Exploring molecular and hydroecology of lithifying and non-lithifying microbial mats

from hypersaline lakes at Rottnest Island, Western Australia

by

Juliana Mendes Monteiro

MSc, BSc

This thesis is presented for the degree of Doctor of Philosophy of The University of Western

Australia

School of Agriculture and Environment

Environmental Sciences

2020

THESIS DECLARATION

I, Juliana Mendes Monteiro, certify that:

This thesis has been substantially accomplished during enrolment in this degree.

This thesis does not contain material which has been submitted for the award of any other degree or diploma in my name, in any university or other tertiary institution.

In the future, no part of this thesis will be used in a submission in my name, for any other degree or diploma in any university or other tertiary institution without the prior approval of

The University of Western Australia and where applicable, any partner institution responsible for the joint-award of this degree.

This thesis does not contain any material previously published or written by another person, except where due reference has been made in the text and, where relevant, in the Authorship

Declaration that follows.

This thesis does not violate or infringe any copyright, trademark, patent, or other rights whatsoever of any person.

The following approvals were obtained prior to commencing the relevant work described in this thesis:

- 2015/244021 – Rottnest Island Research Permit, issued by The Rottnest Island

Authority (RIA);

- SF010314 – License to Take Fauna for Scientific Purposes, issued by The

Department of Biodiversity, Conservation and Attractions, formerly The Department of

Parks and Wildlife.

This thesis contains published work and/or work prepared for publication, some of which has been co-authored.

Signature:

Date: 07 August 2020

ii

AUTHORSHIP DECLARATION: CO-AUTHORED PUBLICATIONS

This thesis contains work that has been published and/or prepared for publication, some of which has been co-authored.

Details of the work: Mendes Monteiro, J., Bischoff, K., Gleeson, D. B., Vogwill, R.,

2020 (in prep.). Hydroecological interactions between microbial mats from hypersaline lakes at Rottnest Island and their surrounding aquatic environment. Prepared to be submitted to Hydrology and Earth System Sciences.

Location in thesis: Chapter 2

Student contribution to work: Conducted the literature review, undertook the experimental work, analyzed the majority of the data, and wrote paper with the contribution of editing and reviewing from the co-authors.

Co-author signatures and dates:

Details of the work: Mendes Monteiro, J., Vogwill, R., Bischoff, K., Gleeson, D. B.,

2019. Comparative metagenomics of microbial mats from hypersaline lakes at Rottnest

Island (WA, Australia), advancing our understanding of the effect of mat community and functional genes on microbialite accretion. Limnology and Oceanography, 00:1–17. doi:10.1002/lno.11323

Location in thesis: Chapter 3

Student contribution to work: Conducted the literature review, undertook the experimental work, analyzed the majority of data, and wrote paper with the contribution of editing and reviewing from the co-authors.

Co-author signatures and dates:

iii

AUTHORSHIP DECLARATION: CO-AUTHORED PUBLICATIONS

Details of the work: Mendes Monteiro, J., Vogwill, R., Bischoff, K., Gleeson, D. B.,

2020 (in prep.). Seasonal variability in taxonomic and functional capacity of lithifying and non-lithifying microbial mats from hypersaline lakes at Rottnest Island (Western

Australia). Prepared to be submitted to Frontiers in .

Location in thesis: Chapter 4

Student contribution to work: Conducted the literature review, undertook the experimental work, analyzed the majority of data, and wrote paper with the contribution of editing and reviewing from the co-authors.

Co-author signatures and dates:

Details of the work: Bischoff, K., Wilson, M. E. J., Mendes Monteiro, J., Vogwill, R.,

George, A. D., 2020 (under second review). Sedimentology of coastal carbonate lakes.

Sedimentology, Manuscript ID SED-2019-OM-036.R1.

Location in thesis: Appendices

Student contribution to work: Contributed to fieldwork and sampling, as well as editing and reviewing the paper (10% contribution).

Co-author signatures and dates

Student signature: Date: 07/08/2020

I, Deirdre Gleeson, certify that the student’s statements regarding their contribution to each of the works listed above are correct.

As all co-authors’ signatures could not be obtained, I hereby authorize inclusion of the co- authored work in the thesis.

Coordinating supervisor signature:

Date: 07/08/2020

iv

ACKNOWLEDGMENTS

This PhD research was supported by an Australian Government Research Training Program

(RTP) Scholarship. It was also supported by the C. F. H and E. A. Jenkins Postgraduate

Research Scholarship (stipend scholarship).

This PhD research was funded by The Rottnest Island Authority and logistically supported by

The Rottnest Island Authority staff, Shane Kearney, Cassyanna Thomas, Luke Wheat, and

Shirley Wills.

I would like to acknowledge the Convocation of UWA Graduates for granting me the

Convocation Postgraduate Research Travel Award, which has supported this PhD research.

I would also like to acknowledge the Graduate Research School for supporting this PhD research with the GRS Travel Award.

I would like to acknowledge all technical assistance that was kindly provided by the technician team from The University of Western Australia School of Agriculture and Environment, Hazel

Gaza, Michael Smirk, Chris Brouwer and Louise Heyworth for the support with DNA extraction, inductively coupled plasma-atomic emission spectrometry (ICP-AES) and high pressure ion chromatography (HPIC) that are described in Material and Methods of Chapter 2,

3 and 4 of this thesis.

I would like to thank my research comrade and friend Belinda Martin for kindly supporting me with part of the statistical analysis described in Chapter 3 and 4 of this thesis. Thank you so much for your patience and friendship.

I would like to thank Karl Bischoff not only for being a research collaborator and co-author in some publications that have arisen from this research but also for his genuine friendship and kind support during challenging times.

I would like to thank my supervisor Dr Ryan Vogwill for the academic support and constant encouragement. Thank you for believing on my potential.

v

ACKNOWLEDGEMENTS

I would like to demonstrate my deepest appreciation and gratitude to my Principal and

Coordinating supervisor Dr Deirdre Gleeson, for the amazing support throughout these challenging last four years. Words cannot express how deeply grateful I am for having a supervisor like you.

Lastly, I would like to thank those who have always supported me behind the scenes. I could have not accomplished this milestone without your love and care. Thank you very much my

Soka Gakkai friends, my dearest family (Brazilian and Australian), and my beautiful and beloved husband.

“None of us can exist in isolation. Our lives and

existence are supported by others in seen and unseen

ways, be it by parents, mentors or society at large. To

be aware of these connections, to feel appreciation for

them, and to strive to give something back to society in

a spirit of gratitude is the proper way for human beings

to live.” Daisaku Ikeda.

vi

TABLE OF CONTENTS

THESIS DECLARATION...... ii AUTHORSHIP DECLARATION: CO-AUTHORED PUBLICATIONS ...... iii ACKNOWLEDGEMENTS ...... v ABBREVIATIONS ...... iix FIGURES ...... xi ABSTRACT ...... xvi Chapter 1 ...... 1 General Introduction and Literature Review ...... 1 1.1 General Introduction ...... 1 1.2 Microbial mats ...... 1 1.3 Microbialites ...... 4 1.4 Microbial mats from hypersaline environments ...... 5 1.5 Microbial mats in Rottnest Island ...... 8 1.6 Thesis aim and research objectives ...... 10 Chapter 2 ...... 11 2.1 Introduction ...... 12 2.2 Material and Methods ...... 15 2.2.1 Study area and sampling sites ...... 15 2.2.2 Lake and pore water sampling and analysis ...... 19 2.2.3 Visualization and statistical analysis ...... 20 2.3 Results ...... 20 2.3.1 Surface water level and water balance ...... 20 2.3.2 Lake and pore water analysis...... 24 2.4 Discussion ...... 36 2.5 Conclusion ...... 42 Chapter 3 ...... 53 3.1 Introduction ...... 54 3.2 Material and methods ...... 59 3.2.1 Study area and sampling sites ...... 59 3.2.2 Lake and pore water sampling and analysis ...... 61 3.2.3 Microbial mat sampling, DNA extraction and high throughput sequencing ...... 62 3.2.4 Visualization and statistical analysis ...... 63 3.3 Results ...... 65 3.3.1 Microbial mats description and water chemistry ...... 65

vii

TABLE OF CONTENTS

3.3.2 Microbial mat taxonomic composition ...... 68 3.3.3 Microbial mat functional gene composition ...... 74 3.4 Discussion ...... 78 3.5 Conclusions ...... 84 Chapter 4 ...... 96 4.1 Introduction ...... 97 4.2 Material and Methods ...... 98 4.2.1 Study area and sampling sites ...... 98 4.2.2 Microbial mat sampling, DNA extraction and high throughput sequencing ...... 99 4.2.3 Visualization and statistical analysis ...... 100 4.3 Results ...... 102 4.3.1 Seasonal variation in the taxonomic composition of the microbial mats ...... 102 4.3.2 Seasonal variation in functional gene composition of the microbial mats ...... 113 4.4 Discussion ...... 120 4.5 Conclusion ...... 127 Chapter 5 ...... 134 Discussion and Conclusion ...... 134 5.1 Discussion ...... 134 5.2 Conclusion ...... 140 References ...... 141 Appendices ...... A1

viii

ABREVIATIONS

0 µg/L Microgram(s) per liter EPS Extracellular polymeric

substances

µm Micrometer(s) FDR False discovery rate

µS/cm Microsiemen(s) per centimeter FRP Filterable reactive phosphate

ºC Degree Celsius G Gram(s)

AGRF Australian Genome Research g/L Gram(s) per liter

Facility

ANOVA Analysis of variance Gbp Giga base pair(s)

BFC Bloom filter correction HPIC High pressure ion

chromatography bp Base pair(s) ICP-AES Inductively coupled plasma-

atomic emission spectrometry

CI Confidence interval Kbp Kilo base pair(s) cm Centimeter(s) mAHD Meter(s) above Australian

Height Datum

DNA Deoxyribonucleic acid Mbp Mega base pair(s)

DL Detection limit mg Milligram(s)

DMSP Dimethylsulfoniopropionate mg/L Milligram(s) per liter

DO Dissolved oxygen MG- Metagenomic Rapid

RAST Annotations using Subsystems

Technology

E East mL Milliliter(s)

EC Electric conductivity mm Millimeter(s)

Eh Redox potential mMDS Metric multidimensional scaling

ix

ABREVIATIONS m3/d Cubic meter(s) per day SEM- Electron microscopy and energy

EDS dispersive spectroscopy mS/cm Millisiemen(s) per centimeter SOB Sulfide oxidizing bacteria

OTU Operational taxonomic unit SRB Sulfate reducing bacteria

PCO Principal coordinate analysis STAMP Statistical analysis of

metagenomic profiles

PERMA Permutational multivariate SWL Surface water level(s)

NOVA analysis of variance pH Potential of Hydrogen TCA Tricarboxylic acid

PSU Practical salinity unit TDS Total dissolved solids

QC Quality control TN Total nitrogen

REE Rare earth elements TP Total phosphorus rRNA Ribosomal ribonucleic acid UV Ultraviolet

S South

x

FIGURES

Figure 1-1. Thin section photomicrographs from light microscope of microbialites at Rottnest Island. (A) with thinly laminated internal macrofabric; (B) Thrombolites with clotted internal macrofabric. Images by Karl Bischoff……………………………………………...…………...………5

Figure 2-1. Study area, sampling sites and microbial mats from where lake and pore water samples were collected at the Rottnest Island hypersaline lakes. (A) Map of Australia illustrating the location of Perth. (B) Aerial image of Rottnest Island, located at the coast of Perth. (C) Sampling sites where lake and pore water samples were collected from Lake Baghdad (site BG), Garden Lake (site GN), Herschel Lake (sites HN and HW), Pink Lake (site PL) and Serpentine Lake (sites SN and SW). (D) Blister mat found at sites HW and SN. (E) Flocculent mat found at site BG. (F) Flocculent non-lithifying mat found at site PL. (G) pustular mat found at sites BG, HN, HW, SN and SW. (H) Non-lithifying loosely cohesive mat found at site GN. Aerial image from Google Earth using satellite data collected on January 1, 2018…………………………………………………………………………………………………18

Figure 2-2. Surface water levels (SWL) recorded as meters above Australian Height Datum (mAHD) of Lake Baghdad from August 2014 to May 2018; including minimum and maximum level criteria….22

Figure 2-3. Surface water levels (SWL) recorded as meters above Australian Height Datum (mAHD) of Garden Lake from August 2014 to May 2018; including minimum and maximum level criteria…23

Figure 2-4. Surface water levels (SWL) recorded as meters above Australian Height Datum (mAHD) of Herschel Lake from August 2014 to May 2018; including minimum and maximum level criteria….22

Figure 2-5. Surface water levels (SWL) recorded as meters above Australian Height Datum (mAHD) of Lake Serpentine from August 2014 to May 2018; including minimum and maximum level criteria..23

Figure 2-6. Overall water balance for Garden, Herschel and Serpentine lakes, in addition to other three lakes that were not included in the present research – Government House, Timperley and Pearse lakes – during the period of September 2014 to Spetember 2015. Data from Gillen (2015, unpub.)……………………………………………………………………………..…………..……...24

Figure 2-7. Electrical conductivity (EC) fluctuation within the lake water samples, throughout the sampling period – July and November 2015, and January, February and March 2016. Each line and color represents probe EC measurement for each lake water collected from the seven different sites – BG (Lake Baghdad); GN (Garden Lake); HN and HW (Herschel Lake North and West); PL (Pink Lake); SN and SW (Serpentine Lake North and West)………………………………...………………25

Figure 2-8. Total dissolved solids (TDS) fluctuation within the lake water samples, throughout the sampling period – July and November 2015, and January, February and March 2016. Each line and color represents probe TDS measurement for each lake water collected from the seven different sites – BG (Lake Baghdad); GN (Garden Lake); HN and HW (Herschel Lake North and West); PL (Pink Lake); SN and SW (Serpentine Lake North and West)………………………………………………...26

xi

FIGURES

Figure 2-9. TDS and EC correlation in the lake water samples seasonally collected from the seven different sites – BG (Lake Baghdad); GN (Garden Lake); HN and HW (Herschel Lake North and West); PL (Pink Lake); SN and SW (Serpentine Lake North and West)………………………………………27

Figure 2-10. Piper diagram showing the relative anions and cations relative composition of lake and pore water samples collected from the seven sampling sites in July and November 2015, and January, February and March 2016. For comparison purposes, Perth rainwater data (Crosbie et al., 2007) and seawater data (Pilson, 2013) were included……………………………………………………………28

Figure 2-11. Anions and cations concentrations variation in lake water samples collected from sites: (A) BG – Lake Baghdad; (B) GN – Garden North; (C) HN – Herschel North; (D) HW – Herschel West; (E) SN – Serpentine North; (F) SW – Serpentine West; and (G) PL – Pink Lake. Lake water samples were collected throughout the sampling period – July and November 2015, and January, February and March 2016. - + + - + + Each line and color represents the major anions and cations analyzed – Cl , Na , Ca , HCO3 , K , Mg 2- and SO4 …………………………………………………………………………………..…………...31

Figure 2-12. Anions and cations concentrations variation in pore water samples collected from sites: (A) BG – Lake Baghdad; (B) GN – Garden North; (C) HW – Herschel West; and (D) SN – Serpentine North. Pore water samples were collected throughout the sampling period – July and November 2015, and January, February and March 2016. Each line and color represents the major anions and cations - + + - + + 2- analyzed – Cl , Na , Ca , HCO3 , K , Mg and SO4 …………………………………………………………………………………………………..……..34

Figure 2-13. Mineral saturation index of (A) Garden Lake; (B) Herschel Lake; and (C) Serpentine Lake during the period of September 2014 to Spetember 2015. Data from Gillen (2015, unpub.)…………...33

Figure 3-1. Sampling sites at Rottnest Island hypersaline lakes – Lake Baghdad (site BG), Lake Vincent (site VC), Garden Lake (sites GN and GS), Herschel Lake (sites HW and HE) and Serpentine Lake (sites SN and SW). Aerial image from Google Earth using satellite data collected on January 1, 2018.…………………………………………………………………………………………………...61

Figure 3-2. Microbial mats collected from Rottnest Island hypersaline lakes. (A) Blister mat collected from sites HW and SN. (B) Flocculent mat collected from sites BG and VC. (C) Pustular mat collected from sites BG, HW, SN and SW. (D) Non-lithifying cohesive mat collected from site GS. (E) Non- lithifying loosely cohesive mat collected from sites HE and GN………………………………………67

Figure 3-3. Beta diversity of metagenomes from microbial mat communities in hypersaline lakes at Rottnest Island. Principal coordinate analysis (PCO) based on Bray-Curtis similarity matrices calculated using square root transformation of OTU abundance data from RefSeq annotation (level VII – species). Different colors represent different microbial mat morphologies – blister, flocculent, pustular, non-

xii

FIGURES lithifying cohesive, non-lithifying loosely cohesive; and different shapes represent microbial mat location within the lake – littoral zone and foreshore………………………………………………….70

