eshet, Y., Looy, C. V., Rampino, M. R., & Traverse, A. (1996). The terminal Paleozoic fungal event: evidence of terrestrial ecosystem destabilization and collapse. Proceedings of the National Academy of Sciences, 93(5), 2155-2158.

Visscher, H., Looy, C. V., Collinson, M. E., Brinkhuis, H., Van Konijnenburg-Van Cittert, J. H., Kürschner, W. M., & Sephton, M. A. (2004). Environmental mutagenesis during the end- Permian ecological crisis. Proceedings of the National Academy of Sciences, 101(35), 12952-12956.

Wagner, N., Eble, C., Hower, J., & Falcon, R. (2019). Petrology and palynology of select coal samples from the Permian Waterberg Coalfield, South Africa. International Journal of Coal Geology, 204, 85-101.

Wang, C., & Visscher, H. (2007). Abundance anomalies of aromatic biomarkers in the Permian– Triassic boundary section at Meishan, China—evidence of end-Permian terrestrial ecosystem collapse. Palaeogeography, Palaeoclimatology, Palaeoecology, 252(1-2), 291- 303.

Waterhouse, J., & Shi, G. (2013). Climatic implications from the sequential changes in diversity and biogeographic affinities for brachiopods and bivalves in the Permian of eastern Australia and New Zealand. Gondwana Research, 24(1), 139-147.

Weaver, L., McLoughlin, S., & Drinnan, A. (1997). Fossil woods from the Upper Permian Bainmedart Coal Measures, northern Prince Charles Mountains, East Antarctica. AGSO Journal of Australian Geology and Geophysics, 16, 655-676.

Webb, J. A. (2001). A new marattialean fern from the Middle Triassic of eastern Australia. In Proceedings-Linnean Society of New South Wales, 123, 215-224.

Wheeler, A., Van de Wetering, N., Esterle, J. S., & Götz, A. E. (2018). Palaeoenvironmental changes recorded in the palynology and palynofacies of a Late Permian marker mudstone (Galilee Basin, Australia). Palaeoworld.

166

Wignall, P. B. (2001). Large igneous provinces and mass extinctions. Earth-Science Reviews, 53(1-2), 1-33.

Wignall, P., Thomas, B., Willink, R., Watling, J., Becker, L., Poreda, R., & Pope, K. (2004). Is Bedout an impact crater? Take 1. Science, 306(5696), 609-611.

Willis, K., & McElwain, J. (2002). The Evolution of Plants. Oxford: Oxford University Press. 378 pp.

Wilson, S. R. (2017). Controls on Sediment Distribution in the Late Permian Rangal Coal Measures of the Nebo Synclinorium. (Unpublished MPhil Thesis). The University of Queensland,

Wood, G. (1996). Palynological techniques-processing and microscopy. In. Jasonius, J. and McGregor, DC eds., Palynology: Principles and Application. American association of stratigraphic palynologists foundation, 1, 29-50.

Xie, S., Pancost, R. D., Huang, J., Wignall, P. B., Yu, J., Tang, X., Chen, L., Huang, X., & Lai, X. (2007). Changes in the global carbon cycle occurred as two episodes during the Permian– Triassic crisis. Geology, 35(12), 1083-1086.

Yang, Z. Y., Wu, S. B., Yin, H. F., Xu, G. R., Zhang, K. X., & Bi, X. M. (1993). Permo-Triassic Events of South China. Beijing: Geological Publishing House. 153 pp.

Yin, H., Feng, Q., Lai, X., Baud, A., & Tong, J. (2007). The protracted Permo-Triassic crisis and multi-episode extinction around the Permian–Triassic boundary. Global and Planetary Change, 55(1-3), 1-20.

Yin, H., Jiang, H., Xia, W., Feng, Q., Zhang, N., & Shen, J. (2014). The end-Permian regression in South China and its implication on mass extinction. Earth-Science Reviews, 137, 19-33.

Yin, H., Sweet, W., Glenister, B., Kotlyar, G., Kozur, H., Newell, N., Sheng, J., Yang, Z., & Zakharov, Y. D. (1996). Recommendation of the Meishan section as Global Stratotype Section and Point for basal boundary of Triassic System. Newsletters on Stratigraphy, 81- 108.

Yin, H., Zhang, K., Tong, J., Yang, Z., & Wu, S. (2001). The global stratotype section and point (GSSP) of the Permian-Triassic boundary. Episodes, 24(2), 102-114.

Yu, J., Peng, Y., Zhang, S., Yang, F., Zhao, Q., & Huang, Q. (2007). Terrestrial events across the Permian–Triassic boundary along the Yunnan–Guizhou border, SW China. Global and Planetary Change, 55(1-3), 193-208.

167

Zavada, M. S. (1991). The ultrastructure of pollen found in the dispersed sporangia of Arberiella (Glossopteridaceae). Botanical Gazette, 152(2), 248-255.

Zavattieri, A. M., Gutiérrez, P. R., & Ezpeleta, M. (2017). Syndesmorion stellatum (Fijałkowska) Foster et Afonin chlorophycean algae and associated microphytoplankton from lacustrine successions of the La Veteada Formation (late Permian), Paganzo Basin, Argentina. Paleoenvironmental interpretations and stratigraphic implications. Review of Palaeobotany and Palynology, 242, 1-20.

Zhang, F., Romaniello, S. J., Algeo, T. J., Lau, K. V., Clapham, M. E., Richoz, S., Herrmann, A. D., Smith, H., Horacek, M., & Anbar, A. D. (2018). Multiple episodes of extensive marine anoxia linked to global warming and continental weathering following the latest Permian mass extinction. Science Advances, 4(4), e1602921.

168

Spatiotemporal variation in the Permian-Triassic pre- and post- extinction palynology of the Bowen and Galilee basins (Australia)

Alexander Thomas Wheeler BSc. (Hons.) (Geology) – Rhodes University MSc. (Geology) – University of Pretoria

Appendices

Appendices of a thesis submitted for the degree of Doctor of Philosophy at The University of Queensland in 2020 School of Earth and Environmental Science

0

Appendix A - Oral and Written communication in conferences

A1. Extended Abstract for oral presentation at the Sydney Basin Symposium (2017)

Palynological records of marine incursions in the Sydney-Gunnedah-Bowen Basin – preliminary investigation

A. Wheeler1, J. Esterle1, & N. van de Wetering1

1School of Earth Sciences, The University of Queensland, QLD 4072, Australia, Corresponding Author: e-mail: [email protected]

Introduction

Recent palynological work in the Permian deposits of the eastern Australian basins has focused on calibrating the regional correlation scheme developed by Price (1997) using zircon age dates from the abundant tuffs that can be found in the Bowen, Gunnedah, Sydney and Galilee basins (Bodorkos et al., 2016; Laurie et al., 2016; Nicoll et al., 2015; Smith & Mantle, 2013). With a better understanding of the timing of deposition, further work is also being done looking at the palaeoenvironment of the coal deposits using palynofacies (van de Wetering, 2013). The appearance and abundance of spinose acritarchs within the marine sequences of these basins offers palynologists another tool for understanding the timing of deposition and palaeoenvironment. This work serves as a short review of marine incursions recorded in the palynological record of the Sydney-Gunnedah-Bowen Basin in the late Permian and at the P-T boundary, and future work to be done in the Galilee.

Spinose acritarchs tend to be indicative of marine or paralic environments with higher abundances and diversities indicative of open marine settings and lower abundances and diversities present in lagoonal or deltaic environments (Smith & Saunders, 1970, Segroves 1967; McLoughlin, 1993). The most widespread and diverse genera of spinose acritarchs in the Permian are the acanthomorph Micrhystridium and polygonomorph Veryhachium, and are the most useful acritarch species in terms of palaeoenvironmental reconstruction (Lei et al., 2013).

Bowen and Gunnedah Basins

Evans (1962) initially developed a biostratigraphic scheme for the southern Bowen basin using both spores and acritarchs as index fossils, but later reconfigured it to just use pollen and spores for

1 better correlation with fully terrestrial deposits (Evans, 1967a). This scheme was further adapted for widespread use in Permian-aged eastern Australian basins (Price, 1983, 1997). The complex series of transgressive and regressive cycles within the Denison Trough suggest a highly dynamic environment which is also reflected in the palynology. Spinose acritarchs are present in the Denison Trough from the Reid’s Dome Beds up to the top of the Peawaddy Formation with the highest abundance in the Ingelara and Peawaddy formations and few to none in the Freitag Formation (Rigby & Hekel, 1977; Fielding & McLoughlin, 1992). The extensive coal deposits of the Bowen Basin also tend to feature low number and diversity of acritarchs, indicating a non-marine depositional environment for the coals (Foster, 1979). McLoughlin (1988) identified two major spikes in spinose acritarch abundance in the marine Ingelara Formation from the well GSQ Springsure 19 (Figure 1). Zircon samples taken from the Ingelara Formation in the well GSQ Eddystone 4 give a date of 257.30±0.5 Ma (Nicoll et al., 2015).

The Catherine Sandstone above is also thought to have been deposited in a shallow marine to deltaic environment (Bann & Fielding, 1993), but features a much lower proportion of spinose acritarchs. The spikes in spinose acritarch abundance appear to coincide with transgressive facies in the Denison Trough. The most notable event reflecting this is the “P3C acme event” (Evans, 1962; Price, 1997) which is defined by a spike in the abundance of the spinose acritarch Micrhystridium evansii at the top of the Peawaddy Formation and Mantuan Productus Beds. This event has also been detected in the Gunnedah Basin in the well Santos Brawboy 1 in the lowermost Trinkey Formation, and zircon dates from this well place the event as occurring between 255.40±0.09 Ma and 255.03±0.05 Ma (Wood & Gallagher, 2012; Laurie et al., 2016). Dating this acme event may provide a strong stratigraphic tie line between the Permian basins as previous attempts to correlate the marine sequences have been difficult due to the complex tectonic history of eastern Australia.

2

Figure 1: Relative abundance (%) of acritarchs in the Late Permian deposits of the Bowen Basin from GSQ Springsure 19 (Dataset from McLoughlin, 1988).

Sydney Basin

The Sydney Basin has much fewer studies focusing on marine acritarchs, due to more limited marine sequences and more focus on studying the coal deposits. Evans (1967b) noted minor proportions of Micrhystridium and Veryhachium in the Newcastle Coal Measures and tentatively suggested a paralic environment for the coal swamps. McMinn (1985) looked at palynological samples from two marine incursions in the middle to late Permian deposits of the Sydney Basin. The first marine incursion occurs in the Kulnura Marine Tongue, Bulga Formation and Ennis Vale Formation and the second in the Dempsey, Denman and Baal Bone Formations (Bembrick 1983, Bamberry 1992). The only spinose acritarch identified from these intervals is Mehlisphaeridium sp. cf. M. fibratum which may be more indicative of brackish environments (McMinn 1985). These findings suggest a direct correlation of the marine intervals of the Bowen and Sydney Basins may be difficult.

Galilee Basin

The Galilee Basin remains relatively understudied compared to the Sydney-Gunnedah-Bowen Basin even in terms of palynology as a whole. While the biostratigraphy has been adapted to fit the

3 scheme of Price (1997) the most extensive palynological study of the basin was done by Norvick (1981). This study identifies and maps a limited marine influence within the south-eastern part of the basin with a proximity to the Springsure Shelf. The Micrhystridium evansii acme event was detected in four wells along the eastern margin of the basin (Jericho 1, Birkhead 1, Cunno 1, and Balfour 1).

Permo-Triassic Boundary

An acritarch spike is recognised in many localities above the Permo-Triassic boundary and is seen to represent a post-extinction pioneer community (Visscher & Brugman, 1981; Balme, 1970, 1979). This spike has been well detailed as a marine transgression event in the marine basins of Western Australia (Gorter et al., 2009; Kemp et al., 1977; Balme, 1963; Dolby & Balme, 1976). In eastern Australia, the Permo-Triassic boundary has been poorly covered due to widespread unconformities and the acritarch acme has only been recorded in the Sydney Basin (Figure 2) suggesting a short- lived marine transgression in the earliest part of the Triassic (Retallack, 1995).

Figure 2: Relative abundance of palynomorphs displaying a sharp increase in acritarchs across the Permo-Triassic boundary in the Sydney Basin (modified from Retallack, 1995).

Conclusions and Outlook

Work focusing on the marine palynology of the Sydney-Gunnedah-Bowen and Galilee basins has been sparse. This is unsurprising due to the nature of the marine sequences, which are limited

4 spatially and temporally, and the complex stratigraphy and tectonic history which makes lithological and palynological correlation difficult. However, the datasets reviewed in this work suggest that acritarch acme events coincide with transgressive events, and may therefore be useful in further refining the interbasinal correlation schemes for the Bowen, Gunnedah, Sydney and Galilee basins. More work needs to be done on a large scale to collect palynological datasets from all eastern Australian basins as well as to develop an interdisciplinary approach possibly incorporating marine invertebrate biostratigraphy together with palynology. Future work in the Galilee Basin will explore correlation of the Denison Trough and the Galilee Basin across the Springsure Shelf and marine acritarchs may play a crucial role in this endeavour, especially as more zircon dates become available.

References Balme, B. E. (1963). Plant microfossils from the Lower Triassic of Western Australia. Palaeontology 6(1) 12-40.

Balme, B.E., (1970). Palynology of Permian and Triassic strata in the Salt Range and Surghar Range, West Pakistan. Stratigraphic boundary problems: Permian and Triassic of West Pakistan. In Kummel, B., Teichert, C. (Eds.), University of Kansas, Department of Geology Special Publications 4, 305-453.

Balme, B. E. (1979). Palynology of Permian-Triassic boundary beds at Kap Stosch, East Greenland. Nyt nordisk forlag 200(6).

Bamberry, W. J. (1992). Stratigraphy and sedimentology of the late Permian Illawarra Coal Measures, southern Sydney basin. New South Wales. PhD. thesis, University of Wollongong (unpublished), Wollongong.

Bann, K. L., & Fielding, C. R. (2004). An integrated ichnological and sedimentological comparison of non-deltaic shoreface and subaqueous delta deposits in Permian reservoir units of Australia. Geological Society, London, Special Publications, 228(1), 273-310.

Bembrick, C. S. (1983). Stratigraphy and sedimentation of the late Permian Illawarra coal measures in the western coalfield, Sydney Basin. New South Wales. Journal and Proceedings of the Royal Society of New South Wales, 116, 105-117.

Bodorkos, S., Crowley, J., Holmes, E., Laurie, J., Mantle, D., McKellar, J., Wood, G. (2016). New dates for Permian palynostratigraphic biozones in the Sydney, Gunnedah, Bowen, Galilee and Canning basins, Australia. Permophiles 63, 19-21.

5

Dolby, J. H., & Balme, B. E. (1976). Triassic palynology of the Carnarvon Basin, Western Australia. Review of Palaeobotany and Palynology, 22(2), 105-168.

Evans, P.R. (1962). Palynological examination of AAO Westgrove 2 Well, Surat Basin, Queensland. Mines Administration Pty Ltd, Well Completion Report, Q/55-56, P/110.

Evans P. R. (1967a). Upper Carboniferous and Permian palynological stages and their distribution in eastern Australia. (1967/99) Bureau of Mineral Resources Geology and Geophysics. Retrieved from www.ga.gov.au.

Evans, P. R. (1967b). Review of the Permian palynology of the Sydney Basin. (1967/103). Bureau of Mineral Resources Geology and Geophysics. Retrieved from www.ga.gov.au.

Fielding, C. R., & McLoughlin, S. (1992). Sedimentology and palynostratigraphy of Permian rocks exposed at Fairbairn Dam, central Queensland. Australian Journal of Earth Sciences, 39(5), 631- 649.

Foster, C. B. (1979). Permian Plant Microfossils of the Blair Athol Coal Measures, Baralaba Coal Measures and Basal Rewan Formation of Queensland. (GSQ Publication 372). Brisbane: Geological Survey of Queensland Retrieved from www.qdexguest.dnrm.qld.gov.au.

Gorter, J. D., Nicoll, R. S., Metcalfe, I., Willink, R. J., & Ferdinando, D. (2009). The Permian- Triassic boundary in Western Australia: evidence from the Bonaparte and northern Perth basins: exploration implications. The APPEA Journal, 49(1), 311-336.

Kemp, E. M., Balme, B. E., Helby, R. J., Kyle, R. A., Playford, G., & Price, P. L. (1977). Carboniferous and Permian palynostratigraphy in Australia and Antarctica: a review. BMR Journal of Australian Geology and Geophysics, 2, 177-208.

Laurie, J. R., Bodorkos, S., Nicoll, R. S., Crowley, J. S., Mantle, J. D., Mory, A. J., . . . Champion, D. C. (2016). Calibrating the middle and late palynostratigraphy of Australia to the geologic time scale via U-Pb zircon CA-IDTIMS dating. Australian Journal of Earth Sciences, 63(6), 701-730.

Lei, Y., Servais, T., & Feng, Q. (2013). The diversity of the Permian phytoplankton. Review of Palaeobotany and Palynology, 198, 145-161.

McLoughlin, S. (1988). Geology of the Inglis Dome, Denison Trough, central Queensland. Department of Geology, University of Queensland, 12, 229-263.

McLoughlin, S. (1993). Plant fossil distributions in some Australian Permian non-marine sediments. Sedimentary Geology, 85(1-4), 601-619.

6

McMinn, A. (1985). Palynostratigraphy of the Middle Permian coal sequences of the Sydney Basin. Australian Journal of Earth Sciences, 32(3), 301-309.

Nicoll, R., McKellar, J., Ayaz, S. A., Laurie, J., Esterle, J., Crowley, J.,& Bodorkos, S. (2015). CA- IDTIMS dating of tuffs, calibration of palynostratigraphy and stratigraphy of the Bowen and Galilee basins. In Bowen Basin Symposium (pp. 211-218).

Norvick, M. (1981). Permian and Late Carboniferious Palynostratigraphy of the Galilee Basin, Queensland. (1974/141). Bureau of Mineral Resources Retrieved from www.ga.gov.au.

Price, P.L. (1997). Permian to Jurassic palynostratigraphic nomenclature of the Bowen and Surat Basins. In Green, P. M. (Ed.), The Surat and Bowen Basins, south-east Queensland (pp. 137-178). Queensland Minerals and Energy Review Series, Queensland Department of Mines and Energy.

Price, P. L. (1983). A Permian palynostratigraphy for Queensland Permian Geology of Queensland (pp. 155-211). Queensland Division, Brisbane: Geological Society of Australia.

Price, P.L. (1997). Permian to Jurassic palynostratigraphic nomenclature of the Bowen and Surat Basins. In Green, P. M. (Ed.), The Surat and Bowen Basins, south-east Queensland (pp. 137-178). Queensland Minerals and Energy Review Series, Queensland Department of Mines and Energy.

Price, P. L. (1983). A Permian palynostratigraphy for Queensland Permian Geology of Queensland (pp. 155-211). Queensland Division, Brisbane: Geological Society of Australia.

Retallack, G. J. (1995). Permian-Triassic life crisis on land. Science, 267(5194), 77.

Rigby, J. F., & Hekel, H. (1977). Palynology of the Permian sequence in the Springsure Anticline, central Queensland (Vol. 363). Queensland, Dept. of Mines.

Segroves, K. L. (1967). Cutinized microfossils of probable nonvascular origin from the Permian of Western Australia. Micropaleontology, 289-305.

Smith, N. D., & Saunders, R. S. (1970). Paleoenvironments and their control of acritarch distribution: of east-central Pennsylvania. Journal of Sedimentary Research, 40(1), 324- 333.

Smith, T. E., & Mantle, D. (2013). Late Permian palynozones and associated CA-IDTIMS dated tuffs from the Bowen Basin, Australia. Geoscience Australia Record, 46.

Van de Wetering, N., Mendonça Filho, J. G., & Esterle, J. (2013). Palynofacies changes and their reflection on preservation of peat accumulation stages in the Late Permian coal measures of the

7

Bowen Basin, Australia: A new system for coal palynofacies characterisation. International Journal of Coal Geology, 120, 57-70.

Visscher, H., & Brugman, W. A. (1981). Ranges of selected palynomorphs in the Alpine Triassic of . Review of Palaeobotany and Palynology, 34(1), 115-128.

Wood, G. R., & Gallagher, L. (2011). Palynostratigraphical analysis, Brawboy No. 1, PEL 456 Gunnedah Basin. Santos Stratigraphic Services, Subsurface Services Department, Palynology Report No, 1.

Acknowledgements

The authors would like to thank the UQ-Vale Fund for continued financial support. We would also like to thank Bob Nicoll for advice and discussion during preparation of this manuscript.

8

A2. Abstract for poster presented at 50th Annual AASP Meeting (2017)

Palynofacies of an end-Permian marker mudstone in the Galilee Basin, Australia

1 1 1 2 A. Wheeler , J. Esterle , N. Van de Wetering , A.E. Götz

1The University of Queensland, QLD, 4067, Australia ([email protected])

2University of Portsmouth, University House, Winston Churchill Ave, Portsmouth PO1 2UP, United Kingdom

The Permo-Triassic Boundary in terrestrial deposits remains difficult to study around the world. In eastern Australia, recent progress has been made in the use of high-resolution zircon dates and the recalibration of the regional Permian biostratigraphic scheme. This has also revealed the need for other techniques to be used in interbasinal correlation and palaeoenvironmental reconstruction. One such tool is palynofacies analysis, which can provide insight into palaeoenvironmental conditions present during end-Permian times.

In parts of the Bowen and Galilee basins, an organic-rich marker mudstone is present above the last coals deposited in the Permian. Above this mudstone lie the coal-free sandstones of the Triassic Rewan Formation. Thusly, this marker mudstone may potentially record the palaeoenvironmental conditions present during the Permo-Triassic transition in Australia. Five samples were obtained from the marker mudstone in the Tambo 1-1A well in the Galilee Basin. These samples were processed for biostratigraphic assessment and palynofacies analysis.

The palynofacies data show a switch from a translucent phytoclast-dominated assemblage at the base of the marker mudstone to an opaque phytoclast-dominated assemblage at the top. This trend is the opposite of palynofacies shifts observed in similar age strata in western Australia, and may be indicative of an extreme shift in the redox conditions of the lake in which the mudstone was deposited rather than reflecting a continent-wide or global event. The palynological assemblage within the marker mudstone is consistent with other end-Permian assemblages found in eastern Australia, though no index taxa have yet been identified, which prevents an accurate age assignment. Though very rare in abundance, some specimens of Micrhystridium acritarchs have been identified within the mudstone. No freshwater algae have been observed. This may be due to preservation biases or extreme environmental conditions but it remains difficult to draw significant conclusions about the palaeoenvironment. Another rare component of the assemblage is Reduviasporinites chalastus, which has commonly been identified in other end-Permian assemblages in Australia, though its affinity to either fungal or algal groups remains questionable.

9

While only from a single locality, the palynofacies data suggest dynamic local environmental conditions. These conditions may be attributed not just to the shifting palaeoclimate but also to regional tectonics as base level rises in response to foreland loading. As more datasets are added to this one, the details about conditions present during the end-Permian in eastern Australia will become clearer.

Keywords: Palynology, Palynofacies, Galilee Basin, Late Permian, Permo-Triassic, Australia

10

A3. Abstract for oral presentation at the 51st Annual AASP Meeting (2018)

Palynology and palynofacies of the Late Permian Galilee Basin: Implications for the end- Permian palaeoenvironment

A. Wheelera, N. Van de Weteringa, J. Esterlea and A.E. Götzb aThe University of Queensland, School of Earth and Environmental Sciences, St. Lucia QLD 4072, Australia bUniversity of Portsmouth, School of Earth and Environmental Sciences, Portsmouth PO1 3QL, United Kingdom

The Permian-Triassic Boundary in the terrestrial basins of Eastern Australia is characterised by palynofloral turnover and shifts in palaeoenvironmental conditions but is often difficult to examine in detail due to widespread erosion and conditions which were unsuitable for preserving sedimentary organic matter. However, an organic-rich carbonaceous mudstone lying just above the uppermost Permian coal deposits has been mapped out in parts of the Bowen and Galilee basins, and may provide insight into end-Permian palaeoenvironmental conditions in the region. This ‘marker mudstone’ has long been used as a reference point by industry, but has only been studied at one locality in the Bowen Basin. The borehole Tambo 1-1A in the Galilee Basin was sampled for palynological, palynofacies and stable carbon isotope analysis, with twenty-two samples taken from the marker mudstone, the underlying Bandanna Formation and the overlying Rewan Formation.

The marker mudstone represents the uppermost expression of the APP5 biozone with the transition to the APP6 biozone occurring over several centimetres above it. The palynofacies shift in the marker mudstone from an assemblage featuring a high proportion of translucent phytoclasts to an assemblage with increased proportions of opaque phytoclasts suggests either changing redox conditions within the water column or increased input from fluvio-deltaic systems. The marker mudstone and the overlying strata contain low percentages of acanthomorph acritarchs, algae and fungal hyphae. The algal forms resemble Botryococcus but differ in morphology from the algal bodies in the underlying deltaic facies. Based on these data, the marker mudstone was likely deposited in a large, shallow lake which would have formed due to subsidence and a rising base level related to foreland loading at the eastern margin of the Bowen Basin. The acritarchs may be indicative of increased salinity caused by a short-term marine transgression.

The presence of degraded and opaque phytoclasts in the same interval as the transition to the APP6 biozone may be indicative of an extinction event in this level, but may also be related to localised changes in the palaeoenvironment. The latter interpretation is supported by the low magnitude of

11 the negative excursion of the δ13C isotope values in the marker mudstone and the low abundances of fungal remains. Based on the biostratigraphic data, the marker mudstone also does not directly correlate with previously studied marker mudstone horizons in the northern part of the Bowen Basin. This may be related to a diachronous onset of Rewan Formation deposition, diachroneity at the base of the APP6 biozone or the marker mudstone in the Galilee Basin may not be a true correlative unit of the one mapped in the Bowen Basin.