Figure 3-4. Differentially abundant OTUs (agglomerated by species level) between lithifying and non- lithifying microbial mats at (A) class, (B) family and (C) genus level. Only OTUs with a significance level of 0.005 are shown. A less than zero log2-fold-change indicates phyla that were more differently abundant in lithifying microbial mats, whereas a greater than zero log2-fold-change indicates phyla that were more differently abundant in non-lithifying mats. Each circle represents one species that was enriched in the two types of mats; not a single OTU. Different colors represent different phyla……….72

Figure 3-5. Functional gene composition among metagenomes of microbial mat communities from hypersaline lakes at Rottnest Island, at function level. Principal coordinate analysis (PCO) based on Bray-Curtis similarity matrices calculated using square root transformation of functional gene abundance data from SEED Subsystems annotation (level IV – function). Different colors represent different microbial mat morphologies – blister, flocculent, pustular, non-lithifying cohesive, non- lithifying loosely cohesive; and different shapes represent microbial mat location within the lake – littoral zone and foreshore……………………………………………………………………………..75

Figure 3-6. Differences in relative abundance of functions from SEED Subsystems annotation (level III – third lowest classification in SEED out of four levels). Extended error plots showing pairwise two- sided White’s non-parametric t-tests, with corrected p-values calculated using Benjamini-Hochberg FDR approach (p<0.05); error bars show 95% confidence intervals. Functions are highlighted according to SEED Subsystems pathways (level I): black – carbohydrates; blue – cofactors, vitamins, prosthetic groups, pigments; green – nitrogen metabolism; purple – photosynthesis; red – respiration and; orange – sulfur metabolism……………………………………………………………………………………77

Figure 4-1. Seasonal taxonomic composition from metagenomes of microbial mat communities in hypersaline lakes at Rottnest Island. Stacked bar charts representing the relative abundance of different Classes in blister, pustular and non-lithifying loosely cohesive mat in the different sampling events – July and November 2015, January, February and March 2016. Classes with relative abundance lower than 2% were represented as “Others”………………………………………………………………104

Figure 4-2. Beta diversity of metagenomes from microbial mat communities in hypersaline lakes at Rottnest Island. (A) Principal coordinate analysis (PCO) and (B) metric multidimensional scaling (mMDS) on difference among centroids based on Bray-Curtis similarity matrices calculated using OTU abundance data from RefSeq annotation (level V – family). Different colors represent different microbial mat morphologies – blister, pustular, non-lithifying loosely cohesive; and different shapes represent sampling events – July and November 2015, January, February and March 2016………….106

xiii

FIGURES

Figure 4-3. Differentially abundant OTUs (agglomerated by species level) in blister mat at class level. (A) November 2015 vs January 2016, (B) January 2016 vs February 2016, (C) February 16 vs March 2016. Only OTUs with a significance level of 0.005 are shown. A less than zero log2-fold-change indicates phyla and classes that were more differently abundant in (A) November 2015, (B) January 2016, (C) February 2016; whereas a greater than zero log2-fold-change indicates phyla and classes that were more differently abundant in (A) January 2016, (B) February 2016, (C) March 2016. Each circle represents one species that was enriched in the paired seasons analyses; not a single OTU. Different colors represent different phyla………………………………………………………………………109

Figure 4-4. Differentially abundant OTUs (agglomerated by species level) in pustular mat at class level. (A) July 2015 vs November 2015, (B) November 2015 vs January 2016, (C) January 2016 vs February 16. Only OTUs with a significance level of 0.005 are shown. A less than zero log2-fold-change indicates phyla and classes that were more differently abundant in (A) July 2015 (B) November 2015, (C) January 2016; whereas a greater than zero log2-fold-change indicates phyla and classes that were more differently abundant in (A) November 2015, (B) January 2016, (C) February 2016. Each circle represents one species that was enriched in the paired seasons analyses; not a single OTU. Different colors represent different phyla………………………………………………………………………112

Figure 4-5. Functional gene profiles among metagenomes from microbial mat communities in hypersaline lakes at Rottnest Island. (A) Principal coordinate analysis (PCO) and (B) metric multidimensional scaling (mMDS) on difference among centroids based on Bray-Curtis similarity matrices calculated using functional gene abundance dada from SEED Subsystems annotation (level IV – function). Different colors represent different microbial mat morphologies – blister, pustular, non- lithifying loosely cohesive; and different shapes represent sampling events – July and November 2015, January, February and March 2016…………………………………………………………………..114

Figure 4-6. Mean proportion of significantly different (± 1 SD; p-value <0.10) functions from SEED Subsystems annotation (level III – third classification in SEED out of four levels) in blister mats. (A) November 2015 vs January 2016, (B) January 2016 vs February 2016. Only level III functions from the following level I key pathways are shown: black – carbohydrates; blue – cofactors, vitamins, prosthetic groups, pigments; green – nitrogen metabolism; purple – photosynthesis; red – respiration and; orange – sulfur metabolism…………………………………………………………………………………..117

Figure 4-7. Mean proportion of significantly different (± 1 SD; p-value <0.10) functions from SEED Subsystems annotation (level III – third classification in SEED out of four levels) in pustular mats. (A) July 2015 vs November 2015, (B) November 2015 vs January 2016, (B) January 2016 vs February 2016. Only level III functions from the following level I key pathways are shown: black –

xiv

FIGURES carbohydrates; blue – cofactors, vitamins, prosthetic groups, pigments; green – nitrogen metabolism; purple – photosynthesis; red – respiration and; orange – sulfur metabolism………………………….118

Figure 4-8. Mean proportion of significantly different (± 1 SD; p-value <0.10) functions from SEED Subsystems annotation (level III – third classification in SEED out of four levels) in non-lithifying loosely cohesive mats – January 2016 vs February 2016. Only level III functions from the following level I key pathways are shown: black – carbohydrates; blue – cofactors, vitamins, prosthetic groups, pigments; green – nitrogen metabolism; purple – photosynthesis; red – respiration and; orange – sulfur metabolism…………………………………………………………………………………………...119

xv

ABSTRACT

Microbial mats are organosedimentary structures organized as multilayered carpets of microbial communities growing in aquatic environments. Within the microbial mat microenvironment, water chemistry influences the occurrence of different biofunctional groups of microorganisms, which via their metabolic activities can lead to carbonate precipitation and potentially, microbialite formation. This process in lithifying microbial mats is to date poorly understood, but studies suggest that the taxonomic composition and functional genes in lithifying mats contribute. In contrast to lithifying mats, non-lithifying mats, which do not form microbialites, only sporadically trap carbonate sand grains that are bound to the microbial mat’s extracellular polymeric substances and do not accrete. Both mat types occur in hypersaline lakes at Rottnest Island (WA). Characterizing the microbial communities and functional genes of both mat types, coupled with water chemistry analysis, has allowed a better understanding of these important ecosystems, hence improving their management in addition to elucidating microbialite formation processes. Metagenomics was used to compare taxonomic and functional diversity of both mat types and observe response to seasonal variation in water chemistry. This has improved our understanding of the environmental limiting factors for microbialites to form and whether distinctions in the microbial mat community structure and their functional gene composition influence microbialites accretion. Results revealed that both mat types harbor taxa and functional genes that are known to play important roles in microbialite formation but seasonal variation in key metabolic activities occurred only in lithifying mats. Non-lithifying mats are potentially impacted due to decreasing water salinity from external processes.

xvi

Chapter 1

Chapter 1

General Introduction and Literature Review

1.1 General Introduction

This thesis investigates the hydroecology of microbial mat communities in hypersaline lakes at Rottnest Island. The main aims of this study were to determine the environmental limits that enable microbial mats to develop and form microbialites and whether distinctions in the microbial mat community structure and their functional gene composition – due to differences in mat morphology or seasonality – may influence their capacity to form microbialites. For context, this chapter provides a review of literature relating to microbial mats and their potential to form microbialites, with particular reference to those found in hypersaline lakes at Rottnest

Island.

1.2 Microbial mats

Microorganisms are commonly found in nature forming communities adhered to a solid surface as complex ecological assemblages in different types of ecosystems across the planet

(Davey and O’Toole, 2000; Neelakanta and Sultana, 2013). For millions of years, microorganisms have used surface adhesion to survive and evolve as biological organizations, allowing the entire community to cope with different surrounding abiotic factors, including stressful ones, such as extreme temperatures, ultra violet (UV) radiation, low nutrient availability, and osmotic stress (Bonilla-Rosso et al., 2012; Inskeep et al., 2013; Tytgat et al.,

2014; Babilonia et al., 2018). Through this adhering mechanism, microorganisms can form a range of different structures, from simple monospecific biofilms to highly diverse microbial

1

Chapter 1 mats, in which complex ecological networks are observed (Dupraz et al., 2009; Bonilla-Rosso et al., 2012).

Microbial mats are vertically multilayered microbial ecosystems that develop in environments with liquid-solid interface and are self-sustaining (Visscher and Stolz, 2005;

Prieto-Bajaras et al., 2018). In addition, microbial mats are composed of a complex assemblage of different species of microorganisms, which are embedded in a self-producing matrix of extracellular polymeric substances (EPS), wherein signal exchange among different microbial niches enables a greater network of resources and energy, increasing the survival of the entire community (Jungblut et al., 2016; Ruvindy et al., 2016; Gutiérrez-Preciado et al., 2018). It is theorized that microbial mats have played a crucial role in global biogeochemical cycling and sedimentation (Dupraz and Visscher, 2005; Peimbert et al., 2012). The complex network of different metabolic pathways present in microbial mats have contributed to the formation of the present atmosphere, by producing, for example, hydrogen, methane and oxygen, and paving the way for oxygen dependent life (Hoehler et al., 2001; Tice and Lowe 2004).

The microorganisms that usually comprise microbial mats are bacteria, and in smaller abundances archaea, eukaryotes and viruses (Prieto-Barajas et al., 2018; Suosaari et al., 2018;

Wong et al., 2018). Among them, the most common biofunctional groups are: (i) the oxygenic photosynthetic bacteria (i.e., cyanobacteria), which fix CO2 and sometimes N2, via photosynthesis using light energy; (ii) anoxygenic photosynthetic bacteria (i.e., non-sulfur green bacteria), that mainly use HS- as electron donor for photosynthesis, and to a lesser extent are able to fix N2; (iii) aerobic heterotrophic bacteria which obtain energy from O2 respitarion and organic carbon; (iv) fermentative bacteria, which utilize as electron donor and acceptor organic carbon or sulfur compunds; (v) anaerobic heterotrophs, mainly sulfate reducing

2- - bacteria (SRB), that respire organic carbon with SO4 while producing HS ; and (vi) sulfide oxidizing bacteria (SOB), many of which are chemolithoautotrophs that oxidize reduced sulfur 2

Chapter 1 compounds with O2 or nitrate while fixing CO2 (Dupraz and Visscher, 2005; Dupraz et al.,

2009; Prieto-Barajas, et al., 2018). All these different biofunctional groups of microorganisms interact as a consortium, wherein biogeochemical processes occur and the products of the metabolism of each group is available as energy source or electron donor/acceptor for the other biofunctional groups (Baumgartner et al., 2006; Cole et al., 2014; Bolhuis et al., 2014; Llorens-

Marès et al., 2015).

Photosynthetic bacteria are the primary producers in microbial mats, and the energy produced during photosynthesis is the main source of energy and the first step for developing a trophic network within the mat. During this process, cyanobacteria, for example, utilize light energy to fix inorganic carbon (CO2) into organic carbon ([CH2O]2), releasing oxygen into the mat matrix (Cole et al., 2014; Prieto-Barajas, et al., 2018). The combination of different metabolisms present in microbial mats have the ability to alter the balance between more reduced (i.e. organic matter) or more oxidized (i.e. CO2) forms of carbon (Dupraz et al., 2009).

When CO2 is present in water as bicarbonate or carbonate ions, due to increased pH and carbonate alkalinity, these ions can combine with cations (i.e., Ca2+, Mg2+) to precipitate carbonate minerals (Glunk et al., 2011; Cabestrero et al., 2018; Sanz-Montero et al., 2019). In general terms, carbonate precipitation depends on the saturation index (SI), which is a function of the solubility product constant Ksp and the ion activity product (IAP), calculated as SI=log

(IAP/Ksp). The IAP is proportional to the ion concentration through an activity coefficient.

When IAP > Ksp, the solution is said to be supersaturated and carbonate precipitation can occur

(Visscher and Stolz, 2005). Laboratory experiments have demonstrated that if SI ≥ 0.8 carbonate may precipitate, depending on the local pH and associated alkalinity (Visscher and

Stolz, 2005; Baumgartner et al., 2006). In circumstances where Ca2+ is sufficiently present and microbial metabolisms induce the alkalinity to increase, the equilibrium is moved toward calcium carbonate precipitation. However, if the active metabolisms lead to alkalinity 3

Chapter 1 decreasing, the opposite happens and carbonate is dissolved (Jiang et al., 2013; Zhu and

Dittrich, 2016). Within microbial mats when the balance between precipitation and dissolution is positive it induces the formation of structures collectively known as microbialites (Burne and Moore, 1987; Riding, 2011).

1.3 Microbialites

Microbialites are organosedimentary deposits formed due to trapping and binding of detrital sediment, or precipitation of carbonate minerals, due to the metabolic activities of resident microorganisms in microbial mats (Burne and Moore 1987; McNamara and Awramik

1994). The classification of different microbialites is determined by their dominant internal macrofabric and according to Riding (2011), they can be classified as stromatolites, thrombolites, dendrolites and leiolites. Stromatolites have an internal macrofabric that is laminated and thinly layered while thrombolites are internally organized with irregular equidimensional clots that can elongate into branches (Figure 1-1). Dendrolites have an internal macrofabric that is dendritic and shrub-like; whereas leiolites have an internal macrofabric that is aphanitic, lacking any observable features at the mesoscale (Riding et al.,

2011). To date, it is not clear what determines the type of macrofabric in each different category of microbialites, though it is believed that the diversification of biofunctional groups within the microbial mats can influence the manner in which the macrofabric is accreted (White et al.,

2016; Suosaari et al., 2016).

4

Chapter 1

Figure 1-1. Thin section photomicrographs from light microscope of microbialites at Rottnest Island. (A)

Stromatolites with thinly laminated internal macrofabric; (B) Thrombolites with clotted internal macrofabric.

Images by Karl Bischoff.

Each occurrence of microbialites and their microbial mats are a distinct community in terms of history, structure, and morphology. Minerals within accreted microbialites can provide information around ancient water chemistries, climates and other environmental parameters

(van Kranendonk 2006). Consequently, microbialites and microbial mats are a significant resource for studying the origin, evolution, and distribution of life, particularly the physiological processes that shape planetary biology (Chilton et al., 2012).

1.4 Microbial mats from hypersaline environments

Microbial mats and microbialites have been present on Earth since the Archaean era, approximately 3.7 billion years ago, and are considered the oldest form of life on the planet

(Awramik, 1976; Green and Jahnke, 2010; Nutman et al., 2016). During Proteozoic era (2.5-

0.57 billion years ago), microbial mats and microbialites were abundant and distributed worldwide (Green and Jahnke, 2010). By the late Proteozoic and early Paleozoic however, the environmental dominance of microbial mats and microbialites sharply declined and it is believed that metazoan grazing and burrowing, nutrients and substrate competition, and changes in the seawater chemistry composition were the main causes (Grotzinger, 1990; 5

Chapter 1

Riding, 1997, Wood et al., 2000). In modern environments these microbial ecosystems are restricted to extreme environments, such as geothermal springs (Nishida et al., 2018; Sharma et al., 2018), oligotrophic environments (Pajares et al., 2015; Ponce-Soto et al., 2015), and hypersaline environments (Farías et al., 2017; Wong et al., 2017). Such environments have the perfect conditions to inhibit or limit the growth of competitor organisms, for example, grazing eukaryotic organisms (Riding, 2006; Dupraz et al., 2011; Revsbech et al., 2016). The best studied microbial mats are those from hypersaline environments (Prieto-Barajas et al., 2018).

These mats are able to survive in harsh conditions, such as extremely high salinity and/or high levels of radiation (Ruvindy et al., 2016; Wong et al., 2015).