Key words: palynology; palynofacies; carbon isotopes; palaeoenvironment; Permian-Triassic Boundary, Galilee Basin

12

A4. Abstract for poster presented at the Australian Geoscience Council Convention (AGCC) (2017)

Palynology of a Late Permian marker mudstone in the Bowen and Galilee basins: Implications for inter-basinal correlation

A. Wheelera, N. Van de Weteringa, J. Esterlea and A.E. Götzb

aThe University of Queensland, School of Earth and Environmental Sciences, St. Lucia QLD 4072, Australia bUniversity of Portsmouth, School of Earth and Environmental Sciences, Portsmouth PO1 3QL, United Kingdom

The coal deposits of the intracratonic Galilee Basin are understudied compared to the Bowen Basin to the East. Even with the well-constrained biostratigraphy developed in eastern Australia, intra- and interbasinal correlation remain challenging due to a lack of detailed palynological studies in the late Permian sections of the Galilee basin. A carbonaceous mudstone, dubbed the marker mudstone, deposited above the uppermost Permian coal seams represents a potentially useful tool for regional correlation of the end-Permian deposits in both basins. The aims of this study were to place the marker mudstone in a biostratigraphic context using samples from the borehole Tambo 1-1A and to reconstruct the end-Permian palaeoenvironment of the Springsure Shelf in the Galilee Basin.

When compared to biostratigraphic data from a section containing the marker mudstone in the Bowen Basin, the samples from the Galilee Basin show a degree of diachroneity. In the Galilee Basin, the marker mudstone represents the uppermost expression of the APP5 biozone (based on the scheme of Price 1997) while in the northern Bowen Basin it represents the base of the APP6 biozone. Palaeoenvironmental data suggests the marker mudstone represents a lacustrine deposit, which formed as base level rose and the alluvial-deltaic system that deposited the Bandanna Formation sediments subsided. Whether the diachroneity reflects changing palaoenvironment or a separate unit is still under study.

A5. Abstract for poster presented at the 19th International Congress on the Carboniferous and Permian (XIX ICCP 2019)

13

Proximal-distal palaeoenvironmental patterns recorded by aquatic palynomophs in the Galilee Basin, Australia

Alexander WHEELER1, Joan S. ESTERLE1, Annette E. GÖTZ2,3

1University of Queensland, School of Earth and Environmental Sciences, St. Lucia, QLD 4072, Australia; [email protected] 2University of Portsmouth, Burnaby Road, Portsmouth PO1 3QL, United Kingdom 3Kazan Federal University, ul. Kremlyovskaya 18, Kazan, 420008 Russia

The Late Permian sedimentary sequence of the Galilee Basin represents a variety of depositional environments, which transition from fluvial through to shallow marine. Recent work within the basin has focused on using sedimentology and geochemistry to refine the lithostratigraphy and reconstruct the palaeoenvironment (Phillips et al. 2017). While palynology has been used as a chronostratigraphic tool within the basin, less emphasis has been placed on the utility of terrestrial and aquatic palynomorphs in palaeoenvironmental reconstruction. This work represents a preliminary study that aims to identify aquatic palynomorphs in the Galilee Basin and, along with detailed sedimentological data, develop a more detailed palaeoenvironmental interpretation for the basin. To this end, three borehole cores, representing a proximal to distal transition, were sampled.

The results suggest a distinct change in the assemblage of aquatic palynomorphs from north to south. Tetraporina and Quadrisporites are limited to the northernmost well suggesting a freshwater environment. Botryococcus is present in all holes in varying abundances but is most common in lagoonal-deltaic facies, suggesting a tolerance to a variety of salinities. In terms of prasinophytes, Cymatiosphaera gondwanensis also appears sparsely in deltaic facies, while Mehlisphaeridium fragile is present in freshwater facies. This suggests the prasinophytes occupy a wider variety of salinity regimes and are not limited to marine environments as observed in modern studies (Mudie et al. 2011). Acanthomorph acritarchs are limited to the distal part of the basin reach their highest abundance during the Micrhystridium evansii acme event which occurred as a result of a regional- scale marine transgression. Just above this, the marine facies gives way to prodelta and delta front facies, wherein the assemblage becomes more diverse featuring several species of Michrhystridium along with Brazilea, Peltacystia and Pilasporites. In the overlying strata of the Bandanna Formation, Micrhystridium is absent until it reappears in the marker mudstone, which likely marks the end of the

14

Permian and suggests increased salinity associated with end-Permian environmental conditions (Wheeler et al. 2018).

The change from freshwater algae in the north to more saline tolerant taxa and acanthomorph acritarchs in the south is in agreement with previous sedimentological studies within the basin (Phillips et al. 2017). Some taxa (i.e. Tetraporina, Cymatiosphaera) appear to be more environmentally sensitive while others (i.e. Brazilea, Pilasporites) occur in quite varied environments and may be more dependant on nutrient input than on salinity or turbidity. Identifying the underlying environemental factors which control the distribution of these aquatic palynomorphs is key to further developing our understanding of ancient environments.

References

MUDIE, P. J., LEROY, S. A. G., MARRET, F., GERASIMENKO, N. P., KHOLEIF, S. E. A., SAPELKO, T., & FILIPOVA-MARINOVA, M. (2011): Nonpollen palynomorphs: indicators of salinity and environmental change in the Caspian–Black Sea–Mediterranean corridor. In Geology and geoarchaeology of the Black Sea region: beyond the flood hypothesis (Vol. 473, pp. 89-115). Boulder, CO: Geological Society of America.

PHILLIPS, L. J., EDWARDS, S. A., BIANCHI, V., & ESTERLE, J. S. (2017): Paleo-environmental reconstruction of Lopingian (upper Permian) sediments in the Galilee Basin, Queensland, Australia. Australian Journal of Earth Sciences, 64(5): 587-609.

WHEELER, A., VAN DE WETERING, N., ESTERLE, J. S., & GÖTZ, A. E. (In press): Palaeoenvironmental changes recorded in the palynology and palynofacies of a Late Permian marker mudstone (Galilee Basin, Australia). Palaeoworld.

15

Appendix B – Sample lists and raw data

B1. Localities sampled for this study

Wellname Basin Latitude(Decimal) Longitude(Decimal) Montani 1 Galilee -21,9589 146,0542 Glue Pot Creek Galilee -22,8013 145,9494 1 Tambo 1-1A Galilee -24,5318 146,6011 Springsure 19 Bowen -24,6862 148,1417 Taringa 7 Bowen -25,9549 148,5297 A41849 Bowen -24,6 150,0369 CGIE0144 Bowen -21,9726 148,1447 CGIN0067 Bowen -21,948 148,1333

B2. List of samples collected for this samples

Top Bottom Sample Formation Analysis Preparation depth (m) depth (m) Name

Montani 1

751,54 751,56 MON-1 Rewan Paly.(barren) Acid-free (UQ)

757,84 757,87 MON-2 Rewan Paly.(barren) Acid-free (UQ)

770,91 770,93 MON-3 Bandanna Paly.(barren) Acid-free (UQ)

774,37 774,4 MON-4 Bandanna Paly.(barren) Acid-free (UQ)

776,54 776,57 MON-5 Bandanna Paly. Acid-free (UQ)

781,18 781,21 MON-6 Bandanna Paly. Acid-free (UQ)

786,55 786,58 MON-7 Bandanna Paly. Acid-free (UQ)

792,83 792,85 MON-8 "Burngrove" Paly. Acid-free (UQ) equivalent

795,54 795,57 MON-9 "Burngrove" Paly.(barren) Acid-free (UQ) equivalent

807,14 807,17 MON-10 Black Alley Paly. Acid-free (UQ) Shale

812,07 812,1 MON-11 Black Alley Paly. Acid-free (UQ) Shale

814,75 814,78 MON-12 "Fair Hill" Paly. Acid-free (UQ) equivalent

16

817,66 817,69 MON-13 "Fair Hill" Paly. Acid-free (UQ) equivalent

824,53 824,56 MON-14 "Fair Hill" Paly. Acid-free (UQ) equivalent

827,7 827,73 MON-15 "Fair Hill" Paly.(barren) Acid-free (UQ) equivalent

Glue Pot Creek 1

582,73 582,75 GPC-1 Rewan Paly. Acid-free (UQ)

592,53 592,55 GPC-2 Rewan Paly.(barren) Acid-free (UQ)

602,63 602,65 GPC-3 Rewan Paly.(barren) Acid-free (UQ)

606,72 606,73 GPC-4 Rewan Paly.(barren) Acid-free (UQ)

615,72 615,75 GPC-5 Rewan Paly.(barren) Acid-free (UQ)

620,29 620,32 GPC-6 Bandanna Paly. Acid-free (UQ)

621,28 621,3 GPC-7 Bandanna Paly. Acid-free (UQ)

622,92 622,94 GPC-8 Bandanna Paly.(barren) Acid-free (UQ)

651,92 651,94 GPC-9 "Burngrove" Paly. Acid-free (UQ) equivalent

658,79 658,81 GPC-10 "Burngrove" Paly.(barren) Acid-free (UQ) equivalent

662,27 662,3 GPC-11 "Burngrove" Paly. Acid-free (UQ) equivalent

664,18 664,2 GPC-12 "Burngrove" Paly. Acid-free (UQ) equivalent

698,33 698,36 GPC-13 "Burngrove" Paly. Acid-free (UQ) equivalent

711,98 712 GPC-14 Black Alley Paly.(barren) Acid-free (UQ) Shale

712,86 712,88 GPC-15 Black Alley Paly.(barren) Acid-free (UQ) Shale

714,96 714,98 GPC-16 Black Alley Paly. Acid-free (UQ) Shale

716,11 716,13 GPC-17 Black Alley Paly. Acid-free (UQ) Shale

718,96 718,98 GPC-18 Black Alley Paly. Acid-free (UQ) Shale

719,77 719,79 GPC-19 Black Alley Paly.(barren) Acid-free (UQ) Shale

17

Tambo 1-1A

696,6 696,62 TAMP-1 Rewan Paly. Standard (MGPalaeo)

697,94 697,97 TAMP-2 Rewan Paly.(barren) Standard (MGPalaeo)

698,33 698,36 TAMP-3 Rewan Paly. Standard (MGPalaeo)

698,86 698,88 TAMP-4 Rewan Paly. Standard (MGPalaeo)

699,13 699,15 TAMP-5 Rewan Paly.(barren) Standard (MGPalaeo)

699,33 699,35 TAMP-6 Rewan Paly.(barren) Standard (MGPalaeo)

700,14 700,16 TAMP-7 Bandanna Paly.(barren) Standard (MGPalaeo)

700,24 700,26 TAMP-8 Bandanna Paly.(barren) Standard (MGPalaeo)

700,39 700,4 TAMP-9 Bandanna Paly. Standard (MGPalaeo)

700,49 700,52 TAMP- Bandanna Paly. Standard (MGPalaeo) 10

700,64 700,67 TAMP- Bandanna Paly. Standard (MGPalaeo) 11

701,69 701,72 TAMP- Bandanna Paly.(barren) Standard (MGPalaeo) 12

723,73 723,76 TAMP- "Burngrove" Paly.(barren) Standard (MGPalaeo) 13 equivalent

729,01 729,03 TAMP- "Burngrove" Paly.(barren) Standard (MGPalaeo) 14 equivalent

729,21 729,23 TAMP- "Burngrove" Paly.(barren) Standard (MGPalaeo) 15 equivalent

733,91 733,93 TAMP- "Burngrove" Paly. Standard (MGPalaeo) 16 equivalent

735,04 735,05 TAMP- "Burngrove" Paly. Standard (MGPalaeo) 17 equivalent

735,4 735,44 TAMP- "Burngrove" Paly. Standard (MGPalaeo) 18 equivalent

738,33 738,37 TAMP- "Burngrove" Paly. Standard (MGPalaeo) 19 equivalent

738,81 738,84 TAMP- "Burngrove" Paly. Standard (MGPalaeo) 20 equivalent

740,54 740,57 TAMP- "Burngrove" Paly. Standard (MGPalaeo) 21 equivalent

741,66 741,67 TAMP- "Burngrove" Paly. Standard (MGPalaeo) 22 equivalent

770,46 770,48 TAMP23 "Burngrove" Paly.(barren) Standard (MGPalaeo) + equivalent Acid-free (UQ)

18

780,99 781,02 TAMP24 Black Alley Paly. Standard (MGPalaeo) + Shale Acid-free (UQ)

791,85 791,88 TAMP25 Black Alley Paly. Standard (MGPalaeo) + Shale Acid-free (UQ)

803,07 803,1 TAMP26 Black Alley Paly.(barren) Standard (MGPalaeo) + Shale Acid-free (UQ)

810,18 810,22 TAMP27 Black Alley Paly.(barren) Standard (MGPalaeo) + Shale Acid-free (UQ)

821,09 821,11 TAMP28 Black Alley Paly. Standard (MGPalaeo) + Shale Acid-free (UQ)

826,98 827,01 TAMP29 Black Alley Paly. Standard (MGPalaeo) + Shale Acid-free (UQ)

828,2 828,23 TAMP30 Black Alley Paly. Standard (MGPalaeo) + Shale Acid-free (UQ)

829,09 829,11 TAMP31 Black Alley Paly. Standard (MGPalaeo) + Shale Acid-free (UQ)

829,25 829,28 TAMP32 Mantuan Paly. Standard (MGPalaeo) + Productus Beds Acid-free (UQ)

699,13 699,15 TAMI-1 Rewan δ13C (org) Isotope UQ Isotope Lab Analysis

699,33 699,35 TAMI-2 Rewan δ13C (org) Isotope UQ Isotope Lab Analysis

700,14 700,16 TAMI-3 Bandanna δ13C (org) Isotope UQ Isotope Lab Analysis

700,24 700,26 TAMI-4 Bandanna δ13C (org) Isotope UQ Isotope Lab Analysis

700,39 700,4 TAMI-5 Bandanna δ13C (org) Isotope UQ Isotope Lab Analysis

700,49 700,52 TAMI-6 Bandanna δ13C (org) Isotope UQ Isotope Lab Analysis

700,64 700,67 TAMI-7 Bandanna δ13C (org) Isotope UQ Isotope Lab Analysis

723,73 701,72 TAMI-8 Bandanna δ13C (org) Isotope UQ Isotope Lab Analysis

729,01 723,76 TAMI-9 "Burngrove" δ13C (org) Isotope UQ Isotope Lab equivalent Analysis

729,21 729,03 TAMI-10 "Burngrove" δ13C (org) Isotope UQ Isotope Lab equivalent Analysis

733,91 729,23 TAMI-11 "Burngrove" δ13C (org) Isotope UQ Isotope Lab equivalent Analysis

19

735,04 733,93 TAMI-12 "Burngrove" δ13C (org) Isotope UQ Isotope Lab equivalent Analysis

735,4 735,05 TAMI-13 "Burngrove" δ13C (org) Isotope UQ Isotope Lab equivalent Analysis

738,33 735,44 TAMI-14 "Burngrove" δ13C (org) Isotope UQ Isotope Lab equivalent Analysis

738,81 738,37 TAMI-15 "Burngrove" δ13C (org) Isotope UQ Isotope Lab equivalent Analysis

740,54 738,84 TAMI-16 "Burngrove" δ13C (org) Isotope UQ Isotope Lab equivalent Analysis

741,66 740,57 TAMI-17 "Burngrove" δ13C (org) Isotope UQ Isotope Lab equivalent Analysis

Springsure 19

320,77 320,8 SPR1 Rewan Paly.(barren) Acid-free (UQ)

346,15 346,18 SPR2 Rewan Paly.(barren) Acid-free (UQ)

359,26 359,28 SPR3 Rewan Paly. Acid-free (UQ)

359,37 359,41 SPR4 Rewan Paly. Acid-free (UQ)

360,03 360,05 SPR5 Rewan Paly. Acid-free (UQ)

360,76 360,78 SPR6 Bandanna Paly.(barren) Acid-free (UQ)

363,67 363,7 SPR7 Bandanna Paly. Acid-free (UQ)

378,77 378,8 SPR8 Bandanna Paly. Acid-free (UQ)

384,91 384,95 SPR9 Bandanna Paly. Acid-free (UQ)

385,85 385,88 SPR10 Bandanna Paly.(barren) Acid-free (UQ)

388,45 388,48 SPR11 Bandanna Paly. Acid-free (UQ)

391,28 391,31 SPR12 Bandanna Paly. Acid-free (UQ)

399,11 399,13 SPR13 Bandanna Paly. Acid-free (UQ)

421,32 421,35 SPR14 Bandanna Paly. Acid-free (UQ)

446,34 446,36 SPR15 Black Alley Paly. Acid-free (UQ) Shale

467,98 468,01 SPR16 Black Alley Paly. Acid-free (UQ) Shale

484,73 484,75 SPR17 Black Alley Paly. Acid-free (UQ) Shale

492,12 492,18 SPR18 Black Alley Paly. Acid-free (UQ) Shale

505,32 505,35 SPR19 Black Alley Paly. Acid-free (UQ) Shale

20

510,94 510,96 SPR20 Black Alley Paly. Acid-free (UQ) Shale

515,92 515,95 SPR21 Mantuan Paly. Acid-free (UQ) Productus Beds

523,6 523,63 SPR22 Peawaddy Paly. Acid-free (UQ)

Taringa 7

949,01 949,06 TAR1 Rewan Paly.(barren) Standard (MGPalaeo)

952,17 952,22 TAR2 Rewan Paly.(barren) Standard (MGPalaeo)

954,2 954,23 TAR3 Rewan Paly. Standard (MGPalaeo)

958,29 958,34 TAR4 Bandanna Paly. Standard (MGPalaeo)

960,7 960,74 TAR5 Bandanna Paly. Standard (MGPalaeo)

963,34 963,39 TAR6 Bandanna Paly.(barren) Standard (MGPalaeo)

967,06 967,13 TAR7 Bandanna Paly.(barren) Standard (MGPalaeo)

970,6 970,65 TAR8 Black Alley Paly. Standard (MGPalaeo) Shale

992,84 992,86 TAR9 Black Alley Paly. Standard (MGPalaeo) Shale

1011,9 1011,92 TAR10 Black Alley Paly. Standard (MGPalaeo) Shale

1039,15 1039,19 TAR11 Black Alley Paly. Standard (MGPalaeo) Shale

1063,1 1063,15 TAR12 Black Alley Paly. Standard (MGPalaeo) Shale

1069,27 1069,3 TAR13 Mantuan Paly. Standard (MGPalaeo) Productus Beds

1072 1072,05 TAR14 Mantuan Paly. Standard (MGPalaeo) Productus Beds

A41859 - Dawson Coal Mine

28.78 28.98 UD1 Rewan Paly.(barren) Standard (MGPalaeo)

36.0 36.2 UD2 Rewan Paly.(barren) Standard (MGPalaeo)

50.79 50.83 UQ1 Rewan Paly. Standard (MGPalaeo)

56.91 57.08 UD3 Rewan Paly. Standard (MGPalaeo)

58.07 58.2 UD4 Rewan Paly. Standard (MGPalaeo)

58.78 58.95 UD5 Rewan Paly. Standard (MGPalaeo)

21

81.39 81.54 UD6 Baralaba Coal Paly. Standard (MGPalaeo) Measures

109.19 109.37 UD7 Baralaba Coal Paly.(barren) Standard (MGPalaeo) Measures

128.3 128.34 UQ2 Baralaba Coal Paly. Standard (MGPalaeo) Measures

132.06 132.1 UQ3 Baralaba Coal Paly. Standard (MGPalaeo) Measures

138.53 138.57 UQ4 Baralaba Coal Paly. Standard (MGPalaeo) Measures

199.07 199.11 UQ5 Baralaba Coal Paly. Standard (MGPalaeo) Measures

214.61 214.65 UQ6 Baralaba Coal Paly. Standard (MGPalaeo) Measures

CGIE0144 - Isaac Plains Coal Mine

34,87 35,1 IPA1 Rewan Paly.(barren) Standard (MGPalaeo)

40,9 41,1 IPA2 Rewan Paly. Standard (MGPalaeo)

44,09 44,35 IPA3 Rangal Coal Paly.(barren) Standard (MGPalaeo) Measures

45,32 45,57 IPA4 Rangal Coal Paly. Standard (MGPalaeo) Measures

CGIN0067 - Isaac Plains Coal Mine

107,2 107,28 IPB1 Rewan Paly. Standard (MGPalaeo)

109,5 109,7 IPB2 Rewan Paly.(barren) Standard (MGPalaeo)

109,8 109,9 IPB3 Rewan Paly.(barren) Standard (MGPalaeo)

110,45 110,61 IPB4 Rewan Paly.(barren) Standard (MGPalaeo)

113,15 113,3 IPB5 Rewan Paly. Standard (MGPalaeo)

B3. Stable carbon isotope data (δ13C org)

ID run time 13C vpdb

TAMI-01.raw 11/1/17 2:00 -25,4

TAMI-01A.raw 11/1/17 2:10 -25,4

TAMI-02.raw 9/1/17 23:26 -25,4

TAMI-02A.raw 11/1/17 3:02 -25,6

22

TAMI-02.raw 11/1/17 17:24 -25,6

TAMI-02A.raw 11/1/17 17:34 -25,3

TAMI-03.raw 9/1/17 23:47 -26,0

TAMI-03A.raw 9/1/17 23:57 -26,0

TAMI-04.raw 10/1/17 0:27 -26,1

TAMI-04A.raw 10/1/17 0:37 -26,2

TAMI-05.raw 10/1/17 0:48 -26,0

TAMI-05A.raw 10/1/17 0:58 -26,4

TAMI-06.raw 10/1/17 1:08 -26,1

TAMI-06A.raw 10/1/17 1:18 -25,7

TAMI-07.raw 10/1/17 1:28 -25,5

TAMI-07A.raw 10/1/17 1:38 -25,3

TAMI-08.raw 11/1/17 3:12 -23,1

TAMI-08A.raw 11/1/17 3:22 -23,4

TAMI-09.raw 10/1/17 2:40 -23,1

TAMI-09A.raw 10/1/17 2:50 -23,2

TAMI-10.raw 11/1/17 3:33 -24,1

TAMI-10A.raw 11/1/17 3:43 -24,0

TAMI-11.raw 11/1/17 3:53 -23,6

TAMI-11A.raw 11/1/17 4:03 -24,0

TAMI-11.raw 11/1/17 17:44 -23,8

TAMI-11A.raw 11/1/17 17:55 -23,9

TAMI-12.raw 11/1/17 18:05 -23,5

TAMI-12A.raw 11/1/17 18:15 -23,4

TAMI-13.raw 11/1/17 4:55 -24,0

TAMI-13A.raw 11/1/17 5:05 -23,3

TAMI-13.raw 11/1/17 18:25 -23,8

TAMI-13A.raw 11/1/17 18:35 -23,3

TAMI-14.raw 11/1/17 5:15 -23,4

TAMI-14A.raw 11/1/17 5:26 -23,5

TAMI-15.raw 11/1/17 18:45 -23,3

TAMI-15A.raw 11/1/17 18:56 -23,3

TAMI-16.raw 11/1/17 5:57 -270,6 (peak split)

23

TAMI-16A.raw 11/1/17 6:07 -23,7

TAMI-17.raw 11/1/17 6:17 -23,8

TAMI-17A.raw 11/1/17 6:27 -23,6

24

B4. Palynological Data (Raw Counts)

B4.1 Montani 1

depth 751,56 757,87 770,93 774,40 776,57 781,21 786,58 792,85 795,57 807,17 812,10 814,78 817,69 824,56 827,73

sample palynom MON-1 MON-2 MON-3 MON-4 MON-5 MON-6 MON-7 MON-8 MON-9 MON- MON- MON- MON- MON- MON- orph 10 11 12 13 14 15 type*

Brazilea plurigenus AL 2

Brazilea scissa AL 1 2 9

Circulisporites spp. AL 1

Maculatasporites sp. AL 1 3 1

Peltacystia venosa AL 1 3

Pilasporites calculus AL 1 1 6

Quadrisporites horridus AL 3

Tetraporina sp. AL 2 6

Cymatiosphaera gondwanensis ALPR 1

Mehlisphaeridium fibratum ALPR 2

Botryococcus spp. ALBO 2 2 9 2

Fungal spore FU 1 4

Baculatisporites comaumensis SP 1 1 3 10

Brevitriletes bulliensis SP +

Brevitriletes cornutus SP + 1

Brevitriletes levis SP + 10 12 4 18 1 2 4

Calamospora sp. SP + 1 3 4 30 5 2 4 2

Camptotriletes warchianus SP 1 1 1 1 3

25

Cyclogranisporites gondwanensis SP 3 4 + 2

Densoisporites sp. SP 2

Dictyophillidites mortonii SP 1

Didecitriletes dentatus SP 1 1 3 1

Didecitriletes ericianus SP 4 1 3 2 3 4

Didecitriletes longispinosus SP 1

Dulhuntyispora parvithola SP 1 2 7 2 27

Grandispora segrovesii SP 1 1 4 5 14

Horriditriletes filiformis SP 4 2 3 6 4

Horriditriletes ramosus SP 12 10 14 3 5 7 8

Horriditriletes superbus SP 1 1

Horritditriletes tereteangulatus SP + 5 10 7 1 7 12 11 5

Indospora clara SP 1 9 5 2 4 4

Indotriradites niger SP 1 + 1 7

Indotriradites reidii SP 1 1 2 2 4

Interradispora daedala SP 1 1 1 3 3

Laevigatosporites scissus SP 1 1 1 3 1

Laevigatosporites vulgaris SP 1 1 3 2 2

Leiotriletes directus SP + 13 18 13 32 20 29 20 155

Lophotriletes novicus SP 3 + 2 4 3 1

Microbaculispora micronodosus SP + 7 1 3 1 4

Microbaculispora tentula SP + 3 2 2 2 4 7 9

Microbaculispora trisina SP + 3 7 2 9 1 1 42

Microbaculispora villosa SP 1

Microfoveolatispora explicita SP 1 2

26

Microreticulatisporites SP 5 bitriangularis

Osmundacidites senectus- SP 8 1 5 4 2 2 wellmanii

Punctatisporites spp. SP 2 1 2 4 1 2

Punctatisporites gretensis SP 1 2 1

Raistrickia sp. A SP 2

Raistrickia crenata SP 1

Retusotriletes diversiformis SP + 1 1 9

Retusotriletes nigritellus SP 2 2 1 9

Secarisporites bullatus SP 2

Verrucosisporites spp. SP + 1 1 1

Alisporites spp. PO 8 11 4 12 5 12 5 1

Alisporites australis PO 2 2 1

Bascanisporites undosus PO 1 1 1 1 1

Cannanoropolis janakii PO 2

Chordasporites spp. PO 1

Cycadopites follicularis PO 2 1 1

Florinites eremus PO 1

Klausipollenites sp. PO 1 1 2 1

Limitisporites rectus PO 2 1 2

Marsupipollenites striatus PO 1 1 1 2 2 1

Marsupipollenites triradiatus PO 4 1 4 1 8 7 8

Plicatipollenites malabarensis PO 1

Praecolpatites sinuosus PO 1 2 + 3 14 7 8

Protohaploxypinus amplus PO 7 7 18 5 9 7

27

Protohaploxypinus bharadwajii PO 2 3

Protohaploxypinus diagonalis PO 1 2

Protohaploxypinus haigii PO 1 1 1 3 3 3 1

Protohaploxypinus hartii PO 4 4

Protohaploxypinus limpidus PO + 47 44 53 26 30 18 41 5

Protohaploxypinus rugatus PO 1 2 3 1

Protohaploxypinus spp. (indet) PO + 6 10 8 13 3 2 7 1

Scheuringipollenites maximus PO 2 2 1 1 2

Scheuringipollenites ovatus PO 2 1 6 10 2 2 2 3

Schizopollis disaccoides PO 4 2 1 1

Striate bisaccate pollen (indet) PO + 8 11 3 8 7 3 2 3

Striatoabieites multistriatus PO 2 1 2 4 1 1 13

Striatopodocarpites cancellatus PO 12 6 16 10 18 4 8 7

Striatopodocarpites fusus PO 6 13 8 3 9 5 4 3

Striatopodocarpites phaleratus PO 1

Striomonosaccites sp. PO 1 1 2 1

Tiwarisporites simplex PO 1 1 1 1 5

Triadispora sp. PO 2 1 4 3 5 7 2 6

Vitreisporites pallidus PO 1 1 2

Vittatina fasciolata PO 11

Vittatina scutata PO 4

Weylandites lucifer PO 1 23

total 0 0 0 0 0 200 200 200 0 200 200 200 200 500 0

*Ac = acritarchs, AL = undifferentiated algae, ALPR = prasinophycean algae, ALBO = Botryococcus, FU = fungal palynomorphs, SP = spores, PO = pollen