Despite the conditions, hypersaline microbial mats harbor a diverse microbial community that comprises mainly bacteria and archaea (Prieto-Barajas et al., 2018; Suosaari et al., 2018). Within these mats, there is a formation of spatiotemporal chemical gradients that determine the vertical distribution of the major bacterial groups, which is related to the presence of light, oxygen and H2S (Golubic, 1976; Kunin et al., 2008, Wong et al., 2015). In Guerrero

Negro, located in Baja California Sur (Mexico), for example the upper layers of the hypersaline mats, comprising the photic-oxic zone (0-3.5 mm), are predominantly composed of phototrophic Cyanobacteria such as Oscillatoriales, Nostocales, and Chroococcales (Harris et al., 2013; Wong et al., 2015; Prieto-Barajas et al., 2018). Within the lower layers, different taxa have been described in the Guerrero Negro mats, for example Bacteroidetes, Proteobacteria,

Planctomycetes, Spirochaetes, Verrucomicrobia, and Chloroflexi; the latter being the most abundant taxon. Chloroflexi are able to perform anoxygenic photoautotrophic processes oxidizing H2S, which can then promote carbonate precipitation (Visscher and Stolz, 2005; Lay et al., 2006; Harris et al., 2013). Guerrero Negro hypersaline mats also harbored some archaeal groups, specifically Crenarchaeota, which play an important role in sulfur cycling within the mats (Wong et al., 2016; Prieto-Barajas et al., 2018). 6

Chapter 1

Another example of a well-studied microbial mat ecosystems is located in Hamelin

Pool in Shark Bay (Western Australia), which comprises one of the most extensive microbial mat systems worldwide (Jahnert et al., 2012; Wong et al., 2018). Hamelin pool microbial mats are subject to high UV radiation and salinity over 60 PSU (Wong et al., 2015; Ruvindy et al.,

2016), in addition to daily tidal cycles, which cause salinity and temperature fluctuations

(Wong et al., 2017). Similar to Guerrero Negro, Hamelin Pool microbial mats have a phototrophic consortium at the surface which represent the main source of energy to lower anoxic layers (Wong et al., 2015; Wong et al., 2017). However, the archaeal communities from these two ecosystems are different; instead of Crenarchaeota, Shark Bay hypersaline mats are dominated by Euryarchaeota, which comprises different methanogenic taxa (Robertson et al.,

2009; Wong et al., 2016). SRB from the Deltaproteobacteria have been also described as part of the Hamelin Pool microbial mats. Both SRB and methanogens appear to cooperate with each other and other groups (i.e., Cyanobacteria) in a spatially-especialized microenvironment organized around chemical gradients (Decho et al., 2010; Gallagher et al., 2012).

Both Guerrero Negro and Shark Bay hypersaline microbial mats are well-studied environments and the characterization of their microbial communities have been contributing to expand scientific knowledge regarding the complexity involved in the processes of microbial mat development and microbialite formation (Lay et al., 2016; Wong et al., 2016; Ruvindy et al., 2018). However, there is still not an explicit understanding of how the communities actually form microbially mediated calcium carbonate, and the study of different microbial mat ecosystems could advance our general knowledge in this area. In Rottnest Island, different morphologies of hypersaline microbial mats are developing but they have not been characterized yet and are still poorly understood.

7

Chapter 1

1.5 Microbial mats in Rottnest Island

Rottnest Island is classified as A-Class Reserve under the Land Administration Act

1997, which provides the island with highest level of protection in Western Australia and restrict and control all activities with potential to disturb the island’s environment (Rottnest

Island Authority, 2014). The Island is a carbonate island located in the Indian Ocean 18 kilometers west of Fremantle (Western Australia) (Brooke, 2001). The Island is 11 kilometers long and 4.5 kilometers at its widest point and the waters surrounding the island includes 3,800 hectares of Marine Reserve (Bryan et al., 2016; Rottnest Island Authority, 2014). The Island bedrock is composed of carbonate eolianite (Tamala Limestone) deposited during the

Pleistocene to mid-Holocene, which is weakly cemented and is typified by dual porosity consisting of numerous connected channels within a matrix of interparticle porosity (Playford and Leech, 1977; Smith et al., 2012).

Inland, Rottnest has a system of hypersaline lakes which cover 10% of the Island at more than 200 hectares (Rottnest Island Authority, 2014). It is the only island among over 200 islands in Western Australia which has hypersaline lake habitats of this size (Rottnest Island

Authority, 2014). It is thought that the hypersaline lakes were formed as a result of repeating cycles of carbonate precipitation and dissolution controlled by sea level fluctuations (Mylroie et al., 1995). The hypersaline water quality of the lakes is believed to be a result of evapoconcentration, allowing hypersaline conditions to develop once the ocean connectivity was initially restricted (Bryan et al., 2016).

The hypersaline lakes at Rottnest Island are of local, national and international significance, providing an important habitat for diverse fauna, for example, frogs, quokkas and migratory birds. The native flora at Rottnest is well adapted to the predominantly nutrient poor soils, in conjunction with the salty and often windy conditions. The Samphire communities around the Rottnest hypersaline lakes have been listed as ‘Vulnerable Ecological Communities’ 8

Chapter 1 under Commonwealth legislation (Rottnest Island Authority, 2014). A number of the hypersaline lakes also support significant microbial mats and microbialite communities

(Mendes Monteiro et al., 2019). These hypersaline lakes, in which microbial mats are actively developing, form the central basis for this thesis and include Lake Baghdad, Garden Lake,

Herschel Lake, Serpentine Lake, Pink Lake and Lake Vincent. Studies suggest that the microbial mats and microbialites developing on the hypersaline lakes substrate restrict interaction with underlying groundwater, which helps to maintain the hypersalinity of the lakes

(Playford, 1997). The microbialite communities on Rottnest contain fossilized structures as well as active the microbial mats that are actively forming modern microbialites (Appendix-

II; K. Bischoff, co-authored by J. Mendes-Monteiro, currently in revision with the Journal

Sedimentology). The form and diversity of both microbial mats and microbialite structures

(both modern and historical) in the Rottnest lakes is unique in south-west Western Australia

(Rottnest Island Authority, 2014). The ecological water requirements of the microbialite communities on Rottnest are currently unknown; however previous research has shown that increased acidity, nutrients or salinity changes can all impact these communities (John et al.,

2009).

The microbial mats in the hypersaline lakes at Rottnest Island have different morphologies which are poorly characterized and little is known about their microbial ecology.

Anthropogenic and development activities occurring at Rottnest Island have been affecting the ecology of the hypersaline lakes, representing a threat to the microbial mats (John et al., 2009).

These threats in combination with the lack of information about these communities and their presence in a restricted and protected area, have qualified them as Priority Ecological

Communities – Priority 1 (Rottnest Island Authority, 2014). Rottnest Island Authority, the government body that manages the island, have been developing strategies to conserve the hypersaline lakes and their microbial mat ecosystems. The present research is part of the 9

Chapter 1

Terrestrial Conservation Action Plan designed by the Rottnest Island Authority to preserve the different ecosystems in the island, including the hypersaline lakes and the microbial mats.

1.6 Thesis aim and research objectives

Considering the global and scientific importance of microbial mats and microbialites, in addition to the environmental pressure that microbial mats at Rottnest Island are undergoing, this research aimed to: unravel the process in with microbial mats and microbialites are formed to enhance scientific knowledge about carbonate accretion; and to better inform the Rottnest

Island Authority to support their work managing these ecosystems.

To achieve this aim, the research objectives were: a) study the hydroecological interactions between the microbial mats from the hypersaline lakes at Rottnest Island by analyzing water quality parameters in different times throughout 2015/2016; b) analyze the microbial community taxonomy and functional genes composition of different microbial mats from the hypersaline lakes at Rottnest Island, examining variations between lithifying and non- lithifying mats and the implications of seasonal environmental changes in the microbial mat communities.

It was hypothesized that (1) due to environmental factors (such as eutrophication) and seasonal changes (from winter to summer) would occur, causing variation in lake water hydrochemical parameters (for example, salinity, ions concentration, nutrient concentration and alkalinity); (2) this variation would affect both lithifying and non-lithifying mats by causing seasonal shifts in their microbial community composition and functional capacity; (3) differences in the two types of mat community structure (lithifying and non-lithifying) and their functional capacity would influence their ability to form microbialites.

10

Chapter 2

Chapter 2

Hydroecological interactions between microbial mats from hypersaline lakes at Rottnest

Island and their surrounding aquatic environment

This chapter has been prepared for publication and will be submitted to Hydrology and Earth

System Sciences:

Mendes Monteiro, J., Bischoff, K., Gleeson, D. B., Vogwill, R., 2020 (in prep.).

Hydroecological interactions between microbial mats from hypersaline lakes at Rottnest Island and their surrounding aquatic environment. Prepared to be submitted to Hydrology and Earth

System Sciences.

11

Chapter 2

2.1 Introduction

Microbial mats are communities of microorganisms that colonize a solid surface and form vertically laminated organosedimentary structures (van Gemerden, 1993; Franks and

Stolz, 2009). The microbial communities within microbial mats are usually associated with the sediement/water interface in a particular environment and the community is developed as a result of microbial growth and activity, sediment trapping and binding in the organic matrix, and sedimentation. (Margulis et al., 1980). Microbial mat communites present high complexity, containing diverse population of Bacteria, Archaea and Eukaryotes (Stolz, 2000). Microbial mats usually thrive in extreme environments where grazing organisms are limited, for example, in extreme pH conditions (e.g., acidic sulfur caves and iron springs), high temperature (e.g., thermal springs and deep sea vents), hypersalinity (e.g., hypersaline lakes) or in environments that support chemolithotrophysm (e.g., methane seeps) (Franks and Stolz, 2009). Ancient microbial mats have shaped planetary biology by changing redox conditions via a variety of metabolic processes including oxygenic photosynthesis, carbon cycling and nitrogen fixing

(Dupraz et al., 2009; Olson et al., 2016; Prieto-Barajas et al., 2018; Thomazo et al., 2018); and modern microbial mats have been studied in evolution and sciences as they represent the modern analogs to ancient life and possibly extraterrestrial ecosystems (Des

Marais, 2003).

Complex microbial mats are usually composed by a variety of microorganism from different functional groups, which due to their specific and differentiated metabolisms (van

Gemerden, 1993), are able to change the local environment’s pH conditions and induce the formation of carbonate minerals (Baumgartner et al., 2006; Dupraz et al., 2009; Prieto-Barajas et al., 2018); which will in turn affect the local environment. Oxygenic phototrophic microorganisms can promote carbonate precipitation by increasing the availability of carbonate ions due to carbon dioxide uptake during photosynthesis (Glunk et al., 2011; Bundeleva et al., 12

Chapter 2

2012; Jiang et al., 2013). Sulfate reducing bacteria can also induce carbonate precipitation by increasing local pH and alkalinity during sulfate reduction and organic carbon oxidation

(Gallagher et al., 2012; Pace et al., 2016; Zhu and Dittrich, 2016). Both processes release carbonate ions into the microbial mat microenvironment, which if coupled with the presence of bivalent cations (e.g., Ca2+, Mg2+), can lead to carbonate mineral precipitation (Dupraz et al., 2009). During the process of extracellular polymeric substances (EPS) formation, Ca2+ and

Mg2+ cations bind to the EPS matrix, inhibiting the process of carbonate mineral precipitation.

However, the presence of heterotrophic microorganisms within the microbial mat allows these cations to be released from the EPS matrix via EPS degradation (Glunk et al., 2011; Cabestrero et al., 2018; Sanz-Montero et al., 2019). This process of microbially-induced mineralization results in the formation of organosedimentary structures known as microbialites (Burne and

Moore, 1987; Riding, 2011). Both microbial mat development and microbialite formation processes can alter local water chemistry, for example by changing local pH and oxygen (O2), sulfide (S2-) and calcium (Ca2+) concentrations (Baumgartner et al., 2009).

Modern microbial mats are considered important model systems that hold a key to understanding the ancient environment and atmosphere, as they could provide insight into the role of microorganisms in biogeochemical processes (e.g., carbon, nitrogen and sulfur cycling), and microbe–mineral interactions (e.g., precipitation of carbonates). Their occurrence is typically restricted to extreme environments, such as hypersaline lakes (Farías et al., 2017;

Mlewski et al., 2018; Shen et al., 2018), hypersaline marine basins (Pagès et al., 2015; Ruvindy et al., 2016; Wong et al., 2017), geothermal springs (Pepe-Ranney et al., 2012; Nishida et al.,

2018; Sharma et al., 2018) and oligotrophic freshwater environments (Bonilla-Rosso et al.,

2012; Pajares et al., 2015; Ponce-Soto et al., 2015); in which competition and grazing pressure are limited, allowing the mats to grow (Riding, 2006; Dupraz et al., 2011). The main factors that support microbial mat development in extreme environments include favorable water 13

Chapter 2 chemistry, including low nutrients; therefore water chemistry is not only important for mineral nucleation but also for the growth of the microbial mats (Baumgartner et al., 2009; Chiriac et al., 2017; Prieto-Barajas et al., 2018).

In Western Australia, modern microbial mats have been found in several hypersaline environments, such as Hamelin Pool (Shark Bay) (Wong et al., 2018; D’Agostino et al., 2019),

Lake Clifton (Gleeson et al., 2016; Warden et al., 2016), Lake Thetis (Grey and Planavsky,

2009; Wacey et al., 2018) and hypersaline lakes at Rottnest Island (John et al., 2009; Mendes

Monteiro et al., 2019). The pH of the water present in these hypersaline environments is usually slightly alkaline (7.5 to 8.0) with the salinity ranging from 70 to 140 mg/L (Suosaari et al.,

2016; Warden et al., 2016; Suosaari et al., 2018).

Previous studies have demonstrated that water quality plays an important role in controlling the microbial mat communities. For example, at Lake Clifton the effects of salinity variation on microbialites/microbial mat communities suggest that due to a considerable increase in salinity (from 18.0 g/L to 47.9 g/L) between 1985 and 2006 the dominant microbial communities within the thrombolites was altered. In earlier studies, the microbialites were dominated by Scytonema sp. and other filamentous cyanobacteria (Smith et al., 2010). In 2006, however, the dominant communities were diatoms and unicellular cyanobacteria (Smith et al.,

2010). Gleeson et al. (2016) also reported increased salinity and nutrient levels at Lake Clifton with Proteobacteria, Bacteroidetes and Firmicutes now being described as the dominant groups of microorganisms. These reports suggest that the water quality variation may be contributing to changes in the dominant microbial populations within those microbialites/microbial mats.

At Rottnest Island, John et al. (2009) compared the microbial communities within mats from four different hypersaline lakes (Garden, Government, Serpentine and Herschel) in association with nutrients levels (total phosphorus and total nitrogen). The results showed that Garden

Lake, which had the highest levels of nutrients (for example 320 µg/L of total phosphorus in 14

Chapter 2 summer, compared to 38 µg/L in Serpentine Lake), also contained the least cohesive microbial mats and a microbial assemblage of cyanobacteria and diatoms that differed from the other lakes, indicating that eutrophication has had a direct adverse impact on the microbial mats

(John et al., 2009).

Current land and water-use activities at Rottnest Island represent environmental pressure to the microbial mats and because of this the microbial mats at Rottnest Island have been categorized as ‘Priority Ecological Communities’ under Western Australian Legislation

– which covers naturally occurring and rare communities with possible threatened status. The

Rottnest Island Authority, who manage and control the island on behalf of the government of

Western Australia, have been developing strategies to protect the microbial mats (Rottnest

Island Authority, 2014). Considering the historical, geological and ecological importance of microbial mats, it is important to assess and monitor the water quality of the hypersaline lakes at Rottnest Island, which will assist on developing an understanding of the influence of water quality changes on microbial mats and microbialite formation, and vice versa. The present study investigated differences in water physicochemical parameters from five different hypersaline lakes at Rottnest Island to determine the current environmental conditions in which the microbial mats are growing and to identify possible threats and interrelationships between microbial assemblage and water chemistry. This assessment will not only support Rottnest

Island Authority to better manage this unique ecosystem, but also expand our knowledge of microbially-induced mineralization, microbialite formation and the relationship between these communities and environmental water chemistry.

2.2 Material and Methods

2.2.1 Study area and sampling sites

A survey was conducted in April 2015 to identify the dominant microbial mats present in the hypersaline lakes at Rottnest Island, located off the coast of Perth (Western Australia; 15

Chapter 2

32°00'7.20" S, 115°31'1.20" E) (Figure 2-1). Five microbial mat morphologies – blister, flocculent, flocculent non-lithifying, pustular and non-lithifying loosely cohesive (Figure 2-1)

– were identified at different locations in five hypersaline lakes – Baghdad, Garden, Herschel,

Serpentine and Pink. Blister, flocculent and pustular mats were defined as lithifying mats growing in association to microbialites structures; whereas flocculent non-lithifying and non- lithifying loosely cohesive mats were defined as sediment associated mats with no apparent lithifying activity. A parallel study conducted in conjunction with the current work and, utilizing microscopy and mineralogical analysis, identified that the blister mat is a laminated mat growing in association with discontinuous layered stromatolites, whereas the flocculent mat grows associated with continuous layered stromatolites. It also reported the pustular mat as a thrombolytic mat developing closely to mottled stromatolites (Appendix-II; K. Bischoff, co-authored by J. Mendes-Monteiro in revision with Sedimentology). The blister mat was observed growing in the littoral zone (depth = 1.0 to ~50.0 cm) of Herschel and Serpentine

Lakes, whereas the flocculent mat was identified in the littoral zone (depth = 1.0 to ~50.0 cm) of Lake Baghdad and the flocculent non-lithifying mat was observed in the littoral zone (depth

= 1.0 to ~25.0 cm) of Pink Lake. The pustular mat was located on the foreshore zone (depth =

0.0 cm) of Baghdad, Herschel and Serpentine Lakes, while the non-lithifying loosely cohesive mat was observed on the foreshore zone of Garden Lake. Both mats, pustular and non-lithifying loosely cohesive, seemed to preferably grow in areas of heavy fresh groundwater discharge, in particular the pustular mat. To better understand the links between water chemistry and specific microbial mats lake water was collected from seven sites in the hypersaline lakes – site

Baghdad (BG) at Lake Baghdad, site Garden North (GN) at Garden Lake, sites Herschel North and West (HN and HW) at Herschel Lake, sites Serpentine North (SN) and West (SW) at

Serpentine Lake, and site Pink Lake (PL) at Pink Lake (Figure 2-1; Supplemental Table 2-

1). Due to the presence of the pustular and non-lithifying loosely cohesive microbial mats 16

Chapter 2 growing on the lake foreshore (depth = 0.0 cm), pore water samples were also collected from

Lake Baghdad (site BG), Garden Lake (site GN), Herschel Lake (site HW) and Serpentine

Lake (site SN) (Figure 2-1; Supplemental Table 2-1). Seven lake water samples and four pore water samples were seasonally collected in July (winter) and November (early-summer) 2015, and in January (mid-summer), February (mid-summer) and March (end-summer) 2016; with the exception of Pink Lake, which had lake water samples collected only at the first four sampling events, as during the last sampling event it was dry. In total, 34 seasonal lake water samples were collected from BG (n=5), GN (n=5), HN (n=5), HW (n=5), PL (n=4), SN (n=5) and SW (n=5); and 20 pore water samples were collected from BG (n=5), GN (n=5), HW (n=5),

SN (n=5) throughout the sampling period (Supplemental Table 2-1).