28

B4.2 Glue Pot Creek 1

depth 582,7 592,5 602,6 606,7 615,7 620,3 621,3 622,9 651,9 658,8 662,3 664,2 698,3 712,0 712,8 714,9 716,1 718,9 719,7 5 5 5 3 5 2 0 4 4 1 0 0 6 0 8 8 3 8 9 sample palynomor GPC1 GPC2 GPC3 GPC4 GPC5 GPC6 GPC7 GPC8 GPC9 GPC1 GPC1 GPC1 GPC1 GPC1 GPC1 GPC1 GPC1 GPC1 GPC1 ph type* 0 1 2 3 4 5 6 7 8 9

Brazilea plurigenus AL + 1 1 1

Brazilea scissa AL 1 1 3

Maculatasporites sp. AL 1 2

Peltacystia venosa AL 1 1

Pilasporites calculus AL 1

Rugaletes playfordii AL 1 +

Botryococcus spp. ALBO + 28 26 + 9 + 87

Brevitriletes bulliensis SP +

Brevitriletes levis SP 4 1 1 1 2

Calamospora sp. SP + 2 29 6 13 6 2

Camptotriletes warchianus SP 2 1

Cavate Trilete Spore (indet.) SP 1

Columnisporites heyleri SP 1

Cyclogranisporites SP + 13 11 4 6 6 4 gondwanensis

Dictyophillidites mortonii SP + 1 2 1

Didecitriletes ericianus SP 1 3 2

Dulhuntyispora parvithola SP 1

Gondisporites raniganjensis SP 2

Grandispora segrovesii SP 2

29

Horriditriletes filiformis SP 3 1 2 1 2

Horriditriletes ramosus SP 1 1 1

Horriditriletes superbus SP 1

Horritditriletes SP 13 15 2 4 1 3 tereteangulatus

Indospora clara SP 1

Indotriradites reidii SP 3 2 8 1

Interradispora daedala SP 2

Leiotriletes directus SP + 2 8 5 2 4 7

Lophotriletes novicus SP 1

Microbaculispora SP 1 1 1 micronodosus

Microbaculispora tentula SP 1 14 1 1

Microbaculispora trisina SP 2 1 3

Microfoveolatispora SP 1 explicita

Microreticulatisporites SP 1 1 1 bitriangularis

Osmundacidites senectus- SP 13 18 11 5 6 2 wellmanii

Punctatisporites spp. SP 2 3

Punctatisporites gretensis SP + 2

Retusotriletes diversiformis SP 1

Thymospora ipsviciensis SP +

Verrucosisporites spp. SP 1 1

Alisporites spp. PO + 17 20 15 13 19 5

Alisporites australis PO + 2 1 2 4 8

30

Barakarites rotatus PO + 2

Bascanisporites undosus PO 2 1

Cannanoropolis janakii PO 1 5 3 1

Chordasporites spp. PO 1 1

Klausipollenites sp. PO 1 2

Marsupipollenites striatus PO 1 1 2

Marsupipollenites PO 4 3 2 3 1 triradiatus

Plicatipollenites PO 2 3 5 gondwanesis

Potonieisporites balmei PO 3

Praecolpatites sinuosus PO 3 1 4 6

Protohaploxypinus amplus PO 8 2 14 21 35 10

Protohaploxypinus PO 1 1 bharadwajii

Protohaploxypinus haigii PO 2 2 1

Protohaploxypinus hartii PO 4 3 3

Protohaploxypinus limpidus PO 59 30 42 52 33 15

Protohaploxypinus rugatus PO 1

Protohaploxypinus spp. PO + 2 4 12 13 5 5 (indet)

Scheuringipollenites PO 4 8 3 3 7 1 maximus

Scheuringipollenites ovatus PO 14 5 1 5 3

Schizopollis disaccoides PO 2

Striate bisaccate pollen PO 2 6 3 3 1 1 (indet)

31

Striatoabieites multistriatus PO 4 1 3 4 2

Striatopodocarpites brevis PO 1 2 1

Striatopodocarpites PO 11 5 5 3 7 1 cancellatus

Striatopodocarpites fusus PO + 1 2 1 4 5 3

Striatopodocarpites PO 2 1 phaleratus

Triadispora sp. PO + 3

Vitreisporites pallidus PO 7 1 5 3

Weylandites lucifer PO 2

total 0 0 0 0 0 200 200 0 200 0 200 0 200 0 0 0 200 0 0

*Ac = acritarchs, AL = undifferentiated algae, ALPR = prasinophycean algae, ALBO = Botryococcus, FU = fungal palynomorphs, SP = spores, PO = pollen

B4.3 Tambo 1-1A (standard processing)

depth 69 69 69 69 69 69 70 70 70 70 70 70 72 72 72 73 73 73 73 73 74 74 770, 781, 791, 803, 810, 821, 827, 828, 829, 829, 6,6 7,9 8,3 8,8 9,1 9,3 0,1 0,2 0,4 0,5 0,6 1,7 3,7 9,0 9,2 3,9 5,0 5,4 8,3 8,8 0,5 1,6 48 02 88 10 22 11 01 23 11 28 2 7 6 8 5 5 6 6 0 2 7 2 6 3 3 3 5 4 7 4 7 7

sample palyno TA TA TA TA TA TA TA TA TA TA TA TA TA TA TA TA TA TA TA TA TA TA TA TA TA TA TA TA TA TA TA TA morph MP MP MP MP MP MP MP MP MP MP MP MP MP MP MP MP MP MP MP MP MP MP MP MP MP MP MP MP MP MP MP MP type* -1 -2 -3 -4 -5 -6 -7 -8 -9 -10 -11 -12 -13 -14 -15 -16 -17 -18 -19 -20 -21 -22 23 24 25 26 27 28 29 30 31 32

Micrhystridi AC 1 3 137 151 um evansii

Micrhystridi AC 1 5 1 1 1 7 9 2 1 um sp.

Brazilea AL 1 helbyi

Brazilea AL 1 1 2 3 1 1 plurigenus

Brazilea AL 1 1 1 2 3 5 6 5 7 scissa

32

Circulisporit AL 1 es spp.

Peltacystia AL 1 venosa

Pilasporites AL 2 1 1 1 calculus

Rugaletes AL 1 playfordii

Tetraporina AL 1 sp.

Cymatiosph ALPR 1 1 1 1 1 2 1 aera gondwanen sis

Leiosphaeri ALPR 1 2 dia spp.

Botryococcu ALBO 3 4 1 1 1 1 1 2 1 13 s spp.

Fungal FU 1 3 2 1 1 spore

Reduviaspo FU 1 1 3 ronites chalasta

Baculatispo SP 1 1 3 1 2 5 4 1 1 rites comaumens is

Brevitriletes SP 21 1 2 5 4 bulliensis

Brevitriletes SP 2 1 1 1 1 2 2 1 1 cornutus

Brevitriletes SP 4 6 13 8 2 5 2 1 2 1 4 4 3 2 levis

Calamospor SP 18 1 3 1 2 1 1 3 5 1 1 9 6 8 4 3 2 4 1 a sp.

Cavate SP 2 3 1 1 1 Trilete Spore (indet.)

Concavissim SP 7 isporites grumulus

33

Cyathidites SP 8 6 5 breviradiatu s

Cyclogranis SP 2 2 1 8 1 2 1 9 7 5 4 2 5 6 5 2 1 7 9 7 2 1 porites gondwanen sis

Densoispori SP 9 1 tes playfordii

Dictyophilli SP 2 1 dites mortonii

Dictyotrilet SP 1 1 1 1 1 1 1 es aules

Didecitrilete SP 1 1 1 2 1 2 3 8 2 1 2 s dentatus

Didecitrilete SP 1 3 1 4 4 4 2 3 2 6 11 1 3 3 7 4 6 2 2 2 s ericianus

Didecitrilete SP 1 1 2 2 1 1 1 1 1 s longispinos us

Dulhuntyisp SP ora dulhuntyi

Dulhuntyisp SP 1 8 5 1 1 1 ora parvithola

Foveosporit SP 1 es sp. A

Gondisporit SP 1 2 1 es raniganjens is

Grandispor SP 1 1 1 1 a segrovesii

Granulatisp SP orites absonus

Horriditrilet SP 5 4 4 3 6 3 10 3 1 6 5 1 8 3 14 3 3 2 2 1 1 es filiformis

Horriditrilet SP 1 es

34 gondwanen sis

Horriditrilet SP 4 2 2 7 1 5 2 2 4 1 2 1 2 10 8 4 9 1 es ramosus

Horriditrilet SP 1 1 1 1 es superbus

Horriditrilet SP 16 12 17 9 7 5 7 2 10 7 8 10 11 18 24 8 21 22 12 3 3 es tereteangul atus

Indospora SP 3 2 2 1 2 1 3 4 1 clara

Indotriradit SP 1 2 es niger

Indotriradit SP 2 2 1 2 es reidii

Indotriradit SP es splendens

Interradispo SP 2 3 2 4 1 2 3 2 2 1 1 ra daedala

Laevigatosp SP 3 2 1 1 orites scissus

Laevigatosp SP 2 2 1 1 2 1 1 1 orites vulgaris

Leiotriletes SP 44 13 21 23 12 16 6 5 5 4 5 9 7 26 27 22 29 65 24 74 8 7 directus

Limatulaspo SP 2 rites limatulus

Lophotrilete SP 4 4 2 2 4 1 2 2 1 4 7 1 3 1 2 4 4 6 3 2 1 s novicus

Lundbladisp SP 2 ora sp. A

Lundbladisp SP 5 2 1 1 ora iphilegna

Lundbladisp SP 2 ora

35 springsuren sis

Lundbladisp SP 1 1 1 ora willmotii

Microbaculi SP 1 1 2 2 1 1 5 3 4 9 4 7 3 spora micronodos us

Microbaculi SP 4 2 3 2 5 6 1 3 4 2 8 4 9 7 5 6 3 2 spora tentula

Microbaculi SP 8 5 5 3 7 6 6 10 6 1 5 3 2 3 3 5 8 1 1 5 2 1 spora trisina

Microbaculi SP 1 2 1 3 spora villosa

Microfoveol SP 1 4 6 1 5 3 1 3 2 2 1 2 1 1 1 atispora explicita

Microreticul SP 1 1 2 1 2 1 1 4 atisporites bitriangular is

Osmundaci SP 4 3 5 7 5 3 6 8 9 9 12 4 6 9 4 10 3 1 2 1 1 dites senectus- wellmanii

Phaselispori SP tes cicatricosus

Playfordias SP 2 5 pora crenulata

Polycingulat SP 1 1 isporites dejerseyi

Polypodiisp SP 7 1 orites mutabilis cf.

Pseudoretic SP 1 ulatispora barakarensi s

36

Pseudoretic SP 1 2 ulatispora pseudoretic ulata

Punctatispo SP 11 3 6 6 1 3 5 2 2 1 4 3 2 1 1 2 1 rites spp.

Punctatispo SP 4 4 4 2 2 5 1 2 1 1 rites gretensis

Raistrickia SP 1 2 1 crenata

Retusotrilet SP 1 es sp. A

Retusotrilet SP 7 3 1 1 1 2 3 5 3 4 es diversiformi s

Retusotrilet SP 10 1 1 3 4 3 2 2 1 es nigritellus

Secarisporit SP 1 es bullatus

Thymospor SP 2 14 6 3 a ipsviciensis

Triplexispori SP 6 tes playfordii

Triquitrites SP 3 1 proratus

Verrucosisp SP 7 1 1 1 orites spp.

Alisporites PO 3 10 10 23 17 34 14 26 22 19 24 19 25 8 8 4 2 2 6 5 2 2 spp.

Alisporites PO 1 5 4 australis

Barakarites PO rotatus

Bascanispor PO 1 1 1 1 1 2 2 ites undosus

37

Cannanoro PO 1 1 1 polis janakii

Chordaspori PO tes spp.

Cycadopites PO 3 1 1 3 1 2 1 1 1 1 1 follicularis

Florinites PO 1 2 1 3 1 2 2 1 3 2 eremus

Klausipollen PO 3 1 4 1 1 2 2 ites sp.

Limitisporit PO 2 2 4 3 1 1 es rectus

Lunatisporit PO 1 1 es sp. A

Marsupipoll PO 1 2 4 1 1 1 3 3 2 1 4 2 1 1 enites striatus

Marsupipoll PO 3 3 5 6 11 7 7 8 8 4 14 9 2 3 1 7 4 2 1 enites triradiatus

Parasaccite PO 1 1 1 1 2 3 1 s diffusus

Plicatipollen PO 1 ites gondwanesi s

Plicatipollen PO 1 ites malabarens is

Praecolpatit PO 3 7 4 5 1 4 6 3 3 2 2 1 1 5 3 2 1 es sinuosus

Protohaplox PO 1 4 9 5 8 13 9 6 6 5 13 5 5 8 3 1 2 2 1 ypinus amplus

Protohaplox PO 2 1 1 1 1 1 1 ypinus bharadwajii

Protohaplox PO 2 1 ypinus diagonalis

38

Protohaplox PO 2 1 1 3 3 2 2 2 4 2 2 ypinus haigii

Protohaplox PO 2 3 5 2 4 1 3 1 2 1 3 1 1 1 1 1 ypinus hartii

Protohaplox PO 7 3 9 18 5 28 20 17 19 23 17 23 37 32 21 18 5 16 14 4 5 ypinus limpidus

Protohaplox PO 1 ypinus pennatulus

Protohaplox PO 1 1 1 1 6 9 9 10 5 4 5 1 1 ypinus rugatus

Protohaplox PO 6 3 3 6 4 3 5 ypinus samoilovich ii

Protohaplox PO 1 6 8 5 18 7 9 6 8 13 8 3 4 6 3 2 5 3 1 ypinus spp. (indet)

Scheuringip PO 1 3 2 1 2 1 2 2 2 2 1 2 1 2 1 ollenites maximus

Scheuringip PO 7 2 3 8 10 5 6 3 7 10 4 4 3 4 6 1 4 12 4 2 2 ollenites ovatus

Schizopollis PO 3 disaccoides

Striate PO 2 7 5 9 18 10 18 10 12 7 14 25 10 5 3 4 3 1 5 2 1 bisaccate pollen (indet)

Striatoabiei PO 1 1 tes sp. A

Striatoabiei PO 2 1 2 3 1 3 3 6 2 4 3 1 1 2 4 2 2 tes multistriatu s

Striatopodo PO 1 1 3 1 1 carpites brevis

39

Striatopodo PO 2 2 1 2 2 3 2 5 2 3 1 9 13 11 7 4 8 3 2 3 1 carpites cancellatus

Striatopodo PO 1 1 9 1 5 1 6 1 5 6 8 6 2 2 7 4 2 carpites fusus

Striatopodo PO 1 carpites phaleratus

Striatopodo PO 1 2 4 1 carpites solitus

Striomonos PO 1 1 accites sp.

Tiwarisporit PO 1 es simplex

Triadispora PO 2 3 1 1 3 3 4 8 3 5 1 sp.

Vitreisporit PO 7 12 3 3 6 8 5 9 9 7 2 es pallidus

Vittatina PO 2 spp.

Vittatina PO 1 3 1 1 1 1 1 1 fasciolata

Vittatina PO 1 1 1 scutata

Weylandite PO 1 1 s lucifer

total 20 0 20 20 0 0 0 0 20 20 20 0 0 0 0 20 20 20 20 20 20 20 0 200 200 200 200 200 200 200 200 200 0 0 0 0 0 0 0 0 0 0 0 0 0

*Ac = acritarchs, AL = undifferentiated algae, ALPR = prasinophycean algae, ALBO = Botryococcus, FU = fungal palynomorphs, SP = spores, PO = pollen

B4.4 Tambo 1-1A (acid-free processing)

40

depth 770,48 781,02 791,88 803,10 810,22 821,11 827,01 828,23 829,11 829,28 sample palynomorph TAMP23 TAMP24 TAMP25 TAMP26 TAMP27 TAMP28 TAMP29 TAMP30 TAMP31 TAMP32 type*

Micrhystridium evansii AC 1 4 152 166

Micrhystridium sp. AC 1 4 1

Brazilea helbyi AL 3

Brazilea plurigenus AL 2 4 8 2 6 3

Brazilea scissa AL 2 1 2 2 7 19 5 1

Circulisporites spp. AL 1 2 2

Peltacystia venosa AL 3 2

Pilasporites calculus AL 1 1 4 3 7 7 2

Rugaletes playfordii AL 1 1

Leiosphaeridia spp. ALPR 2 5

Botryococcus spp. ALBO 76

Baculatisporites comaumensis SP 1 2 1

Brevitriletes cornutus SP 1 1

Brevitriletes levis SP 3 3 4 2

Calamospora sp. SP 3 2 1

Cyclogranisporites gondwanensis SP 1 5 7 3 3 1 3

Dictyophillidites mortonii SP 1 1

Dictyotriletes aules SP 1

Didecitriletes dentatus SP 1 1 3 4 1 1 1 1

Didecitriletes ericianus SP 4 5 3 5 4 1 2 1

Didecitriletes longispinosus SP 1

Dulhuntyispora dulhuntyi SP 1

41

Dulhuntyispora parvithola SP 1 1 4 11 34 6 5 3

Granulatisporites absonus SP 3 1

Horriditriletes filiformis SP 4 8 4 3 12 1 2

Horriditriletes ramosus SP 1 9 1 1 10 1

Horriditriletes superbus SP 1

Horriditriletes tereteangulatus SP 3 8 21 5 6 21 6 1 2

Indospora clara SP 2 1 1

Indotriradites reidii SP 1 3 3 2

Indotriradites splendens SP 2

Interradispora daedala SP 1 3 3 1 1

Laevigatosporites vulgaris SP 1

Leiotriletes directus SP 13 14 24 19 33 19 26 7 3

Lophotriletes novicus SP 3 1 2 2

Microbaculispora micronodosus SP 2 5 11

Microbaculispora tentula SP 1 2 10 8 3 1

Microbaculispora trisina SP 7 5 3 13 12 6 4 1

Microbaculispora villosa SP 2 2 1 1

Microfoveolatispora explicita SP 2 1 3 8 1 2

Microreticulatisporites bitriangularis SP 2 3 3 1

Osmundacidites senectus-wellmanii SP 6 4 5 5 2 5 3

Phaselisporites cicatricosus SP 1

Pseudoreticulatispora pseudoreticulata SP 1

Punctatisporites spp. SP 2 1 1 2 1 1

Punctatisporites gretensis SP 1 1

Raistrickia crenata SP 1 1

42

Retusotriletes diversiformis SP 2 2

Retusotriletes nigritellus SP 1 3 2 3 4 1

Verrucosisporites spp. SP 1 3

Alisporites spp. PO 16 17 6 2 4 5 4 1

Barakarites rotatus PO 3 1

Bascanisporites undosus PO 1

Cannanoropolis janakii PO 1

Chordasporites spp. PO 1

Cycadopites follicularis PO 3 1 2

Klausipollenites sp. PO 1 1

Limitisporites rectus PO 1 4 1 3

Marsupipollenites striatus PO 1 3 1 1

Marsupipollenites triradiatus PO 2 10 3 1 2 1

Plicatipollenites gondwanesis PO 1

Praecolpatites sinuosus PO 1 3 2 7 8 9 1

Protohaploxypinus amplus PO 23 9 1 5 3 3

Protohaploxypinus bharadwajii PO 1

Protohaploxypinus diagonalis PO 1

Protohaploxypinus haigii PO 3 3 2 5

Protohaploxypinus hartii PO 1

Protohaploxypinus limpidus PO 41 25 16 4 5 18 23 3 2

Protohaploxypinus pennatulus PO 1 1

Protohaploxypinus rugatus PO 4 1 2 1

Protohaploxypinus spp. (indet) PO 8 11 2 1 3 1 1

Scheuringipollenites maximus PO 2 2

43

Scheuringipollenites ovatus PO 4 5 2 1 2 2 1

Schizopollis disaccoides PO 1 1

Striate bisaccate pollen (indet) PO 7 7 3 3 1 4 2

Striatoabieites multistriatus PO 7 7 2 10 1 10 2 2

Striatopodocarpites brevis PO 3 2 1 1 1 1

Striatopodocarpites cancellatus PO 12 12 11 2 16 6 6 1

Striatopodocarpites fusus PO 9 6 7 2 2 3 2 4

Striatopodocarpites phaleratus PO 1 1 4

Striatopodocarpites solitus PO 2 3 1

Striomonosaccites sp. PO 1 3 1

Tiwarisporites simplex PO 1 2

Triadispora sp. PO 2 1 1 7 3 5

Vitreisporites pallidus PO 3

total 202 200 200 200 200 200 200 200 200

*Ac = acritarchs, AL = undifferentiated algae, ALPR = prasinophycean algae, ALBO = Botryococcus, FU = fungal palynomorphs, SP = spores, PO = pollen

B4.5 Springsure 19

depth 320,8 346,1 359,2 359,4 360,0 360,7 363,7 378,8 384,9 385,8 388,4 391,3 399,1 421,3 446,3 468,0 484,7 492,1 505,3 510,9 515,9 523,6 0 8 8 1 5 8 0 0 5 8 8 1 3 5 6 1 5 8 5 6 5 3

sample palynomorp SPR1 SPR2 SPR3 SPR4 SPR5 SPR6 SPR7 SPR8 SPR9 SPR10 SPR11 SPR12 SPR13 SPR14 SPR15 SPR16 SPR17 SPR18 SPR19 SPR20 SPR21 SPR22 h type*

Micrhystridium sp. AC 4

Brazilea plurigenus AL 1 3 1 1

Brazilea scissa AL 1 4 1 4 1 2 1

Circulisporites spp. AL 1

Maculatasporites sp. AL 1 1 1

Peltacystia venosa AL 1 3 2 1

44

Pilasporites calculus AL 1 2 1 1 1 2

Quadrisporites horridus AL 1

Tetraporina sp. AL 1

Leiosphaeridia spp. ALPR 1 1 1

Mehlisphaeridium fibratum ALPR 1

Botryococcus spp. ALBO 12 5 37 1 1 7 21 3 43 5 24 11 59 2

Fungal spore FU 1 2 1

Reduviasporonites chalasta FU 14

Baculatisporites SP 1 1 8 1 1 4 2 2 comaumensis

Brevitriletes bulliensis SP 42 48 11

Brevitriletes cornutus SP 2 2

Brevitriletes levis SP 16 13 9 4 2 10 2 3 3 2 1 1 5

Calamospora sp. SP 6 23 13 1 2 10 1 1 2 7 5 6 16 9 5 15 6

Camptotriletes warchianus SP 1 2 2 4

Cavate Trilete Spore (indet.) SP 4 9 2 1 1 1 1

Columnisporites heyleri SP 1 1 2

Concavissimisporites SP 3 2 8 1 grumulus

Converrucosisporites sp. SP 1 14

Cyclogranisporites SP 2 4 3 12 2 3 5 1 4 1 7 5 gondwanensis

Densoisporites playfordii SP 4

Densoisporites sp. SP 1 5

Dictyophillidites mortonii SP 1 1

Dictyotriletes aules SP 1 1

Didecitriletes dentatus SP 1 1 1 11 4 14 13 2

Didecitriletes ericianus SP 1 1 1 2 4 1 2 1 1 7 12 8 16 5 3 5

Didecitriletes longispinosus SP 1 3 2

Dulhuntyispora dulhuntyi SP 1

45

Dulhuntyispora parvithola SP 1 1 3 6 1 3 9 1 7 79 4 3

Gondisporites raniganjensis SP 1

Grandispora segrovesii SP 2 1 1 1 4 1

Granulatisporites absonus SP 1 1 1 3 4 3

Horriditriletes filiformis SP 1 1 4 1 1 2 1 3 5 4 2 11 1 6

Horriditriletes ramosus SP 7 2 1 5 2 4 2 6 5 7 14 8

Horriditriletes SP 5 6 10 2 1 1 6 13 8 3 5 9 5 5 tereteangulatus

Indospora clara SP 3 1 1 1

Indotriradites niger SP 1

Indotriradites reidii SP 3 2 2 1 1 3 1 1 1

Indotriradites splendens SP 3 12 3

Interradispora daedala SP 1 1 1 1 2 1 1

Laevigatosporites scissus SP 2

Laevigatosporites vulgaris SP 1 3 1

Leiotriletes directus SP 8 10 7 11 27 27 13 15 3 10 11 15 26 17 14 3 15 7

Lophotriletes novicus SP 2 2 1 1 1 1 1 1 5 2 1 1 2 1

Limbosporites balmei SP 1 1

Lundbladispora iphilegna SP 1

Lundbladispora SP 5 10 1 springsurensis

Lycopodiumsporites SP 2 22 3 "crassus"