17

Chapter 2

Figure 2-1. Study area, sampling sites and microbial mats from where lake and pore water samples were collected at the Rottnest Island hypersaline lakes. (A) Map of Australia illustrating the location of Perth. (B) Aerial image of Rottnest Island, located at the coast of Perth. (C) Sampling sites where lake and pore water samples were collected from Lake Baghdad (site BG), Garden Lake (site GN), Herschel Lake (sites HN and HW), Pink Lake

(site PL) and Serpentine Lake (sites SN and SW). (D) Blister mat found at sites HW and SN. (E) Flocculent mat found at site BG. (F) Flocculent non-lithifying mat found at site PL. (G) pustular mat found at sites BG, HN, HW,

SN and SW. (H) Non-lithifying loosely cohesive mat found at site GN. Aerial image from Google Earth using satellite data collected on January 1, 2018.

Throughout the survey and sampling period, it was observed that when the lake surface water level rose that the blister mat followed the raised surface water level and grew over the pustular mat in both Herschel and Serpentine Lakes. Therefore, surface water level data from

18

Chapter 2 the sampled lakes was obtained (reported as meters above Australian Height Datum; mAHD), from Rottnest Island Authority gauges, to monitor the distribution of the microbial mats, particularly pustular and non-lithifying loosely cohesive mats, which were observed to require fresh groundwater input.

2.2.2 Lake and pore water sampling and analysis

For lake water sampling, five aliquots were collected, stored in sterile 50 mL centrifuge tubes and each aliquot was used for (1) total nutrients; (2) dissolved nutrients; (3) total alkalinity; (4) anions and cations; and (5) metals, metalloids and rare earth elements (REE) analysis, respectively. For total nutrients analysis, a 40 mL aliquot of lake water was collected using sterile syringes and directly transferred to centrifuge tubes without being filtered. For each of dissolved nutrients and total alkalinity a 40 mL aliquot of lake water was collected and filtered using 0.45 µm pore size membrane filter cartridges attached to the syringe and transferred to centrifuge tubes. For anions and cations analysis, a fourth aliquot of 55 mL was collected and filtered as described previously. Finally, for metal, metalloids and REE 45 mL of water was collected and filtered as described previously and 5 mL of HNO3 20% was added.

New filter cartridges were used for each sampling site and lake water samples were stored at -

18 °C until analysis.

For pore water sampling, microbial mat samples were collected, stored in sterile 50 mL centrifuge tubes at -18 °C, and later (within 24 to 48 hours after sampling) centrifuged at 3,000 rpm for 2 hours. The supernatant fluid was considered pore water that was then sequentially double filtered through 8.0 µm and 0.45 µm pore size membranes. The filtered pore samples were then stored at -18 °C. Samples were then analyzed for total nutrients; dissolved nutrients; total alkalinity; anions and cations; and metals, metalloids and REE.

Flow injection analysis, inductively coupled plasma-atomic emission spectrometry

(ICP-AES) and high pressure ion chromatography (HPIC) were used for total nutrient; 19

Chapter 2 dissolved nutrients; anions and cations; and metals metalloids and REE analysis. Alkalinity

Test Kit, Model AL-DT (HACH®), was used to measure alkalinity in the field and the probe

HI 9829 Water Quality Meter (Hanna Instruments®) was used to measure water depth, pH, temperature, dissolved oxygen (DO), redox potential (Eh), electric conductivity (EC) and total dissolved solids (TDS). Due to the methodology used to collect pore water, it was not possible to use the probe to measure the cited water parameters, as only small volumes were produced during centrifuging, which was all required for detailed analysis.

2.2.3 Visualization and statistical analysis

Lake and pore water environmental data were analyzed using RStudio (version

1.2.1335). Differences among lake and pore water data from the different lakes throughout the sampling period were visualized as line graphs created in RStudio (version 1.2.1335) using ggplot2 package (version 3.1.1). The script for all water data analysis is available at https://github.com/Microbial-Ecology/WaterData. The relative anion composition of lake and pore water samples were visualized as a piper diagram that was created using GW_Chart

(version 1.29.0.0) (Winston, 2000).

2.3 Results

2.3.1 Surface water level and water balance

Surface water level measurements, recorded as meters above Australian Height Datum

(mAHD), from Baghdad, Garden, Herschel and Serpentine Lakes were obtained from Rottnest

Island Authority, during the period of August 2014 to May 2018. Surface water level data from

Pink Lake were not available. Minimum (-1.00 mAHD) and maximum (0.00 mAHD) water level criteria were set based on the distribution of the microbial mats occurring on the foreshore zone of the lakes, specifically the pustular and non-lithifying loosely cohesive mats. For all lakes assessed, the surface water level was generally within the minimum and maximum level criteria, with a few exceptions (Figures 2-2 to 2-5). In August 2017, Lake Baghdad surface 20

Chapter 2 water level was registered below the minimum level criterion, at -1.73 mAHD (Figure 2-2).

This result is likely to be a data error, as in July 2017 – the previous month – the surface water level in Lake Baghdad was -0.20 mAHD and in September 2017, it was -0.16 mAHD. The registration of a surface water level of -1.73 mAHD would have meant a drop of 1.53 mAHD from July to August 2017, then rising back another 1.57 mAHD in September 2017. This is a highly unlikely scenario and thus this data point is to be treated as a recording error. For Garden

Lake, in November 2017, the surface water level registered above the maximum level criterion, at 0.30 mAHD (Figure 2-3), but only for one month. Herschel Lake surface water level registered above the maximum level criterion at 0.40 mAHD in May 2017 and below the minimum level criterion at -1.30 mAHD, in September and October 2017 (Figure 2-4).

Serpentine Lake registered a surface water level above the maximum level criterion at 0.06 mAHD and 0.02 mAHD in April and May 2017, respectively (Figure 2-5).

Lake Baghdad

06/2014 12/2014 07/2015 01/2016 08/2016 03/2017 09/2017 04/2018 10/2018

0.4

0.2 Max level 0.0

-0.2

-0.4

-0.6

-0.8 Min level -1.0 SWL SWL (mAHD) -1.2

-1.4

-1.6

-1.8

-2.0

Figure 2-2. Surface water levels (SWL) recorded as meters above Australian Height Datum (mAHD) of Lake

Baghdad from August 2014 to May 2018; including minimum and maximum level criteria.

21

Chapter 2

Garden Lake

11/2013 06/2014 12/2014 07/2015 01/2016 08/2016 03/2017 09/2017 04/2018 10/2018 0.4

0.2 Max level 0.0

-0.2

-0.4

-0.6 SWL SWL (mAHD) -0.8 Min level -1.0

-1.2

-1.4

Figure 2-3. Surface water levels (SWL) recorded as meters above Australian Height Datum (mAHD) of Garden

Lake from August 2014 to May 2018; including minimum and maximum level criteria.

Herschel Lake 22/11/2013 10/06/2014 27/12/2014 15/07/2015 31/01/2016 18/08/2016 6/03/2017 22/09/2017 10/04/2018 27/10/2018 0.6

0.4

0.2 Max level 0

-0.2

-0.4

-0.6 SWL SWL (mAHD) -0.8

-1 Min level

-1.2

-1.4

Figure 2-4. Surface water levels (SWL) recorded as meters above Australian Height Datum (mAHD) of Herschel

Lake from August 2014 to May 2018; including minimum and maximum level criteria.

22

Chapter 2

Serpentine Lake

11/2013 06/2014 12/2014 07/2015 01/2016 08/2016 03/2017 09/2017 04/2018 10/2018 0.4

0.2 Max level 0.0

-0.2

-0.4

-0.6 SWL SWL (mAHD) -0.8 Min level -1.0

-1.2

-1.4

Figure 2-5. Surface water levels (SWL) recorded as meters above Australian Height Datum (mAHD) of

Serpentine Lake from August 2014 to May 2018; including minimum and maximum level criteria.

The water balance for Garden, Herschel and Serpentine lakes, in addition to other three lakes that were not included in the present research – Government House, Timperley and Pearse lakes – was studied by Gillen (2015, unpub.), during the period from September 2014 to

September 2015 (Figure 2-6). For this period, the water balance analysis showed that the combined lakes storage decreases during summer when the evaporation fluxes are high and rainfall fluxes are low. Storage increases in winter/spring due to lower evaporation and higher rainfall. The net groundwater inflow is low to none when lake levels are high but as lake levels drop groundwater inflow increases until rainfall inflows occur in winter to spring and lake levels start to rise. The cycle then repeats the following year but timing of inflows will have some variability depending a particular year’s climate.

23

Chapter 2

8000 8000

6000 6000

4000 4000 /d) 3 2000 2000

0 0

-2000 -2000 Daily fluxes Daily fluxes (m -4000 -4000

-6000 -6000

-8000 -8000

Lake storage change (m3/d) Evaporation (m3/d) Groundwater inflow (m3/d) Rainfall (m3/d)

Figure 2-6. Overall water balance for Garden, Herschel and Serpentine lakes, in addition to other three lakes that were not included in the present research – Government House, Timperley and Pearse lakes – during the period of September 2014 to Spetember 2015. Data from Gillen (2015, unpub.).

2.3.2 Lake and pore water analysis

The pH of all lake water samples was slightly alkaline, ranging from 7.55 at site PL

(Pink Lake) in February 2016 to 8.60 at site PL in November 2015, and their temperature varied from 14.17 °C at site GN (Garden Lake) in July 2015 to 36.0 °C at site PL in January 2016

(Supplemental Table 2). Alkalinity measurements in lake water ranged from 184.00 mg/L (at site PL in November 2015) to 866.00 mg/L (at site PL in February 2016) and in pore water from 680.68 mg/L (at Herschel Lake in July 2015) to 1721.72 mg/L (at site BG at Baghdad

Lake in March 2016). Electrical conductivity (EC) varied throughout the sampling period and was similar among the lake water samples collected, registering a substantial increase from

November 2015 to January 2016, and then declining again in February 2016 (Figure 2-7). All the measurements for EC were above 120 mS/cm, i.e. hypersaline. The lowest EC obtained

24

Chapter 2 was 121.8 mS/cm for site GN in July 2015 and the highest EC was 314.3 mS/cm for site PL in

January 2016 (Figure 2-7). Among all the lakes, Garden Lake (site GN) exhibited generally the lowest EC, ranging from 121.8 mS/cm in July 2015 to 180.1 mS/cm in March 2016 (Figure

2-7).

350.0

300.0 LW-BG LW-GN 250.0 LW-HW LW-HN 200.0 LW-PL EC EC (mS/cm) LW-SN 150.0 LW-SW

100.0 15/07/2015 3/09/2015 23/10/2015 12/12/2015 31/01/2016 21/03/2016

Figure 2-7. Electrical conductivity (EC) fluctuation within the lake water samples, throughout the sampling period

– July and November 2015, and January, February and March 2016. Each line and color represents probe EC measurement for each lake water collected from the seven different sites – BG (Baghdad); GN (Garden North);

HN (Herschel North); HW (Herschel West); PL (Pink Lake); SN (Serpentine North); and SW (Serpentine West).

Similar patterns were observed for total dissolved solids (TDS), in which measurements during January 2016 were also the highest observed among the different sampling events for all lakes (Figure 2-8). TDS values were all above 35 g/L, which is a characteristic of hypersalinity (Shadrin and Anufriieva, 2019), and confirms that all lakes were hypersaline throughout the sampling period. The lowest and highest TDS results were also observed at site

GN in November 2015 (65.4 mg/L) and at site PL, in January 2016 (130.0 g/L), respectively

(Figure 2-8). Similar to EC analysis, site GN also presented the lowest range for TDS measurements – 65.4 g/L (July 2015) to 85.3 g/L (March 2016) (Figure 2-8).

25

Chapter 2

140.0

130.0

120.0 LW-BG LW-GN 110.0 LW-HW 100.0 LW-HN

TDS (g/L) 90.0 LW-PL LW-SN 80.0 LW-SW 70.0

60.0 15/07/2015 3/09/2015 23/10/2015 12/12/2015 31/01/2016 21/03/2016

Figure 2-8. Total dissolved solids (TDS) fluctuation within the lake water samples, throughout the sampling period – July and November 2015, and January, February and March 2016. Each line and color represents probe

TDS measurement for each lake water collected from the seven different sites – BG (Baghdad); GN (Garden

North); HN (Herschel North); HW (Herschel West); PL (Pink Lake); SN (Serpentine North); and SW (Serpentine

West).

Due to the impossibility of measuring EC and TDS concentration in the pore water samples using the probe HI 9829 Water Quality Meter (Hanna Instruments®), TDS was calculated by summing the major anion and cation concentrations in both lake and pore water samples (Supplemental Figure 2-1). Overall, the pore water samples showed lower TDS concentration than the respective lake water samples. For example, pore water from site GN showed TDS concentrations of 69.5 g/L in July 2015 and 148.9 g/L in March 2016; which were lower than the respective lake water samples from the same site – 96.7g/L in July 2015 and

190.9 g/L in March 2016 (Supplemental Figure 2-1). Lake water samples from site GN had the lowest EC and TDS concentrations, in comparison to the lake water samples from other lakes (Figures 2-6 and 7; Supplemental Figure 2-1).

26

Chapter 2

The relationship between TDS and EC is shown in Figure 2-9. TDS and EC are related, since more solids dissolved in the water means higher electrical conductivity. The general relationship that holds between electrical conductivity (EC) and TDS is of the form: TDS

(mg/L) = α x EC (µS/cm), where 0.55 < α < 0.75 (Freeze and Cherry, 1979), in which EC varies as function of the ions that account for TDS.

The field EC measurements collected for the study lakes were compared to the TDS data to assess whether a linear relationship held at high salinity (Figure 2-9) – data collected suggested that it does not. The most suitable trend line for the data is: TDS = 13.97872 x

(EC)0.7260452.

Figure 2-9. TDS and EC correlation (R2 = 0.789) in the lake water samples seasonally collected from the seven different sites – BG (Baghdad); GN (Garden North); HN (Herschel North); HW (Herschel West); PL (Pink Lake);

SN (Serpentine North); and SW (Serpentine West).

Anions and cations analysis revealed that all lake and pore water samples were Na-Cl dominant. The overall anions and cations concentration distribution for lake water was Cl- >> 27

Chapter 2

2- - + + + + - 2- - SO4 >> HCO3 and Na >> Mg >> K > Ca ; and for pore water was Cl >> SO4 >> HCO3 and Na+ >> Mg+ > K+ > Ca+, respectively (Supplemental Table 2-3). Anions and cations concentrations were generally higher in lake water samples when compared to pore water,

- except for bicarbonate ions (HCO3 ), which had higher concentrations in the pore water samples (Supplemental Table 2-3).

The piper diagram showed that, in terms of relative abundance of anions and cations, lake and pore water had similar ionic composition to both local rainwater (Crosbie et al., 2012) and seawater (Pilson, 2013) (Figure 2-10).

Figure 2-10. Piper diagram showing the relative anions and cations relative composition of lake and pore water samples collected from the seven sampling sites in July and November 2015, and January, February and March

2016. For comparison purposes, Perth rainwater data (Crosbie et al., 2012) and seawater data (Pilson, 2013) were included.

Seasonal water data for each study site (Figures 2-11 and 2-12) showed that major ion concentrations varied month to month with slightly different trends. Overall, major ion 28

Chapter 2 concentrations varied in all lake water samples and had a similar trend throughout the sampling period, with the exception of site PL (Figures 2-11). For pore water samples, the major ion concentration variation from the sites BG and GN were very similar (Figures 2-12A and 12B), whereas the pore water samples from site HW presented similar trend as site SN (Figures 2-

12C and 2-12D).

The seasonal lake water anion and cation concentration analysis revealed that from July

2015 to March 2016, the anion Cl- concentration increased in the sites BG, GN, HW, HN, SW and SN, from 53.12 g/L to 140.00 g/L (Figures 2-11A to 2-11F). The same was observed for the cations Na+ and Mg+, for which the lowest concentrations were 30.10 g/L and 3.93 g/L in

July 2015 and the highest were 76.00 g/L and 8.60 g/L in March 2016, respectively (Figures

- + 2-11A to 2-11F). For the same sites – BG, GN, HW, HN, SW and SN –, the ions HCO3 , Ca and K+ did not exhibit much variation, registering steady concentrations throughout the

2- sampling period (Figures 2-11A to 2-11F). The anion SO4 was the only ion to markedly exhibit different patterns in seasonal concentration. For Lake Baghdad and Garden Lake, the

2- concentration of SO4 decreased in January 2016, then considerably increased in February

2- 2016 (Figures 2-11A and 2-11B); whereas for Herschel and Serpentine Lakes, SO4 concentrations showed an overall constant increase throughout the sampling period (Figures

- 2- + + + 2-11C to 2-11F). For site PL, the anions Cl and SO4 and cations Na , Mg and K concentrations slightly increased from July to November 2015, followed by a considerable

2- increase in January 2016. In March 2015, SO4 concentration was even higher than the

- + previous month (Figure 2-11G). Similar to the other sites, HCO3 and Ca concentrations registered throughout the sampling period were generally steady (Figure 2-11G).

- 2 For pore water anions and cations analysis, at the sites BG and GN, the ions Cl , SO4 ,

Na+ and Mg+ concentrations showed a considerable increase in January 2016 compared to earlier sampling times, followed by a decline in February 2016, increasing again in March 2016 29

Chapter 2

2- (Figures 2-12A and 2-12B). The exception was SO4 at site BG, which maintained a higher concentrations than November 2015, from January 2016 onwards (Figure 2-12A). Overall, the

- + + ions HCO3 , K and Ca concentrations were steady throughout the sampling period, in the sites

BG and GN (Figures 2-12A and 2-12B). Pore water samples from sites HW and SN presented

- 2- + + similar seasonal results, having Cl , SO4 , Na and Mg concentrations declined from

November 2015 to February 2016, followed by a concentration increase in March 2016

- + + (Figures 2-12C and 2-12D). Once again, for sites HW and SN, the ions HCO3 , K and Ca concentrations were fairly steady throughout the sampling period (Figures 2-12C and 2-12D).

30

Chapter 2

B LW-GN A 180 LW-BG 35 180 35 160

160 ) ) ) 30 ) 3

30 3 3 3 140 140 25 25 120 120 100 20 100 20 80 15 80 15 60 60 10 10 40 40 5 5 Cl and Na (mg/L x 10 x (mg/LNa and Cl Other Ions (mg/L x 10 x (mg/LIons Other Other ions (mg/L x 10 x (mg/L ions Other Cl and Na (mg/L x 10 x (mg/LNa and Cl 20 20 0 0 0 0 26/05/2015 3/09/2015 12/12/2015 21/03/2016 26/05/2015 3/09/2015 12/12/2015 21/03/2016 LW-HN D LW-HW C 180 35 180 35 160 160 ) 30 ) ) ) 3

30 3 3 140 3 140 25 25 120 120 100 20 100 20 80 15 80 15 60 60 10 10 40 40

5 10 x (mg/LNa and Cl 5 Other ions (mg/L x 10 x (mg/L ions Other Cl and Na (mg/L x 10 x (mg/LNa and Cl 20

20 10 x (mg/L ions Other 0 0 0 0 26/05/2015 3/09/2015 12/12/2015 21/03/2016 26/05/2015 3/09/2015 12/12/2015 21/03/2016 LW-SW E 180 LW-SN 35 F 180 35 160 160 ) ) )

30 ) 30 3 3 3 140 3 140 25 25 120 120 100 20 100 20 80 15 80 15 60 60 10 10 40 40

Cl and Na (mg/L x 10 x (mg/LNa and Cl 5 5 Other ions (mg/L x 10 x (mg/L ions Other 20 10 x (mg/LNa and Cl 20 Others ions (mg/L x 10 x (mg/L ions Others 0 0 0 0 26/05/2015 3/09/2015 12/12/2015 21/03/2016 26/05/2015 3/09/2015 12/12/2015 21/03/2016 LW-PL G 180 35 160

30 ) 3

) 140 Anions and Cations 3 25 120 Cl HCO3 SO4 100 20 80 15 Na K 60 10 40 Ca Mg 20 5 Other ions (mg/L x 10 x (mg/L ions Other

Cl and Na (mg/L x 10 x (mg/LNa and Cl 0 0 15/07/2015 23/10/2015 31/01/2016 Figure 2-11. Anions and cations concentrations variation in lake water samples collected from sites: (A) BG –

Lake Baghdad; (B) GN – Garden North; (C) HN – Herschel North; (D) HW – Herschel West; (E) SN – Serpentine

North; (F) SW – Serpentine West; and (G) PL – Pink Lake. Lake water samples were collected throughout the sampling period – July and November 2015, and January, February and March 2016. Each line and color

- + + - + + 2- represents the major anions and cations analyzed – Cl , Na , Ca , HCO3 , K , Mg and SO4 .

31

Chapter 2

A PW-BG B PW-GN 180 12 180 12 160 160 ) ) ) 3 10 3

10 3 )

3 140 140 120 8 120 8 100 100 6 6 80 80 60 4 60 4 40 40 Other ions (mg/L x 10 x (mg/L ions Other Other ions (mg/L x 10 x (mg/L ions Other

2 10 x (mg/LNa and Cla 2 Cl and Na (mg/L x 10 x (mg/LNa and Cl 20 20 0 0 0 0 26/05/2015 3/09/2015 12/12/2015 21/03/2016 26/05/2015 3/09/2015 12/12/2015 21/03/2016

C PW-HW D PW-SN 180 12 180 12 160 160 ) ) ) 3

3 10

10 3 ) 3 140 140 120 8 120 8 100 100 6 6 80 80 60 4 60 4 40 40 Cl and Na (mg/L x 10 x (mg/LNa and Cl 2 10 x (mg/L ions Other 2 10 x (mg/L ions Other Cl and Na (mg/L x 10 x (mg/LNa and Cl 20 20 0 0 0 0 26/05/2015 3/09/2015 12/12/2015 21/03/2016 26/05/2015 3/09/2015 12/12/2015 21/03/2016

Anions and Cations

Cl Na Ca HCO3 K Mg SO4

Figure 2-12. Anions and cations concentrations variation in pore water samples collected from sites: (A) BG –

Lake Baghdad; (B) GN – Garden North; (C) HW – Herschel West; and (D) SN – Serpentine North. Pore water samples were collected throughout the sampling period – July and November 2015, and January, February and

- + + - + March 2016. Each line and color represents the major anions and cations analyzed – Cl , Na , Ca , HCO3 , K ,

+ 2- Mg and SO4 .

Mineral saturation indicies (SI) for the period of September 2014 to September 2015 in the lakes Garden, Herschel and Serpentine were analysed by Gillen (2015, unpub.), using

PHREEQC software (Parkhurst and Appleo, 2013). The three lakes were saturated with respect to dolomite (SI: 2.6 – 4.0), calcite (SI: 0.8 – 1.5), and aragonite (SI: 0.5 – 1.5), which indicates these lakes are thermodynamically favourable for precipitation of these minerals throughout the year (Figure 2-13). Conversly, the lakes seasonally fluctuate between positive and negative 32

Chapter 2

SI with respect to gypsum and anhydrite (suggesting they may form seasonally and then dissolve) and unsaturated with respect to sylvite and halite (Figure 2-13). Bischoff et al. (2020) have found that the lakes are typically saturated (positive SI) with respect to aragonite and calcite and have found ample evidence of their formation but not dolomite.

33

Chapter 2

A

B

C

Figure 2-13. Mineral saturation index of (A) Garden Lake; (B) Herschel Lake; and (C) Serpentine Lake during the period of September 2014 to Spetember 2015. Data from Gillen (2015, unpub.).

34

Chapter 2

Data from nutrients analysis (Supplemental Table 2-4) were compared with the guideline values from the Australian and New Zealand Guidelines for Fresh and Marine Water

Quality (ANZEEC & ARMCANZ, 2000). The results from nutrient analysis for lake and pore water samples were compared to the default trigger values for physical and chemical stressors for south-west Australia at slightly disturbed ecosystems. The nutrients considered were: total phosphorus (TP), filterable reactive phosphate (FRP), total nitrogen (TN), oxides of nitrogen

+ (NOx) and ammonium (NH4 ). For TP, only lake water samples from the sites GN and PL were below the guideline trigger value of 60 µg/L; and for FRP, all lake water samples analyzed presented values under the trigger value of 30 µg/L, whereas pore water samples were all above the guideline value (Supplemental Table 2-4). In addition, all lake and pore water samples had TN concentrations over the trigger value of 1,500 µg/L (Supplemental Table 2-4). For pore water samples, high concentrations of total phosphorus and nitrogen were most likely observed due to freeze-thaw during water sampling and storing, which may have caused cell lysis. The guideline trigger value for NOx concentration is 100 µg/L and, with the exception of lake water set of samples from sites GN and BG, all the other sites presented at least one lake

+ or pore water seasonal sample with values over the guideline limit. For NH4 , all pore water samples registered concentrations above the trigger value of 40 µg/L and only four lake water samples were below the guideline limit - site HW (July 2015), site GN (July and November

2015) and PL (November 2015) (Supplemental Table 2-4).

Metals, metalloids and rare earth elements (REE) analysis data were also compared with guideline values from the Australian and New Zealand Guidelines for Fresh and Marine

Water Quality (ANZEEC & ARMCANZ, 2000). The guideline shows trigger values for freshwater and marine water for different levels of protection. Considering that the hypersaline lakes at Rottnest Island are protection areas of high importance, the trigger values used for comparison were the ones set for level of protection equals to 99%. The guideline shows 35

Chapter 2 insufficient data (ID) for most of the metals and the results of the analysis for all lake and pore water samples were in majority below the detection limit (data not shown). For lake water analysis, among all metals, metalloids and rare earth elements (REE) analyzed, Boron was the only one that occurred over the guideline trigger value for all sites throughout the sampling period (data not shown). For all pore water analysis, the results for the following specimens were recorded as above the guideline limit throughout the sampling period – Aluminum,

Arsenic, Boron, Chromium, Copper, Nickel, Selenium and Zinc (data not shown).

2.4 Discussion

Microbial mats represent a natural model for studying microbial diversity and adaptation to extreme environments, and microbially induced mineral precipitation (Inskeep et al., 2013; Cabestrero and Sanz-Montero, 2018; Prieto-Barajas et al., 2018). Although progress has been made in understanding the major groups of microorganisms that influence microbial mat growth and microbialite formation (Dupraz et al., 2009; Pace et al., 2016; Wilmeth et al.,

2018), it is still challenging to unravel the integration of the various physicochemical processes occurring in microbialite‐accretion ecosystems. To better understand the ecological aspects that lead to the development of microbial mats and further microbialite formation, it is important to understand the links between microbial mat communities with the physicochemical parameters of the aquatic environment the live in. It is also relevant to understand the implications of seasonal changes in the physicochemical environment conditions onto to microbial mat growth and microbialite accretion. The present study advances our understanding on the current environmental limits that allow microbial mat development at Rottnest Island and how seasonal changes in physicochemical factors can affect the growth of this sensitive ecosystem.

At Rottnest Island, the microbial mats, especially from Garden and Herschel Lakes, are currently under environmental pressure, mainly due to land and water-use activities on the 36

Chapter 2

Island. In order to monitor the distribution of the microbial mats and whether changes in surface lake water level would affect the development of the mats (in particular pustular and non- lithifying loosely cohesive mats, found on the lake foreshore) surface water level and water balance data were assessed. It was observed that during the period of August 2014 to May 2018 at Garden, Herschel and Serpentine Lakes, that the surface water level was registered at least once above the maximum criterion (0.00 mAHD); and in Herschel Lake it was also registered twice during the same period surface water level below the minimum criterion (-1.00 mAHD), but not persistently. The likely causes for the surface water level being higher than the maximum criterion or lower than the minimum criterion are primarily related to the rainfall dominated winter and evaporation dominated summer. However, other influences on lake hydrology, such as increased recharge from irrigation or stormwater disposal, declining groundwater levels due to land use (increased vegetation density) and/or water use change may be also contribuiting (Vogwill et al., 2019, unpub.). Garden Lake, for example, appears to be experiencing a rising lake level trend (Figure 2-3), potentially due to increased groundwater input from irrigation activities on the nearby golf course, which may be leading to freshening the lake water, hence inhibiting the microbial mats in this lake to accrete carbonate and form microbialite structures. In addition, prolonged periods of raised surface water level in the lakes where the pustular mat is growing could inhibit the ability of this mat to produce microbially mediated carbonate, as they are most active in areas of fresh groundwater input, which occurs at or above the lake level. If the lake levels rise more quickly than groundwater levels, hypersaline lake water will flow into the littoral zone, reducing or halting groundwater inflow to the lake. A reduced flux of groundwater into the lake, in addition to the presence of hypersaline water on the littoral zone, may represent a threat to the pustular mats, which are more active when fresh water is present. Conversely, prolonged periods of lowered surface water level in the lakes may inhibit the ability of the blister and flocculent mats to develop as 37

Chapter 2 they would not be seasonally covered by hypersaline water and the zone of groundwater discharge may not include these mats. Therefore, water level fluctuations have to occur within a certain range in order to maintain pustular, blister and flocculent mats health status and current distribution. Recently, in Shark Bay (Western Australia), it was observed that seasonal changes in water level and ground water influx may be influencing the ephemeral nature of living dendrolitic microbial mats and limiting their organomineralization activity (Suosaari et al., 2016; Suosaari et al., 2018). In the present study, although changes in surface water level above or below the criteria were registered for sustained periods (two consecutive months), this did not occur frequently during the monitoring period. However, long-term monitoring is needed to determine whether surface water level alterations could affect the development of the microbial mats.

Blister, flocculent and pustular mat found in some of the studied sites are growing in association with microbialites and are actively precipitating carbonate (Appendix-II; K.

Bischoff, co-authored by J. Mendes-Monteiro in revision with Sedimentology). Microbialite formation is dependent on high alkalinity coupled with appropriate local concentrations of bivalent cations (e.g., Ca2+, Mg2+), and sulfur cycling in the microbial mats, which induce carbonate precipitation (Baumgartner et al., 2009; Dupraz et al., 2009). Zeyen et al. (2017) suggested that the minimum alkalinity value required for modern microbialites formation is

75.03 mg/L. The lowest alkalinity registered among all lake and pore water samples was more than two times higher (184.00 mg/L at Pink Lake). Therefore, it is not surprising that blister, flocculent and pustular mats are precipitating carbonate. For Garden Lake however, although the minimum alkalinity registered was 188.40 mg/L, the microbial mat found in this lake – non-lithifying loosely cohesive – does not seem to be actively precipitating carbonate, as no microbialite structures were found. The lack of carbonate precipitation in the microbial mats at

Garden Lake could be attributed to their microbial community composition. Recent studies 38

Chapter 2 compared the microbial taxonomy among blister, flocculent, pustular and non-lithifying loosely cohesive mats (Mendes Monteiro et al., 2019). It was observed that, although the non- lithifying loosely cohesive mat had higher abundance of Firmicutes and Archaea, which are known to influence carbonate precipitation, other important groups of microorganisms that contribute to microbialite accretion were less abundant, especially sulfur cycle-associated

Alpha- and Gammaproteobacteria, and Bacteroidetes (Mendes Monteiro et al., 2019). Other environmental factors (e.g., salinity and nutrients concentration) could be influencing the differences in the microbial mat taxonomic composition in Garden Lake in comparison to the other lakes; hence inhibiting the non-lithifying loosely cohesive mat to form carbonate structures. It is known, for example, that shifts, either up or down, in salinity and nutrient concentrations (both of which have occurred in Garden Lake) can cause changes in the microbial community composition of hypersaline mats and, as consequence, the lithification process can potentially be constrained (John et al., 2009; Dupraz et al., 2013; Gleeson et al.,

2016; Preisner et al., 2016).