Microbaculispora SP 2 2 1 8 4 7 5 2 6 2 3 2 1 4 1 4 1 micronodosus

Microbaculispora tentula SP 2 4 4 7 17 1 1 3 3 2 2 3 1 4 2 17 3

Microbaculispora trisina SP 2 1 9 5 9 5 2 1 1 6 4 6 1 2 8 14 1

Microfoveolatispora SP 1 2 2 explicita

Microreticulatisporites SP 2 1 4 1 bitriangularis

46

Osmundacidites senectus- SP 1 1 6 7 14 6 2 8 13 10 2 1 11 4 9 wellmanii

Phaselisporites cicatricosus SP 4

Playfordiaspora crenulata SP 1 1

Polypdiisporites leopardus SP 1

Punctatisporites spp. SP 2 3 2 5 3 1 4 1 3

Punctatisporites gretensis SP 2 5 4 2 1 4

Raistrickia crenata SP 1 1

Retusotriletes diversiformis SP 4 3 1 1 4 1 3 1 1

Retusotriletes nigritellus SP 2 1 1 3 1 1 1 1

Thymospora ipsviciensis SP 4 5

Triplexisporites playfordii SP 2 6 3

Verrucosisporites spp. SP 19 4 5 6 35 2 1 2

Alisporites spp. PO 16 1 24 6 5 5 4 5 7 7 1 6 2 1 1

Alisporites australis PO 3 4

Bascanisporites undosus PO 1 2

Cannanoropolis janakii PO 1 1 3 2 1 1 1

Cycadopites follicularis PO 1

Florinites eremus PO 1

Guttulapollenites sp. PO 2

Krauselisporites sp. PO 1

Limitisporites rectus PO 3 1 5

Marsupipollenites striatus PO 1 1 2 1 3 1 1 1

Marsupipollenites PO 1 1 11 8 1 1 3 4 3 12 2 13 4 4 7 triradiatus

Plicatipollenites PO 1 1 gondwanesis

Potoniesporites balmei PO 1 1

Praecolpatites sinuosus PO 7 2 5 1 1 1 4 4

Protohaploxypinus amplus PO 9 2 11 14 6 18 15 8 5 7 2 15

47

Protohaploxypinus PO 1 2 2 1 3 1 1 bharadwajii

Protohaploxypinus PO 1 1 1 diagonalis

Protohaploxypinus haigii PO 1 1 2 2 2 1 3 2 2 7

Protohaploxypinus hartii PO 1 1 3 6 2 2 3 1 1

Protohaploxypinus limpidus PO 3 53 24 39 50 43 67 56 43 40 9 61 14 1 10 60

Protohaploxypinus rugatus PO 1 1 3

Protohaploxypinus PO 2 samoilovichii

Protohaploxypinus spp. PO 1 2 7 6 4 3 2 7 3 1 5 11 (indet)

Scheuringipollenites PO 4 4 3 2 2 7 2 1 1 maximus

Scheuringipollenites ovatus PO 8 1 6 6 10 4 4 3 9 2 10 2 3 5

Schizopollis disaccoides PO 2 1 1

Striate bisaccate pollen PO 3 2 1 7 4 1 3 3 1 1 1 2 2 (indet)

Striatoabieites sp. A PO 1

Striatoabieites multistriatus PO 1 1 3 1 1 1 1 3 1 1 2 2 2

Striatopodocarpites brevis PO 1 1 1 1 1 2

Striatopodocarpites PO 18 5 14 14 6 25 12 24 14 7 8 9 1 10 11 cancellatus

Striatopodocarpites fusus PO 13 3 10 11 7 19 7 9 4 11 5 1 21 4

Striatopodocarpites PO 1 2 1 1 1 5 1 phaleratus

Striatopodocarpites solitus PO 2

Tiwarisporites simplex PO 1 1 1 2 1 1

Triadispora sp. PO 1 1 1 1 9

Vitreisporites pallidus PO 2 1 1 3 1 1 1

total 0 0 200 200 200 0 200 200 200 0 200 200 200 200 200 200 200 200 200 200 200 200

*Ac = acritarchs, AL = undifferentiated algae, ALPR = prasinophycean algae, ALBO = Botryococcus, FU = fungal palynomorphs, SP = spores, PO = pollen

48

B4.6 Taringa 7

depth 949,01 952,17 954,2 958,29 960,7 963,34 967,06 970,6 992,84 1011,9 1039,15 1063,1 1069,27 1072 sample palynomorph TAR 1 TAR 2 TAR 3 TAR 4 TAR 5 TAR 6 TAR 7 TAR 8 TAR 9 TAR 10 TAR 11 TAR 12 TAR 13 TAR 14 type*

Micrhystridium sp. AC 48 8 13

Brazilea plurigenus AL 2 2 1 1 7

Brazilea scissa AL 1 7 1 1 3 1

Circulisporites spp. AL 1

Brazilea helbyi AL 1

Peltacystia venosa AL 1

Pilasporites calculus AL 1

Botryococcus spp. ALBO 1 3 25

Cymatiosphaera gondwanensis ALPR 1

Leiosphaeridia spp. ALPR 1 1

Fungal spore FU 1 1

Alisporites australis PO 1

Protohaploxypinus microcorpus PO +

Leuckisporites sp. PO 1

Protohaploxypinus rugatus PO 1 1

Protohaploxypinus bharadwajii PO 1 1

Striatopodocarpites phaleratus PO 2 1 1 1 2

Striatopodocarpites solitus PO 1 1 1 2 1

Protohaploxypinus amplus PO 12 1 9 8 5 2 1

Protohaploxypinus haigii PO 1 3 3 1

Protohaploxypinus hartii PO 5 1 8 1 1 3 1

49

Protohaploxypinus spp. (indet) PO 4 8 6 3 1 1

Scheuringipollenites ovatus PO 7 8 3 3 1 1

Striatopodocarpites fusus PO 4 1 2 3 3 1 3 4

Alisporites spp. PO 12 2 4 6 4 6 7 5 5

Marsupipollenites triradiatus PO 2 1 1 5 12 10 9

Protohaploxypinus limpidus PO 76 6 61 31 13 15 8 10 6

Scheuringipollenites maximus PO 2 1 1 1 3 1 1 1

Striatopodocarpites cancellatus PO 10 10 4 6 7 1 4 1

Cycadopites follicularis PO 2

Striatoabieites multistriatus PO 1 1 3 12 7 4 3 1

Triadispora sp. PO 3 1 1 11 2 2 5 2

Protohaploxypinus diagonalis PO 2

Schizopollis disaccoides PO 1

Vitreisporites signatus PO 1 2

Marsupipollenites striatus PO 1 2 1 1

Platysaccus sp. PO 1 2 1

Praecolpatites sinuosus PO 2 2 1 1 1

Klausipollenites sp. PO 1

Plicatipollenites gondwanesis PO 1

Bascanisporites undosus PO 1 1

Plicatipollenites densus PO 1 1 1

Striate bisaccate pollen (indet) PO 3 1 1 4

Striomonosaccites sp. PO 1

Limitisporites rectus PO 1 2 4

Tiwarisporites simplex PO 1 1

50

Cannanoropolis janakii PO 2 3 1

Barakarites rotatus PO 1

Potonieisporites methoris PO 1

Dictyophillidites mortonii SP 1

Lundbladispora springsurensis SP 1

Triplexisporites playfordii SP 1

Calamospora sp. SP 1 4 4 1 6 1 1 5 6

Horriditriletes filiformis SP 1 2 7 4 2 1 5 1 1 5

Horriditriletes ramosus SP 1 1 3 2 5 1 1 1 1 2

Laevigatosporites flexus SP 1 3 2 1 1

Brevitriletes cornutus SP 1

Columinisporites sp. SP 1

Granulatisporites absonus SP 3 2

Indotriradites splendens SP 1 5 1

Microreticulatisporites SP 1 5 2 bitriangularis

Brevitriletes levis SP 3 9 5 2

Osmundacidites senectus- SP 2 7 10 3 4 3 wellmanii

Cyclogranisporites gondwanensis SP 2 3 8 2 2 2 5 1 3

Horriditriletes tereteangulatus SP 6 3 13 7 1 5 5 5 4

Leiotriletes directus SP 12 67 25 48 79 36 19 14 11

Microbaculispora micronodosus SP 5 6 1 6 4 6 6 2 2

Microbaculispora tentula SP 7 2 3 8 3 1 7 4

Microbaculispora trisina SP 2 5 2 3 14 3 5 4 6

Punctatisporites gretensis SP 1 1 1 2

51

Punctatisporites spp. SP 1 1 1 1 3 4

Retusotriletes diversiformis SP 1 9 1 2 1 3 2 1

Horriditriletes gondwanensis SP 2

Limatulasporites limatulus SP 1

Horriditriletes superbus SP 1 1 1

Indotriradites reidii SP 5 1

Interradispora daedala SP 4 1 6 1 6 1 3

Didecitriletes dentatus SP 5 2 6 8 6 11

Didecitriletes ericianus SP 17 3 4 10 31 34 23

Didecitriletes longispinosus SP 7 1 2

Didecitriletes uncinatus SP 1 3 1 1 1

Microfoveolatispora explicita SP 3 1 1

Retusotriletes nigritellus SP 2 1 1 4 1

Lophotriletes novicus SP 1 1 1 1 1

Indospora clara SP 1 1 1

Microbaculispora villosa SP 1 1 1 2 3

Verrucosisporites spp. SP 1 2 1

Dulhuntyispora parvithola SP 3 7 35 48

Dictyotriletes aules SP 1 1 2

Cavate Trilete Spore (indet.) SP 1

Phaselisporites cicatricosus SP 1

total 0 0 9 200 200 0 0 200 200 200 202 200 205 200

*Ac = acritarchs, AL = undifferentiated algae, ALPR = prasinophycean algae, ALBO = Botryococcus, FU = fungal palynomorphs, SP = spores, PO = pollen

52

B4.7 CGIE0144 depth 35,10 41,10 44,35 45,57 sample palynomorph IPA1 IPA2 IPA3 IPA4 type*

Rugaletes playfordii AL 1

Fungal spore FU 1

Reduviasporonites chalasta FU 1

Baculatisporites comaumensis SP 1

Brevitriletes bulliensis SP 28 1

Brevitriletes cornutus SP 1

Brevitriletes levis SP 4 3

Calamospora sp. SP 13 1

Camptotriletes warchianus SP 1

Cyclogranisporites gondwanensis SP 2 15

Didecitriletes ericianus SP 1

Granulatisporites absonus SP 2

Horriditriletes filiformis SP 1

Horriditriletes ramosus SP 2 1

Horriditriletes tereteangulatus SP 3 1

Indospora clara SP 1

Indotriradites niger SP 1

Indotriradites splendens SP 2

53

Leiotriletes directus SP 31 6

Lundbladispora iphilegna SP 1

Lundbladispora springsurensis SP 8

Microbaculispora micronodosus SP 1 1

Microbaculispora tentula SP 8 8

Microbaculispora trisina SP 8 6

Microbaculispora villosa SP 1

Microfoveolatispora explicita SP 3

Microreticulatisporites bitriangularis SP 1

Osmundacidites senectus-wellmanii SP 9

Playfordiaspora crenulata SP 1

Punctatisporites spp. SP 9 4

Punctatisporites gretensis SP 1

Retusotriletes diversiformis SP 1

Triplexisporites playfordii SP 10

Verrucosisporites spp. SP 1

Alisporites spp. PO 16 15

Alisporites australis PO 14

Leuckisporites spp. PO 1

Lunatisporites noviaulensis PO 2

Lunatisporites pellucidus PO 5

Marsupipollenites triradiatus PO 3

54

Praecolpatites sinuosus PO 3

Protohaploxypinus amplus PO 12

Protohaploxypinus hartii PO 1

Protohaploxypinus limpidus PO 59

Protohaploxypinus microcorpus PO 6

Protohaploxypinus pennatulus PO 1

Protohaploxypinus spp. (indet) PO 5

Scheuringipollenites maximus PO 2

Scheuringipollenites ovatus PO 9

Striate bisaccate pollen (indet) PO 10 5

Striatoabieites multistriatus PO 1 4

Striatopodocarpites brevis PO 3

Striatopodocarpites cancellatus PO 2 8

Striatopodocarpites fusus PO 1 3

Triadispora sp. PO 2

total 0 200 0 200

*Ac = acritarchs, AL = undifferentiated algae, ALPR = prasinophycean algae, ALBO = Botryococcus, FU = fungal palynomorphs, SP = spores, PO = pollen