Electrical conductivity (EC) and Total dissolved solids (TDS) concentration was also monitored and it was observed that all the lake and pore water samples had EC above 120 mS/cm and TDS above 35 g/L, characteristic of hypersaline water (Shadrin and Anufriieva,

2019). Microbial mats and microbialites are known to be common in hypersaline environments, because their microbial communities are able to thrive in high-salinity, in which competitors/grazers are rare to absent (Dupraz et al., 2011; Hubert et al., 2018). Therefore, the lakes at Rottnest Island have the adequate hypersaline environment that allows the development of microbial mats and microbialite structures. Although Garden lake water EC and TDS were within the range of hypersaline water, 121.8 mS/cm to 180.1mS/cm and 65.4g/L to 85.3g/L, respectively (Supplemental Table 2-2), this lake showed the lowest salinity among all the studied lakes and had a decreasing salinity trend. 39

Chapter 2

In San Salvador Island (Bahamas), microbial mats have experienced large shift in salinity (230g/L to 65 g/L) during 2011-2012, and it was observed significant changes in the microbial community structures and in the expression of genes involved in nitrogen and sulfur cycling (Preisner et al., 2016). In Lake Clifton (Western Australia), increases in salinity have been registered from late 1990s to recent years, which might also have caused changes in the dominant microbial communities in the system and possibly restrained microbialite formation

(Smith et al., 2010; Gleeson et al., 2016). Considering that shifts in salinity can potentially alter microbial mat communities, hence their development and capacity of forming microbialites, ongoing monitoring is recommended in the hypersaline lakes in Rottnest Island, specially

Garden Lake, to assess variations in both EC and TDS and whether it is impacting the microbial mats in these lakes. It is also important to monitor pore water salinity, as the current study revealed that pore water samples had lower TDS concentration than the respective lake water samples, indicating that pore water is fresher than lake water (Figure 2-9; Supplemental Table

2-2). These results, combined with surface water level rises inundating the pustular mat and inhibiting its growth, are a strong indication that this mat growing on the foreshore zone of the lakes, might require fresh groundwater inflow to develop/persist. Therefore, alterations in the pore water EC and TDS could also potentially impact pustular mat health, as it would show a change in inflowing groundwater quality or a reduction/cessation of groundwater inflow.

Anions and cations analysis revealed that pore water samples presented higher

- concentrations of ion bicarbonate (HCO3 ), when compared to lake water samples

(Supplemental Table 2-3). These results may suggest that groundwater could be a source of bicarbonate and the microbial mats, from which pore water was extracted, contain high

- concentration of this ion. It may also indicate that HCO3 is being liberated in the mats via carbonate dissolution or due to heterothophic activity. Therefore, these microbial mats are most likely involved in active carbonate cycling. 40

Chapter 2

Another line of evidence for this is the difference in seasonal variation between sulfate

2- (SO4 ) and other analyzed ions. For example, in lake water samples from Lake Baghdad and

Garden lake, while all other ions concentration follow the same pattern of enrichment towards

2- summer, SO4 concentrations decrease. This may be related to seasonal changes in activity of sulfate reducing bacteria, which are known to induce carbonate precipitation (Gallagher et al.,

2012; Pace et al., 2016; Zhu and Dittrich, 2016). Gillen (2015, unpub.) demonstrated similar

2- results for Garden, Herschel and Serpentine Lakes, where ion SO4 concentrations showed a notable drop during mid and end-summer 2015. Although, the same phenomena was observed for Garden Lake in both current and Gillen (2015, unpub.) studies, the microbial mats that are growing in this lake do not appear to be forming microbialites. This suggests that differences in the microbial community composition among the mats might be the reason for the mats in

Garden Lake not be actively precipitating carbonate. For example, the Archaea present in higher abundance in the non-lithifying microbial mat community, in comparison with the other mats (Mendes Monteiro et al., 2019) might be reducing sulfate but the low abundance of other microorganisms that also participate in the lithification process might be constraining the microbialite formation due to insufficient carbonate cycling. Changes in the hydrology of

Garden Lake (for example, salinity decrease), could be contributing to differences in the microbial community structure in the mats present in this lake; hence affecting carbonate precipitation. It is important to note that this may have been caused by the water quality changes in the measured data (presented herein) or by previous changes not observed in this study (John et al., 2009).

Data from nutrients analysis revealed that phosphorus and nitrogen compounds were generally over the default trigger values from the Australian and New Zealand Guidelines for

Fresh and Marine Water Quality (ANZEEC & ARMCANZ, 2000). It is important to note however, that this guideline is for freshwater and marine water environments only and it was 41

Chapter 2 used due to the inexistence of a guideline for hypersaline wetlands. Nonetheless, comparing the current results with John et al. (2009) at Serpentine, Herschel and Garden Lakes, it is evident that shifts in nutrients concentration have happened. For example, in Herschel and

Serpentine Lakes, phosphorus and nitrogen concentrations have increased in comparison to the results obtained at John et al. (2009); whereas in Garden Lake, nutrients levels were lower in the present study than 10 years ago (John et a., 2009). It is known that shifts in nutrient concentrations (in particular, increases), can affect the microbial community composition, which can then disturb carbonate accretion and microbialite formation (Dupraz et al., 2013;

Gleeson et al., 2016). Therefore, it is important to keep monitoring water quality to detect shifts in nutrients on ions in the hypersaline lakes at Rottnest Island, preferably paired with microbial mat taxonomic and functional analysis.

2.5 Conclusion

Understanding the interactions between microbial mat development and water quality is relevant for progressing scientific knowledge of microbialite formation as well as providing insights into better managing these sensitive ecosystems. The present study analyzed water physicochemistry parameters of hypersaline lakes at Rottnest Island, which sustain the growth of lithifying and non-lithifying mats. The results revealed that Garden Lake, in which only non- lithifying mats are growing, had the lowest salinity measurements and a decreasing salinity trend; which may be one of the factors that are inhibiting microbialite accretion in this lake. In addition, Graden Lake also presented a rising lake level trend that might have been caused by the increment of groundwater input due to the proximity of the golf course and its irrigation activities. This fact could also be contributing to decreasing Garden Lake’s salitinity, due to freshening, which in turn could be inhibiting the microbial mats in this lake to accrete carbonate. It was also observed that sulfur reduction, a key component of microbialite formation process, is actively happening in Lake Baghdad and Garden Lake, however there is 42

Chapter 2 no evidence of mat lithification at Graden Lake. It is possible that the microbial mat in Garden

Lake has been disturbed by current land and water-use activities at Rottnest Island, affecting the microbial community; hence carbonate precipitation. Water level (and water quality) fluctuations have to occur within a certain range in order to maintain pustular, blister and flocculent mats health status and current distribution.

Further studies are required to determine the environmental factors that might be inhibiting Garden Lake mats to form carbonate structures. In addition, it is important to keep monitoring these sensitive ecological communities, as it will guide the Rottnest Island

Authority on how to better manage these unique ecosystems, and provide further insights into the mechanisms by which they form microbialites.

43

Chapter 2

Supplemental Material

Supplemental Table 2-1. Lake and pore water samples collected from seven sampling sites on five hypersaline lakes at Rottnest Island.

Sample ID Site Water Coordinate Sampling events Number of

collected samples

LW-BG BG Lake water 31°59’41” S, 115°31’25” E Jul-15, Nov-15, Jan-16, 5

Feb-16, Mar-16

LW-GN GN Lake water 31°59’44” S, 115°32’12” E Jul-15, Nov-15, Jan-16, 5

Feb-16, Mar-16

LW-HN HN Lake water 31°59’44” S, 115°31’41” E Jul-15, Nov-15, Jan-16, 5

Feb-16, Mar-16

LW-HW HW Lake water 31°59’47” S, 115°31’53” E Jul-15, Nov-15, Jan-16, 5

Feb-16, Mar-16

LW-PL PL Lake water 32°0’3” S, 115°30’46” E Jul-15, Nov-15, Jan-16, 4

Feb-16

LW-SN SN Lake water 32°0’10” S, 115°31’33” E Jul-15, Nov-15, Jan-16, 5

Feb-16, Mar-16

LW-SW SW Lake water 32°0’25” S, 115°31’13” E Jul-15, Nov-15, Jan-16, 5

Feb-16, Mar-16

PW-BG BG Pore water 31°59’41” S, 115°31’25” E Jul-15, Nov-15, Jan-16, 5

Feb-16, Mar-16

PW-GN GN Pore water 31°59’44” S, 115°32’12” E Jul-15, Nov-15, Jan-16, 5

Feb-16, Mar-16

PW-HW HW Pore water 31°59’47” S, 115°31’53” E Jul-15, Nov-15, Jan-16, 5

Feb-16, Mar-16

PW-SN SN Pore water 32°0’10” S, 115°31’33” E Jul-15, Nov-15, Jan-16, 5

Feb-16, Mar-16

44

Chapter 2

Supplemental Table 2-2. Rottnest Island hypersaline lake water and microbial mat pore water chemistry.

Sample ID Depth pH Temperature Dissolved Oxygen Dissolved Oxygen Redox potential EC TDS TDS (m) (°C) (ppm) (%) (mV) (mS/cm) (g/L)* (g/L)** LW-HW-Jul15 0.48 7.95 15.96 2.31 66.10 137.40 165.1 99.7 167.7 LW-HW-Nov15 0.48 7.99 27.84 2.59 75.50 81.00 185.7 88.1 196.6 LW-HW-Jan16 0.28 7.85 26.03 1.65 77.60 140.20 256.0 125.6 238.1 LW-HW-Feb16 0.32 7.80 24.10 2.64 72.30 61.10 170.7 86.9 240.6 LW-HW-Mar16 0.27 7.92 30.35 3.05 100.20 46.60 211.3 95.9 244.5 LW-HN-Jul15 0.29 7.93 15.50 2.01 66.40 124.40 180.0 109.8 134.1 LW-HN-Nov15 0.30 7.92 26.31 5.82 163.70 117.20 184.2 89.6 197.5 LW-HN-Jan16 0.18 7.90 27.85 1.84 87.10 99.50 262.5 124.5 233.6 LW-HN-Feb16 0.20 7.89 25.20 2.75 73.80 18.10 171.1 85.3 223.7 LW-HN-Mar16 0.15 8.00 24.85 1.78 58.40 59.60 173.5 88.7 208.0 LW-GN-Jul15 0.29 8.20 14.17 4.78 98.00 172.30 121.8 76.7 96.7 LW-GN-Nov15 0.29 8.32 25.60 9.06 195.10 146.10 133.8 65.4 114.6 LW-GN-Jan16 0.14 7.97 29.29 3.95 109.30 63.00 179.6 83.1 154.7 LW-GN-Feb16 0.18 8.02 23.07 3.51 78.00 67.10 133.7 69.5 166.6 LW-GN-Mar16 0.21 7.81 27.93 2.45 68.70 68.70 180.1 85.3 190.9 LW-BG-Jul15 0.32 7.96 16.20 2.86 78.00 118.50 158.3 95.1 133.2 LW-BG-Nov15 0.30 8.44 27.53 8.43 206.80 103.40 159.5 76.2 161.2 LW-BG-Jan16 0.25 8.45 24.10 2.67 105.50 68.10 224.2 113.9 183.7 LW-BG-Feb16 0.21 8.06 24.37 2.86 71.10 109.40 156.1 79.1 196.3 LW-BG-Mar16 0.19 7.59 25.85 1.64 46.30 125.50 178.4 87.9 207.7 LW-SN-Jul15 0.18 8.06 17.40 3.35 95.60 98.10 181.4 105.7 126.3

45

Chapter 2

LW-SN-Nov15 0.20 7.93 27.85 6.27 175.20 108.20 187.5 89.4 193.9 LW-SN-Jan16 0.28 7.97 28.32 1.62 78.70 118.70 265.6 124.8 221.3 LW-SN-Feb16 0.25 7.79 26.70 3.31 93.20 60.90 174.9 84.8 221.2 LW-SN-Mar16 0.26 7.86 28.30 1.25 38.00 47.10 200.0 94.1 229.7 LW-SW-Jul15 0.25 7.98 16.50 3.25 98.60 154.30 168.5 100.4 130.3 LW-SW-Nov15 0.25 7.99 25.51 3.96 116.70 125.30 181.1 89.9 198.7 LW-SW-Jan16 0.20 8.07 28.63 1.57 76.40 88.90 267.9 125.3 228.8 LW-SW-Feb16 0.23 7.85 27.99 2.03 58.80 96.90 180.9 85.6 235.5 LW-SW-Mar16 0.21 7.87 27.21 1.13 35.00 55.20 196.3 94.4 244.2 LW-PL-Jul15 0.22 8.02 17.50 3.54 82.40 88.20 145.1 85.6 95.5 LW-PL-Nov15 0.25 8.60 26.29 8.52 224.00 78.50 165.4 81.4 163.4 LW-PL-Jan16 0.05 7.62 36.00 0.65 35.60 134.90 314.3 130.0 319.8 LW-PL-Feb16 0.05 7.55 25.91 2.61 70.10 152.10 164.9 81.3 284.6 PW-HW-Jul15 ------159.3 PW-HW-Nov15 ------180.3 PW-HW-Jan16 ------156.0 PW-HW-Feb16 ------79.6 PW-HW-Mar16 ------264.2 PW-GN-Jul15 ------72.4 PW-GN-Nov15 ------69.5 PW-GN-Jan16 ------214.5 PW-GN-Feb16 ------91.7 PW-GN-Mar16 ------148.9 PW-BG-Jul15 ------134.0

46

Chapter 2

PW-BG-Nov15 ------78.2 PW-BG-Jan16 ------106.0 PW-BG-Feb16 ------57.0 PW-BG-Mar16 ------143.8 PW-SN-Jul15 ------144.6 PW-SN-Nov15 ------173.8 PW-SN-Jan16 ------131.6 PW-SN-Feb16 ------66.1 PW-SN-Mar16 ------183.4 *TDS measurement from using the probe **TDS measurement from summing major anions and cations - Due to the methodology used to collect pore water, it was not possible to use the probe HI 9829 Water Quality Meter (Hanna Instruments®) to measure the parameter in this table.

47

Chapter 2

Supplemental Table 2-3. Major anions and cations concentration in the hypersaline lake water and microbial mat pore water samples collected from Rottnest Island hypersaline lakes.

Sample ID Ca Cl Mg K Na SO4 HCO3 (mg/L) (mg/L) (mg/L) (mg/L) (mg/L) (mg/L) (mg/L) LW-HW-Jul15 1009.0 98947.0 7085.2 2123.1 52524.5 5629.8 348.2 LW-HW-Nov15 986.0 109527.6 7766.8 2342.2 63878.6 11669.5 348.4 LW-HW-Jan16 1075.2 129869.9 9286.6 2762.7 76455.3 18156.5 363.4 LW-HW-Feb16 1143.3 131911.6 9553.3 2873.5 78488.9 16160.1 410.3 LW-HW-Mar16 930.0 140000.0 8600.0 2500.0 76000.0 16000.0 542.0 LW-HN-Jul15 796.4 83080.8 5650.3 1703.3 40659.6 1952.1 308.7 LW-HN-Nov15 995.8 110375.7 7563.3 2328.9 62422.5 13408.0 322.3 LW-HN-Jan16 1056.7 130967.1 8977.2 2751.6 73713.8 15589.7 387.3 LW-HN-Feb16 1088.1 123491.0 8829.1 2746.5 72100.7 15065.8 387.4 LW-HN-Mar16 790.0 120000.0 6800.0 2000.0 64000.0 14000.0 386.5 LW-GN-Jul15 1103.6 53119.1 3928.9 1180.6 30101.6 7047.3 230.6 LW-GN-Nov15 1257.5 63427.2 4365.9 1387.3 35733.1 8197.2 225.1 LW-GN-Jan16 1564.1 89328.1 6219.9 1890.2 48480.6 6940.7 276.3 LW-GN-Feb16 1501.2 91198.2 6230.0 1862.4 51598.9 13880.1 231.8 LW-GN-Mar16 970.0 110000.0 6600.0 2000.0 58000.0 13000.0 310.3 LW-BG-Jul15 1272.2 73794.7 5396.9 1598.4 41592.6 9322.8 309.5 LW-BG-Nov15 1371.5 84911.0 5870.7 1832.6 48915.1 18050.8 246.0 LW-BG-Jan16 1604.1 105865.6 7268.4 2169.3 57795.1 8610.5 332.2 LW-BG-Feb16 1638.6 106858.7 7666.4 2391.1 63111.1 14289.5 308.8 LW-BG-Mar16 1100.0 120000.0 7100.0 2100.0 63000.0 14000.0 350.0 LW-SN-Jul15 995.1 69824.3 5818.4 1709.4 42786.2 4478.6 298.6 LW-SN-Nov15 1047.3 112000.4 7837.0 2331.8 62348.1 7964.9 332.9 LW-SN-Jan16 1164.2 126631.4 9020.0 2761.9 71390.5 9821.0 425.5 LW-SN-Feb16 1170.0 120216.5 8849.0 2753.3 73138.7 14695.2 349.2 LW-SN-Mar16 960.0 130000.0 7900.0 2400.0 72000.0 16000.0 426.3 LW-SW-Jul15 915.7 73561.5 5393.7 1560.2 40758.8 7853.4 348.0 LW-SW-Nov15 1052.1 113277.8 7487.4 2335.7 62157.2 12033.2 334.9 LW-SW-Jan16 1190.7 125646.9 9017.9 2818.9 74200.9 15406.7 434.1 LW-SW-Feb16 1890.1 129049.2 9944.1 2769.8 73774.1 17598.2 387.6 LW-SW-Mar16 990.0 140000.0 8300.0 2500.0 76000.0 16000.0 465.0 LW-PL-Jul15 1124.1 52912.9 3726.9 1022.7 30064.4 6438.9 285.4 LW-PL-Nov15 1482.7 88017.6 6688.5 1921.6 54867.4 10039.7 215.9 LW-PL-Jan16 519.9 172388.7 26211.7 7016.3 89477.1 23096.0 855.3

48

Chapter 2

LW-PL-Feb16 366.8 148071.8 23292.8 5977.4 72574.9 33271.9 1052.8 PW-HW-Jul15 528.0 95300.0 3651.4 1103.1 50700.0 7433.1 830.4 PW-HW-Nov15 780.0 110000.0 6600.0 2000.0 55000.0 5600.0 1207.8 PW-HW-Jan16 696.5 97340.0 5566.6 2074.8 47291.3 2637.0 1343.3 PW-HW-Feb16 281.5 45143.5 3268.2 1289.3 26518.0 2869.1 976.6 PW-HW-Mar16 1335.1 159885.5 9540.0 3321.4 82800.0 6671.0 976.7 PW-GN-Jul15 825.7 41241.6 2567.6 966.6 22500.8 4196.1 903.7 PW-GN-Nov15 698.9 41570.0 2355.9 895.4 22422.0 1393.0 1159.0 PW-GN-Jan16 1133.1 125810.0 9225.0 1717.4 68719.5 7434.0 1978.4 PW-GN-Feb16 845.9 54630.0 3876.4 999.2 28200.0 2884.0 1880.7 PW-GN-Mar16 1047.5 90950.0 7103.4 1399.8 41265.1 6838.0 1856.3 PW-BG-Jul15 1220.4 75714.1 5258.3 1671.6 40608.9 9231.9 1135.7 PW-BG-Nov15 589.6 46360.0 1766.3 707.5 26100.0 2107.0 951.6 PW-BG-Jan16 806.1 62000.0 4109.0 1191.1 33630.5 4046.0 1514.3 PW-BG-Feb16 690.0 30000.0 2200.0 940.0 18000.0 4900.0 1978.4 PW-BG-Mar16 689.3 87466.8 3052.0 873.0 46950.0 4441.0 2100.5 PW-SN-Jul15 1046.1 82995.8 5903.0 1914.2 42997.0 9430.5 1160.2 PW-SN-Nov15 846.9 105830.0 6347.5 2039.3 54241.8 4159.0 976.0 PW-SN-Jan16 925.3 81380.0 4063.2 1570.3 40200.0 3146.0 1734.1 PW-SN-Feb16 785.3 39900.0 2543.9 932.0 19627.4 2201.0 1831.8 PW-SN-Mar16 861.4 109470.0 3355.0 1287.2 64925.0 3058.0 1502.1

49

Chapter 2

Supplemental Table 2-4. Nutrients analysis data from the hypersaline lake water and microbial mat pore water samples collected from Rottnest Island hypersaline lakes.

Total Filterable Total Nitrogen Dissolved Phosphorous Reactive Nitrogen (N- Oxides (N- Ammonium (N- + (P-TP) (µg/L) Phosphate (P- TN) (µg/L) NOx) (µg/L) NH4 ) (µg/L) PO4) (µg/L) Guideline trigger 60 µg/L 30 µg/L 1500 µg/L 100 µg/L 40 µg/L value Sample ID LW-HW-Jul15 93.54 1.41 5,385.37 251.36 361.48 LW-HW-Nov15 110.35 1.45 5,911.18 136.93 271.21 LW-HW-Jan16 142.85 8.22 7,151.63 47.58 222.12 LW-HW-Feb16 151.16 8.99 6,767.90 49.78 109.77 LW-HW-Mar16 156.02 4.76 7,064.69 17.81 16.16 LW-HN-Jul15 103.92 10.62 5,546.14 256.13 386.47 LW-HN-Nov15 120.46 10.00 6,477.25 13,622.90 301.50 LW-HN-Jan16 141.08

Chapter 2

PW-HW-Feb16 24,329.06 897.82 873,834.59 9.67 2,468.88 PW-HW-Mar16 13,894.92 12,345.08 903,940.03 368.30 46,827.80 PW-GN-Jul15 526.75 52.06 8,150.37 5.03 107.41 PW-GN-Nov15 729.87 569.97 19,759.00 66.34 3,031.32 PW-GN-Jan16 4,489.44 3,938.23 230,160.51 97.05 46,667.70 PW-GN-Feb16 5,550.66 4,858.72 393,062.01 144.03 43,093.02 PW-GN-Mar16 6,859.53 5,858.64 288,434.01 169.97 47,537.10 PW-BG-Jul15 1,000.72 100.49 65,506.44 4.50 1,731.66 PW-BG-Nov15 1,233.79 942.84 32,119.93 180.90 841.71 PW-BG-Jan16 11,764.78 1,078.32 338,964.84 14.33 2,141.51 PW-BG-Feb16 50,895.00 46,415.27 638,877.51 306.67 53,679.36 PW-BG-Mar16 19,086.66 17,080.50 456,225.51 826.85 39,841.35 PW-SN-Jul15 1,772.65 175.79 76,579.14 12.66 263.01 PW-SN-Nov15 708.27 649.53 10,821.75 81.14 304.43 PW-SN-Jan16 22,629.76 2,139.42 226,239.84 105.57 5,887.43 PW-SN-Feb16 30,390.00 28,105.22 568,604.01 684.44 56,242.74 PW-SN-Mar16 7,237.92 6,210.36 398,298.51 477.03 39,827.80 Results in bold are above the guideline trigger value.

51

Chapter 2

Supplemental Figure 2-1. Total dissolved solids (TDS) fluctuation within lake and pore water samples, throughout the sampling period – July and November 2015, and January, February and March 2016. TDS concentrations were measured as the sum of major anion and cation concentrations. Each line and color represents

TDS measurement for each lake or pore water collected from the seven different sites – BG (Lake Baghdad); GN

(Garden North); HN (Herschel North); HW (Herschel West); PL (Pink Lake); SN (Serpentine North); and SW

(Serpentine West).

52

Chapter 3

Chapter 3

Comparative metagenomics of microbial mats from hypersaline lakes at Rottnest Island

(WA, Australia), advancing our understanding of the effect of mat community and

functional genes on microbialite accretion

This chapter has been accepted for publication:

Mendes Monteiro, J., Vogwill, R., Bischoff, K., Gleeson, D. B., 2019. Comparative metagenomics of microbial mats from hypersaline lakes at Rottnest Island (WA, Australia), advancing our understanding of the effect of mat community and functional genes on microbialite accretion. Limnology and Oceanography, 00:1–17. doi:10.1002/lno.11323

(Accepted 14 August 2019).

53

Chapter 3

3.1 Introduction

Microbial mats are complex organosedimentary structures, usually organized as vertically stratified layers of specialized microbial consortia (Cantrell and Duval-Pérez, 2013;

Wong et al., 2016). They consist of self-sustaining communities of bacteria, archaea and, to a lesser extent, eukarya and viruses (Prieto-Barajas et al., 2018; Suosaari et al., 2018; Wong et al., 2018), in which all major biogeochemical processes occur (Bolhuis et al., 2014; Preisner et al., 2016). Microbial mat communities are embedded in a self-produced matrix of extracellular polymeric substances (EPS), wherein steep physicochemical gradients are created as a result of microbial community metabolism (Dupraz et al., 2009; Jungblut et al., 2016; Gutiérrez-

Preciado et al., 2018). These complex gradients, coupled with niche differentiation among the communities, enables a greater network of resources and energy exchange, increasing their survival capabilities (Paerl and Yannarell, 2010; Ruvindy et al., 2016).

Microbial mats are considered to be the earliest biological communities on Earth

(Awramik, 1976; Green and Jahnke, 2010), having first appeared in the Archaean Eon, approximately 3.7 billion years ago, and persisting for ~85% of the geological history of the planet (Dupraz and Visscher, 2005; Nutman et al., 2016). It is thought that these ancient ecosystems may have been the first occurrence of cellular organization on the planet (Reitner,

2011), having played a major role in modifying the composition of the atmosphere, by changing its redox condition via oxygenic photosynthetic process, nitrogen fixing and hydrogen production (Hamilton et al., 2016; Revsbech et al., 2016; Thomazo et al., 2018).

These changes in the ancient atmosphere have influenced biological evolution by, for example, paving the way for oxygen-dependent life (Dupraz et al., 2009; Olson et al., 2016). Modern microbial mats are considered relevant analogues of the early microbial mat ecosystems, hence, an important resource for studying the evolution of the ancient biosphere and biogeochemical cycles, and investigating microbial-microbial and microbial-environment interactions (Noffke

54

Chapter 3 et al., 2013; Preisner et al., 2016; Prieto-Barajas et al., 2018). For several decades, sedimentologists and geomicrobiologists have been studying these interactions to explain, for example, how microorganisms from microbial mats have shaped the planet’s sedimentary record and their role in mineral precipitation (Peimbert et al., 2012; Nutman et al., 2016;

Cabestrero and Sanz-Montero, 2018).

Microbially induced mineral precipitation processes involve major functional groups of Bacteria, including photolithoautotrophs, sulfate reducers, sulfide oxidizers, aerobic heterotrophs and extracellular polymeric substances (EPS) producers (Dupraz et al., 2009; Pace et al., 2016; Wilmeth et al., 2018). Within the modern microbial mat microenvironment, the occurrence of these different metabolic processes can lead to alterations in the local chemistry and induce carbonate precipitation (Dupraz et al., 2009; Zhu and Dittrich, 2016).

Photolithoautotrophs can favor carbonate precipitation by increasing the local alkalinity via raising the amount or availability of carbonate ions due to carbon dioxide uptake during photosynthesis (Glunk et al., 2011; Bundeleva et al., 2012; Jiang et al., 2013). Likewise, sulfate reducers can also increase pH, hence promoting carbonate precipitation, via sulfate consumption and organic matter decomposition (Gallagher et al., 2012; Pace et al., 2016; Zhu and Dittrich, 2016). Finally, EPS producers can influence carbonate precipitation by binding cations, such as calcium (Ca2+), to anionic functional groups present in the EPS matrix. These cations can then become available for calcium carbonate precipitation when EPS is degraded via heterotrophic metabolism (Glunk et al., 2011; Cabestrero et al., 2018; Sanz-Montero et al.,

2019). Archaea, eukarya and viruses also contribute to microbially induced mineral precipitation, directly or indirectly (Edgcomb et al., 2014, Prieto-Bajaras et al., 2018, White et al., 2018). For example, some archaeal metabolisms are linked to the reduction of sulfate, which is directly related to calcium carbonate precipitation (Prieto-Bajaras et al., 2018).

Eukaryotic organisms, such as metazoa and protists are potential bioturbators of microbialite

55

Chapter 3 mat structure and may contribute to more clotted structures and actively degrade/consume EPS

(Edgcomb et al., 2014). To date it is not clear how viruses influence microbialite accretion but it is known that viruses present in microbial mats may play key roles in modulating microbial diversity, impacting ecosystem function through infection and participating in biogeochemical cycles, for example, by recycling important nutrients within the microbial mats (White et al.,

2018).

The accretion of precipitated calcium carbonate in lithifying microbial mats may induce the formation of structures collectively known as microbialites (Burne and Moore, 1987;

Riding, 2011). Microbialites are classified according to their dominant internal macrofabric and the two main morphological types include stromatolites and thrombolites. Stromatolites are defined as having a laminated and thinly layered internal fabric whereas thrombolites have an unlaminated irregular clotted internal fabric (Kennard and James, 1986; Riding, 2011).

Although the processes of microbialite formation and how the different structure types are shaped are still not completely understood, studies have suggested that the taxonomic composition of the microbial community and the predominant metabolic pathways and surrounding hydrological and geochemical conditions can influence both (White et al., 2016;

Suosaari et al., 2016; Babilonia et al., 2018). In contrast to lithifying mats that form microbialites, non-lithifying microbial mats are composed of a large biomass usually enriched in photosynthetic microorganisms – e.g., cyanobacteria – with substantial amounts of EPS production (Khodadad and Foster, 2012; Fernandez et al., 2016). Within non-lithifying mats, instead of carbonate precipitation and microbialite formation, sporadic trapping of carbonate sand grains may occur due to the presence of EPS, which provides structural scaffolding, a means for microbial-sedimentary adherence and binding with the mats (Khodadad and Foster,

2012; Farías et al., 2014, Suosaari et al., 2018). To elucidate potential controls of microbialite formation processes, recent studies have assessed and compared the microbial community

56

Chapter 3 taxonomic and metabolic profiles of lithifying and non-lithifying microbial mats (Khodadad and Foster, 2012; Warden et al., 2016). Results have shown that both mat types present distinct taxonomic and functional genes profiles, revealing new insight into the taxa and functional capabilities of microbial mats that can potentially form microbialites in relation to non- lithifying mats (Warden et al., 2016; Farías et al., 2017; Babilonia et al., 2018).

Worldwide, modern analogues of microbial mats have their distribution limited to only a few locations compared to their abundance in the ancient Earth (Foster et al., 2009; Gerdes,

2010; Peimbert et al., 2012). Non-lithifying microbial mats that trap and bind sediments are more prevalent than microbial mats that promote lithification and form microbialites

(Havemann and Foster, 2008). Both mat types are restricted in their occurrence to a few extreme environments, such as, hypersaline lakes (Farías et al., 2017; Mlewski et al., 2018;

Shen et al., 2018), marine basins (Pagès et al., 2015; Ruvindy et al., 2016; Wong et al., 2017), geothermal springs (Pepe-Ranney et al., 2012; Nishida et al., 2018; Sharma et al., 2018) and oligotrophic freshwater environments (Bonilla-Rosso et al., 2012; Pajares et al., 2015; Ponce-

Soto et al., 2015). Previous studies that have compared the microbial communities and functional gene composition of lithifying and non-lithifying mats from hypersaline (Warden et al., 2016), marine (Khodadad and Foster, 2012; Babilonia et al., 2018), high temperature

(White et al., 2015) and freshwater environments (Bonilla-Rosso et al., 2012; Russell et al.,

2014) have revealed that the two different mat types – lithifying and non-lithifying – harbor distinct microbial communities and functional composition. The most well studied microbial mats are from hypersaline environments, including those from: Hamelin Pool (Shark Bay)

(Wong et al., 2018; D’Agostino et al., 2019); Lake Clifton (Gleeson et al., 2016; Warden et al.,

2016), in Western Australia; and Highborne Cay, in the Exuma Sound, The Bahamas (Casaburi et al 2016; Louyakis et al., 2017). These three locations are habitats for both lithifying and non- lithifying microbial mats (Khodadad and Foster, 2012; Warden et al., 2016; Babilonia et al.,

57

Chapter 3

2018). At Lake Clifton, metagenome analysis revealed that Cyanobacteria and genes associated with photosynthesis were more abundant in thrombolite-associated mats, whereas non- lithifying mats were enriched mainly with Bacteroidetes and Deltaproteobacteria, and genes associated with carbohydrate metabolism (Warden et al., 2016). A similar study at Shark Bay showed that non-lithifying mats were also more enriched with Proteobacteria in comparison to -associated mats (Babilonia et al., 2018). The same study also revealed that >1,000 functional genes were differently abundant between non-lithifying and lithifying microbialite- associated mats; for example stromatolite-forming mats had an abundance of genes associated with dissimilatory sulfur metabolism when compared to non-lithifying mats (Babilonia et al.,

2018). In Highborne Cay, Khodadad and Foster (2012) demonstrated that the two mat types presented few taxonomic differences in archaeal populations; however within the eukaryote communities, lithifying mats were more abundant in green algae than non-lithfying mats

(Khodada and Foster, 2012). Previous studies comparing both microbial mat types in marine environments have also recovered a number of viral sequences, but these represented less than

1% of the community within both lithifying and non-lithifying mats (Khodadad and Foster,

2012; Babilonia et al., 2018).

Rottnest Island, located 18km west of Fremantle in Western Australia (latitude 32o00

S, longitude 115o30 E), also harbors both lithifying and non-lithifying mats in the hypersaline lakes at the Island. These microbial mats are unique to the south-west region of Western

Australia (Rottnest Island Authority 2014) and are considered to be highly susceptible to environmental change, for example, alteration in salinity and nutrient concentrations (John et al., 2009). Historical and current land and water-use activities on the island may have led to elevated nutrient inputs to the hypersaline lakes via discharge of nutrient-rich fresh groundwater, potentially impacting the health of the microbial mats (Rottnest Island Authority,

2014). The microbial mats at Rottnest Island have been categorized as ‘Priority Ecological

58

Chapter 3

Communities’ under Western Australian Legislation – which covers naturally occurring and rare communities with possible threatened status – and are protected under the regulation of the Rottnest Island Authority (Rottnest Island Authority, 2014). Despite their relevance and threatened status, Rottnest Island microbial mat communities have not been characterized and are still poorly understood. In order to support Rottnest Island Authority to develop appropriate strategies to better manage and protect the existing microbial mats it is essential to understand how the resident microbial communities and their functional gene capacity contributes to microbialite formation. In addition, understanding the differences in the microbial diversity and functional capacity of both lithifying and non-lithifying microbial mats is relevant in elucidating the microbial processes associate with carbonate precipitation and microbialite formation.