B4.8 CGIN0067

depth 107,28 109,70 109,90 110,61 113,30

55 sample palynomorph IPB1 IPB2 IPB3 IPB4 IPB5 type*

Reduviasporonites chalasta FU 1

Baculatisporites comaumensis SP 1

Brevitriletes bulliensis SP 18 17

Brevitriletes cornutus SP 2

Brevitriletes levis SP 7 4

Calamospora sp. SP 1 12

Camptotriletes warchianus SP 2 1

Cavate Trilete Spore (indet.) SP 4 15

Concavissimisporites grumulus SP 6

Cyathidites breviradiatus SP 2

Cyclogranisporites gondwanensis SP 1 2

Densoisporites playfordii SP 15 15

Densoisporites sp. SP 3

Dictyophillidites mortonii SP 1

Horriditriletes ramosus SP 4

Horriditriletes tereteangulatus SP 1 5

Indospora clara SP 1 1

Indotriradites splendens SP 1 1

Krauselisporites saeptatus SP 1

Leiotriletes directus SP 14 17

56

Limatulasporites fossulatus SP 5 2

Limatulasporites limatulus SP 5 1

Lophotriletes novicus SP 3 1

Lundbladispora sp. A SP 11 16

Lundbladispora iphilegna SP 2 2

Lundbladispora springsurensis SP 4 7

Microbaculispora micronodosus SP 8 3

Microbaculispora tentula SP 17 7

Microbaculispora trisina SP 13 5

Osmundacidites senectus-wellmanii SP 2 1

Playfordiaspora crenulata SP 2

Polypodiisporites mutabilis cf. SP 1

Punctatisporites spp. SP 3 5

Rewanispora foveolata SP 13 11

Secarisporites bullatus SP 2

Thymospora ipsviciensis SP 5

Triplexisporites playfordii SP 1 12

Verrucosisporites spp. SP 1

Alisporites spp. PO 22 4

Alisporites australis PO 4 1

Lunatisporites noviaulensis PO 1 2

Lunatisporites pellucidus PO 1 1

57

Lunatisporites sp. A PO 1

Protohaploxypinus microcorpus PO 1

Protohaploxypinus samoilovichii PO 2 1

Protohaploxypinus spp. (indet) PO 1 1

Striate bisaccate pollen (indet) PO 1 4

Vitreisporites pallidus PO 2 2

Weylandites lucifer PO 1

total 200 0 0 0 200

*Ac = acritarchs, AL = undifferentiated algae, ALPR = prasinophycean algae, ALBO = Botryococcus, FU = fungal palynomorphs, SP = spores, PO = pollen

B4.9 A41859

depth 28,98 36,2 50,83 57,08 58,2 58,95 81,54 109,37 128,34 132,1 138,57 199,11 214,65 sample palynomorph type* UD1 UD2 UQ1 UD3 UD4 UD5 UD6 UD7 UQ2 UQ3 UQ4 UQ5 UQ6

Micrhystridium evansii AC 4 1

Rugaletes playfordii AL 1 1

Maculatasporites sp. AL 1

Leiosphaeridia spp. ALPR 1

Reduviasporonites chalasta FU + +

Densoisporites playfordii SP 3

Gondisporites bharadwadjii SP 1

Indotriradites niger SP 7

Indotriradites rallus SP 7

Indotriradites splendens SP 10

Limatulasporites fossulatus SP 3

Limatulasporites limatulus SP 3

Lundbladispora sp. A SP 4

58

Polycingulatisporites dejerseyi SP 1

Secarisporites bullatus SP 1

Thymospora ipsviciensis SP 3

Triquitrites proratus SP 1

Concavissimisporites grumulus SP 1 11

Brevitriletes bulliensis SP 14 + 18 13

Cavate Trilete Spore (indet.) SP 7 1 3

Lundbladispora iphilegna SP 2 1 1

Lundbladispora springsurensis SP 15 1 16

Playfordiaspora crenulata SP 7 + 3 9

Triplexisporites playfordii SP 12 11 21

Camptotriletes warchianus SP 8 2 1 2

Retusotriletes diversiformis SP 1 1 3 5

Retusotriletes nigritellus SP 1 3 7 11

Calamospora sp. SP 1 13 23 3 2 7 4 7

Dictyophillidites mortonii SP 4 1 1

Indospora clara SP 3 4 1 1

Brevitriletes levis SP 6 17 25 4 4 4 2 4 6

Cyathidites breviradiatus SP 2 6 1 2 1

Cyclogranisporites gondwanensis SP 1 5 2 2 1 1 2 2 6

Didecitriletes ericianus SP 1 4 2 1 1 1

Horriditriletes filiformis SP 1 + 3 9 2 1 1 6 5

Horriditriletes ramosus SP 4 11 4 2 1 2 5 1

Horriditriletes superbus SP 1 1 1 2 1 1 1

Horriditriletes tereteangulatus SP 4 13 2 4 3 1 1 7 9

Interradispora daedala SP 2 7 3 4 8 3 3

Leiotriletes directus SP 21 + 23 24 20 55 62 3 9 48

Lophotriletes novicus SP 1 1 2

Microbaculispora micronodosus SP 2 1 5 19 12 9 8 3

Microbaculispora trisina SP 1 6 1 2 2 2 4 4 7

Osmundacidites senectus-wellmanii SP 1 5 8 12 13 13

Punctatisporites spp. SP 10 7 14 4 5 3 2 1

59

Raistrickia crenata SP 6 2 1 1

Verrucosisporites spp. SP 2 + 1 1 1 1

Indotriradites reidii SP + 1 2

Granulatisporites absonus SP 3

Limbosporites balmei SP 1

Baculatisporites comaumensis SP 1 2 2

Microbaculispora tentula SP 5 4 4 22 25 12 15 6

Brevitriletes cornutus SP 2 6

Didecitriletes longispinosus SP 1

Punctatisporites gretensis SP 2 1 4 1

Microreticulatisporites bitriangularis SP 2 1 ? 1

Microfoveolatispora explicita SP 1 1 1

Dictyotriletes aules SP 1

Pseudoreticulatispora pseudoreticulata SP 1

Lunatisporites noviaulensis PO 1

Lunatisporites pellucidus PO 2

Protohaploxypinus microcorpus PO 1

Protohaploxypinus samoilovichii PO 2

Alisporites australis PO 4 3 2 2 2 1 2

Striate bisaccate pollen (indet) PO 1 4 5 6 1

Alisporites spp. PO 1 + 8 5 17 6 8 36 11 25

Protohaploxypinus spp. (indet) PO 1 2 3 11 4 1 3 17 5

Striatopodocarpites cancellatus PO 1 11 1 2

Striatoabieites multistriatus PO 1 3 3 11 8 2 11

Striatopodocarpites brevis PO 1 4 1 2 3

Vitreisporites pallidus PO 2 1 11 5 11 4 4

Schizopollis disaccoides PO 2 2

Protohaploxypinus limpidus PO 2 45 7 3 26 25 10

Triadispora sp. PO 1 1 11 1

Klausipollenites sp. PO 3

Protohaploxypinus bharadwajii PO 1

Protohaploxypinus diagonalis PO 1

60

Protohaploxypinus rugatus PO 2

Vittatina fasciolata PO 1

Bascanisporites undosus PO 1 2 1

Praecolpatites sinuosus PO 5 1 1 1

Scheuringipollenites ovatus PO 2 3 3 7 2

Marsupipollenites triradiatus PO 2 8 5 23 8 15

Protohaploxypinus amplus PO 5 1 1 3 8 2

Protohaploxypinus haigii PO 1 1 1

Striatopodocarpites fusus PO 5 3 8 4

Marsupipollenites striatus PO 2 2 1 1

Protohaploxypinus pennatulus PO 1

Chordasporites spp. PO 4 2

Striomonosaccites sp. PO 1 1

Protohaploxypinus hartii PO 1

Barakarites rotatus PO 1 1

Striatopodocarpites rarus PO 1

Total 0 0 200 0 200 200 200 0 200 199 200 200 200 *Ac = acritarchs, AL = undifferentiated algae, ALPR = prasinophycean algae, ALBO = Botryococcus, FU = fungal palynomorphs, SP = spores, PO = pollen

B5. Palynofacies data (raw counts)

B5.1 Tambo 1-1A (samples 1 – 22)

TAM TAM TAM TAM TAM TAM TAM TAM TAM TAM TAM TAM TAM TAM TAM TAM TAM TAM TAM TAM TAM TAM P1A P2A P3A P4A P5A P6A P7A P8A P9A P10 P11 P12 P13 P14 P15 P16 P17 P18 P19 P20 P21 P22 A A A A A A A A A A A A A

Opaque Lath 47 55 22 19 39 23 37 25 42 20 12 35 22 36 24 29 45 42 27 35 29 69

PHYTOCLASTS

61

Opaque 190 208 120 115 180 227 145 105 78 75 62 153 130 171 108 161 106 136 89 95 56 130

Equidimensional

Opaque Subrounded- 24 28 15 15 27 11 37 35 37 23 13 22 15 10 9 13 15 34 16 23 16 31

subangular

Striped 11 5 4 11 2 4 5 15 11 12 2 10 3 8 4 8 8 3 6 3

Striates 1 1 1 5 1 1 9 6 2 3 5 13 4 7 10 3 1 5

Banded 5 3 1 2 14 1 1 3 5 7 4 4 8 4 3

Pitted 1 1 1 3 1 12 3 2 1 1 2 4 20 3 3 2 1

Non-biostructured 13 19 118 110 28 27 33 53 76 98 155 55 90 71 119 85 70 49 45 22 45 21

Cuticles/membranes 3 1 14 11 2 4 14 21 37 2 15 2 3 3 8 9 19 24 25 3

ORGANIC DP (Degraded 18 23 9 61 4 8 MATTER Particles)

AOM (homogenous 2 7 11 9 11 6 12 21 37 12 2 1 1 1 4 4 2 macrophytes)

Spores (trilete) 13 5 16 19 8 10 15 8 9 9 6 8 7 8 8 23 19 22 24 31 28 15

PALYNOMORPHS

Spores (other) 1 1 1 1 2 2 3 4 1 8 9 3 4

Pollen (monosaccate) 1 1 1 2 1 2 4 7

Pollen (bisaccate) 1 5 2 1 1 4 1 2 2 6 1 3 9 12 18 40 50 43 10

striate

Pollen (bisaccate) non- 5 3 2 3 4 2 2 6 2 2 3 5 2 9 9 16 13 22 32 1

striate

Indeterminate 1 1 4 6 8 5 2 1 1 1 2 5 3 1 8 16 21 48

Spore/Pollen

Fresh-water algae 1 1 1 1 1 1

62

Architarchs 1 1 1 1

Fungal hyphae and 1 1 1 1 spores

total 310 327 332 326 313 332 310 318 310 301 335 333 316 315 303 359 332 363 307 346 322 332

B5.2 Tambo 1-1A (Samples 23 – 32)

Standard Processing

ACID TAMP23 TAMP24 TAMP25 TAMP26 TAMP27 TAMP28 TAMP29 TAMP30 TAMP31 TAMP32

Op 174 214 170 127 245 224 179 138 267 72

Tra 111 212 245 133 94 132 197 114 105 119

Cut/mem 2 3 6 9 14 2 1 2 6 9

DP 233 1 2 6 1 74 147 1

AOM 3 1 3 3 2

Spore 9 36 50 101 88 51 81 76 21 27

Pollen 14 35 31 138 60 20 45 27 20 36

Algae 1 3 1 6 1 3 4 2 1

Acritarch 2 3 80 234

total 547 501 508 515 511 504 508 511 505 500

Acid-free Processing

63

ACID- TAMP23 TAMP24 TAMP25 TAMP26 TAMP27 TAMP28 TAMP29 TAMP30 TAMP31 TAMP32 Free

Op 251 292 255 296 342 222 247 269 318 198

Tra 253 221 244 165 99 234 167 199 115 153

Cut/mem 2 2 3 4 2 5 9 6 3 9

DP 26 19 1 1 1

AOM 1 2 1 2

Spore 2 17 25 33 32 60 62 15 17 12

Pollen 3 23 15 20 10 34 39 21 10 6

Algae 9 2 3 30 7 7 4 2 3

Acritarch 2 60 124 total 546 555 544 521 516 583 531 517 527 508

64

Appendix C1 - Step-by-step methodology for the acid-free palynological processing technique

1. Crush rock and sedimentary samples to a fine sand size fraction (<250 μm). 2. Place crushed samples in 100 ml beakers and add 40 ml of 10% sodium hexametaphosphate, then a) stir and heat for 30 minutes then cool and/or b) leave overnight until sediment is disaggregated. 3. Sieve disaggregated samples through 250 μm mesh sieve to remove coarse fraction, then into a 5-8 μm nylon filter mesh held over large beaker with rubber band being careful not to discard material collecting on the filter mesh. Continue rinsing 5-8 μm mesh under running tapwater to wash out fine silt and clay particles into the large beaker. Rinse collected matter (5-250 μm range particles) into large clean beakers, cover with film or foil and leave to settle overnight. 4. Drain excess supernatant from beakers and wash sediment into labelled 15 ml polypropylene (PP) centrifuge tubes with lab grade water. Fill to mark with lab grade water and vortex and invert to mix. 5. Centrifuge for 5 minutes at 3000 rpm, discard supernatant carefully and repeat water washing and centrifuging. 6. Discard supernatant carefully and add 10 ml of sodium polytungstate (s.g. 1.9) and ensure solids are fully suspended by vortex mixing and shaking. 7. Centrifuge for 30 minutes at 2500 rpm. Transfer the floating organic layer and heavy liquid into a new 15 ml tube containing approx. 9 ml lab grade water. The old tube containing clays and silts may be disposed of (though step 6 can be repeated to attempt to extract more organic material). 8. Centrifuge for 5 minutes at 3000 rpm. Carefully discard supernatant into a recycling bottle (sodium polytungstate can later be reclaimed after filtering and boiling). 9. Add 10 ml lab grade water to retained solids and mix by vortex or inverting. Repeat step 8. 10. Repeat water washing, but after centrifuging excess supernatant can be discarded. 11. Carefully wash solids into a 5 ml storage tube, fill with water and add a few drops of weak (5%) HCl to stop fungal growth. 12. Samples are now ready for storage and slide mounting.

65