In this study, the microbial communities of lithifying and non-lithifying mats from five different hypersaline lakes at Rottnest Island were characterized using metagenomic sequencing approaches, revealing their taxonomic diversity, as well as their functional gene composition. Both lithifying and non-lithifying microbial mats were compared to determine whether differences in their community structure and their functional gene composition may influence their capacity to form microbialites. This work reports for the first time the microbial composition and functional gene complexity of the Rottnest Island microbial mats, providing baseline information for future work. In addition, by comparing both lithifying and non- lithifying mats metagenomes, we expand our understanding of the processes underlying microbialite formation.

3.2 Material and methods

3.2.1 Study area and sampling sites

Microbial mat and water samples were collected in February 2016 from five hypersaline lakes (Garden Lake, Herschel Lake, Lake Baghdad, Lake Vincent and Serpentine

59

Chapter 3

Lake) at Rottnest Island, located off the coast of Perth (Western Australia; 32°00'7.20" S,

115°31'1.20" E) (Figures 2-1A and 2-1B, Chapter 2). In total eight different sampling sites were identified across the five lakes based on the predominant mat morphology observed in each lake. For each of Baghdad (BG) and Vincent (VC), microbial mat and water samples were collected from a single site. For Garden, Herschel and Serpentine, samples were collected from two sites each – sites Garden North (GN) and Garden South (GS); sites Herschel West (HW) and Herschel East (HE) and; sites Serpentine North (SN) and Serpentine West (SW) (Figure

3-1). Herschel, Garden and Serpentine were sampled at two separate sites due to the presence of different microbial mat morphologies at the two sites within each lake. At each site microbial mat samples were collected according to their appearance and texture, thus at some sites more than one mat morphology was collected. For example at the sites Baghdad (BG), Herschel

West (HW) and Serpentine North (SN) two microbial mats different in appearance and texture were present, hence both were collected. At the other five sampling sites (Garden North – GN;

Garden South – GS; Herschel East – HE; Serpentine West – SW; Vincent – VC) only one morphology of microbial mat was collected (for full details of all mats collected, see

Supplemental Table 3-1).

60

Chapter 3

Figure 3-1. Sampling sites at Rottnest Island hypersaline lakes – Lake Baghdad (site BG), Lake Vincent (site

VC), Garden Lake (sites GN and GS), Herschel Lake (sites HW and HE) and Serpentine Lake (sites SN and SW).

Aerial image from Google Earth using satellite data collected on January 1, 2018.

3.2.2 Lake and pore water sampling and analysis

Samples of lake and pore water were collected from the eight sampling sites. For lake water sampling, four aliquots were collected, stored in sterile 50 mL centrifuge tubes and each aliquot was used for total nutrients (nitrogen and phosphorus), dissolved nutrients

(ammonium), alkalinity and anions and cations analysis, respectively. Lake and pore water sampling procedure were performed as described in Chapter 2. Inductively coupled plasma- atomic emission spectrometry (ICP-AES) and high pressure ion chromatography (HPIC) were used for total nitrogen and phosphorus, dissolved ammonium, salinity and anions and cations analysis; and an Alkalinity Test Kit, Model AL-DT (HACH®), was used to measure alkalinity in the field. The probe HI 9829 Water Quality Meter (Hanna Instruments®) was used to measure pH (see Supplemental Table 3-2 for full details).

61

Chapter 3

3.2.3 Microbial mat sampling, DNA extraction and high throughput sequencing

Microbial mat samples were collected in triplicate using sterile plastic spoons and stored in labelled sterile bags at -18 °C until processed for DNA extraction. To avoid contamination, nitrile disposable gloves were worn during sampling, with new gloves used for each microbial mat sample collection.

Total DNA was extracted from microbial mat samples and then used as a template for high throughput metagenome sequencing. For DNA extraction each microbial mat replicate was separately homogenized using a sterile mortar and pestle by mixing thoroughly.

Approximately 50 mg of homogenized mat material was transferred to sterile 2 mL tubes containing sterile glass beads (0.2 g of 0.1 mm beads, 0.2 g of 0.5 mm beads, 8 units of 2.5 mm beads) and DNA was extracted using the PowerSoil® DNA Isolation Kit (QIAGEN, Inc.), following the manufacturer’s instructions. To increase the yield of DNA available for sequencing, DNA was extracted from multiple subsamples and pooled. The final DNA concentration of each pooled sample was checked using a Qubit® 2.0 Fluorometer (Life

Technologies Corporation, California). Total DNA from each sample (concentration ≥20 ng/µL) was subsequently sequenced at the Australian Genome Research Facility (AGRF) using the Illumina platform – HiSeq2500 paired end (100/150).

Paired-end Illumina reads were trimmed and screened using the open source BBTools package (Bushnell, 2017), then corrected using the high-performance tool for correcting sequence errors BFC version r181 (Li, 2015). Reads with no mate pair were removed. The resulting reads were then assembled with SPAdes assembler version 3.11.1-check (Nurk et al.,

2017), using a range of different k-mers. The entire filtered read set was mapped to the final assembly and coverage information generated using BBMap version 37.76 (Bushnell, 2017) using default parameters.

62

Chapter 3

The assembled reads were annotated using Metagenomic Rapid Annotations using

Subsystems Technology (MG-RAST) pipeline version 4.0.3 (Meyer et al., 2008). Before annotation, a quality control (QC) filtering was performed, which included dereplication

(Gomez-Alvarez et al., 2009), host specific species sequences screening (Langmead et al.,

2009), and low quality sequences trimming using a modified DynamicTrim (Cox et al., 2010).

Taxonomic and functional annotations were made based on sequence similarity searches against RefSeq and SEED Subsystems databases, respectively. Hierarchical taxonomic and functional abundance profiles were generated using Best Hit Classification, with a minimum alignment length of 15 bp, minimum e-value cutoff of 10-5, minimum abundance of 1 and a minimum percentage of identity cutoff of 60%, then downloaded for further analysis.

Taxonomic and functional annotated data are available on MG-RAST under the study ID mgp84373.

3.2.4 Visualization and statistical analysis

Taxonomic hierarchical data were plotted as multi-layered Krona charts (Ondov et al.,

2011) and for visualization purposes only the taxa with more than 1% of abundance are shown.

To assess statistical differences in microbial mat taxonomic diversity and their functional gene composition, data were analyzed using Primer v7 (Anderson et al., 2008) with the

PERMANOVA+ (permutational multivariate ANOVA) package (Clarke and Gorley, 2015),

DESeq2 package (version 1.24.0) (McMurdie and Holmes, 2014) in RStudio (version

1.2.1335) and the Statistical Analysis of Metagenomic Profiles (STAMP) package (version

2.1.3) (Parks et al., 2014).

MG-RAST annotations using RefSeq and SEED Subsystems databases were loaded into PRIMER v7 and a Bray-Curtis similarity resemblance matrix of both taxonomic (species level) and functional (level 4 – function level) profiles were created. In order to decrease the dominant contribution of abundant species and functions to Bray-Curtis similarities the data

63

Chapter 3 was square root transformed. Principal coordinate analysis (PCO) was performed based on the resemblance matrix constructed. Permutational multivariate analysis of variance

(PERMANOVA) with 999 unrestricted permutations of raw data were used to test the statistical significance of the differences in the taxonomic and functional profiles from the various microbial mat morphologies studied. Main and pair-wise tests were performed using type III sums of squares condition.

Differential abundance analysis was conducted to identify taxa more abundant in lithifying versus non-lithifying microbial mats, and among paired microbial mat morphologies, using the DESeq2 (version 1.24.0) (McMurdie and Holmes, 2014), Phyloseq (version 1.28.0)

(McMurdie and Holmes, 2013) and ggplot2 (version 3.1.1) packages in RStudio (version

1.2.1335). Differential abundance of OTUs between lithifying and non-lithifying mats, and among paired mat morphologies, were performed on variance-stabilized data with a significance level set to 0.005 and 0.01, respectively. The script for this analysis is available at https://github.com/Microbial-Ecology/DESeq2-for-Metagenome.

MG-RAST annotations were loaded into STAMP (version 2.1.3) (Parks et al., 2014) to determine functions that were significantly different between mat types and among microbial mat morphologies. The lithifying mats were tested against non-lithifying mats using pair-wise two-sided White’s non-parametric t-tests, which were conducted applying a confidence interval (CI) of 95% and using Benjamini-Hochberg FDR for multiple test corrections and the results were visualized as extended error bar plot. In order to determine the functions that were significantly different among microbial mat morphologies – blister, flocculent, pustular, non- lithifying cohesive and non-lithifying loosely cohesive – multiple group analyses were also performed in STAMP using Tukey-Kramer post-hoc test (confidence interval – 95%) and

Benjamini-Hochberg FDR for multiple test corrections. Multiple group statistics tables from

STAMP containing mean proportions and corrected p-values were used to plot mean

64

Chapter 3 proportion bar plots in RStudio (version 1.2.1335) using ggplot2 package (version 3.1.1). The script for this analysis is available at https://github.com/Microbial-Ecology/STAMP-outputs- barplots.

3.3 Results

3.3.1 Microbial mats description and water chemistry

Five different morphologies of microbial mats – blister, flocculent, pustular, non- lithifying cohesive and non-lithifying loosely cohesive – were identified at different locations within the five hypersaline lakes investigated at Rottnest Island (Figure 3-2; Supplemental

Table 3-1). Blister, flocculent and pustular mats can be defined as lithifying mats growing in association with microbialite structures, whereas non-lithifying cohesive and non-lithifying loosely cohesive can be defined as sediment-associated mats with no apparent lithifying activity. This has been investigated by a parallel study into the sedimentology of these deposits utilizing microscopy and mineralogical analysis (Appendix-II; K. Bischoff, co-authored by J.

Mendes-Monteiro, currently in revision with the Journal Sedimentology). The blister mat is a laminated orange-pink-purple mat with blister covered surface, usually growing in association with discontinuous layered stromatolites, found in the littoral zone around Herschel Lake and in the northern littoral zone of Serpentine Lake (Figure 3-2A; Supplemental Table 3-1). The flocculent mat is a brown-green-purple mat with a smooth-flocculent texture, growing associated with continuous layered stromatolites, found in the eastern littoral zone of Baghdad

Lake and the littoral zone around Vincent Lake (Figure 3-2B; Supplemental Table 3-1). The pustular mat is a black mat with visible pustules forming on the surface, growing in association with mottled thrombolites, found close to groundwater discharge on the lake foreshore of

Baghdad Lake (eastern area), Herschel Lake (throughout, except the eastern area) and

Serpentine Lake (throughout) (Figure 3-2C; Supplemental Table 3-1). The non-lithifying cohesive mat is a thin white-orange mat, growing on a thick layer of sediment (fine bioclastic

65

Chapter 3 sand), found in the southern littoral zone of Garden lake, only visible when the water level in this lake is low (dry season) (Figure 3-2D; Supplemental Table 3-1). The non-lithifying loosely cohesive mat is a white-blackish-green-purple layered mat full of sediment throughout with low cohesiveness, found on the lake foreshore of Garden Lake (northern area) and

Herschel Lake (eastern area) (Figure 3-2E; Supplemental Table 3-1). In total, 11 microbial mat samples were collected as biological triplicates (n=33) across all sampling sites – blister

(n=6); flocculent (n=6); pustular (n=12); non-lithifying cohesive (n=3); non-lithifying loosely cohesive (n=6). Technical replicates were not performed for this work due to the expense that would have been incurred for metagenomics sequencing.

66

Chapter 3

Figure 3-2. Microbial mats collected from Rottnest Island hypersaline lakes. (A) Blister mat collected from sites

HW and SN. (B) Flocculent mat collected from sites BG and VC. (C) Pustular mat collected from sites BG, HW,

SN and SW. (D) Non-lithifying cohesive mat collected from site GS. (E) Non-lithifying loosely cohesive mat collected from sites HE and GN.

Lake and pore water were also collected from the same sampling sites as the microbial mats. For each microbial mat sampled, a corresponding lake or pore water sample was collected. For example, lake water samples were collected to correspond to the microbial mats found in the lake littoral zones; whereas pore water samples were collected to correspond to the microbial mats growing on the lake foreshore, thus not covered by lake water. Although the non-lithifying cohesive mat was found in the littoral zone, as it was not covered by lake water during the field sampling (due to the low lake water level) a corresponding pore water sample was collected instead. The non-lithifying loosely cohesive mat collected from Herschel

East (HE) was particularly dry, hence no pore water was obtained after centrifuging. Therefore,

10 water samples (lake water – n=4; pore water – n=6) were collected in total (Supplemental

Table 3-2). Lake water pH was slightly alkaline, ranging from 7.82 to 8.06 (Supplemental

Table 3-2). Total alkalinity, total nutrients (nitrogen and phosphorus), and dissolved ammonium was generally higher in pore water than in lake water samples; whereas salinity was higher in lake water samples when compared to pore water (Supplemental Table 3-2).

67

Chapter 3

Anions and cations concentrations were generally higher in lake water samples, when

3- compared to pore water, except for anion phosphate (PO4 ), which presented higher concentrations in pore water samples (Supplemental Table 3-2). The overall ion concentration

- + 2- + + + 3- distribution for lake water and pore water were Cl > Na >> SO4 > Mg > K > Ca >PO4

- + + 2- + + 3- and Cl > Na >> Mg > SO4 > K > Ca >PO4 , respectively (Supplemental Table 3-2).

3.3.2 Microbial mat taxonomic composition

Assembled reads from 33 microbial mat metagenomes were annotated using MG-

RAST. A summary of the MG-RAST analysis statistics is provided in Supplemental Table 3-

3. The total base pair (bp) count for the assembled sequences was 1.27 Gbp and 1.06 Gbp before and after quality control, respectively (Supplemental Table 3-3). The total number of assembled sequences was 1.63 Mbp before quality control and 1.58 Mbp after (Supplemental

Table 3-3). After annotation, 2.87 Kbp rRNA features and 0.73 Mbp functional categories were identified (Supplemental Table 3-3). The MG-RAST statistics summary for each microbial mat morphology separately is also provided in Supplemental Table 3-3.

Microbial mat metagenomes from hypersaline lakes at Rottnest Island were predominantly composed of Bacteria (97.2%) with only 2.3% of the sequences assigned to

Archaea and the remaining 0.5% was representative of other sequences, including Viruses

(0.02%) and Eukaryota (0.48%). Among the Bacteria and Archaea a total of 34 phyla, 65 classes and 130 orders were recovered. The 3 main phyla that contributed to the taxonomic diversity within the microbial mats were Proteobacteria (52.9%), Bacteroidetes (20.8%) and

Cyanobacteria (9.9%). The phyla Actinobacteria, Firmucutes, Chloroflexi, Euryarchaeota and

Planctomycetes had minor representation of 3.5%, 3.3%, 2.1%, 2.0% and 1.5%, respectively.

The main classes in the analyzed metagenomes were Alphaproteobateria (36.4%),

Sphingobacteria (15.9%), an unclassified Cyanobacterial class (9.7%) and

Gammaproteobacteria (9.2%). Rhizobiales (20.7%), Sphingobacteriales (15.9%),

68

Chapter 3

Rhodobacterales (10.1%) and Chroococcales (4.6%) were the most dominant orders found in the microbial mat communities. Although Archaea represented only 2.3% of the total community within the microbial mats metagenomes, it is relevant to highlight that approximately 2.0% of the Archaea taxa were Euryarchaeota from Halobacteria class.

The lithifying and non-lithifying microbial mat metagenomes shared several annotated taxa, however the taxonomic composition and the relative abundance of the microbial communities in different mat types were distinct (Supplemental Figure 3-1). For example, blister, flocculent and pustular (all lithifying mats) were more abundant (p<0.05) in

Bacteroidetes from Rhodothermaceae family – 37%, 19%, 21%, respectively; whereas non- lithifying sediment associated microbial mats (both cohesive and loosely cohesive) were more abundant in (p<0.10) Cyanobacteria – 22% and 34%, respectively. Archaea from the

Halobacteriaceae family were also more abundant (p<0.10) in non-lithifying microbial mats

(3%-14%), in comparison to less than 1% of abundance in the lithifying mats. Proteobacteria were highly abundant in all microbial mat metagenomes, except for the non-lithifying loosely cohesive mat, which exhibited higher relative abundance of Cyanobacteria (34%) than

Proteobacteria (31%). Alphaproteobacteria were the major class of Proteobacteria across all five microbial mat morphologies. Gammaproteobacteria were the second major Proteobacteria class in the non-lithifying mats (11%-12%), being significantly more abundant (p<0.01) in those mats than in lithifying mats (2%-5%) (Supplemental Figure 3-1).

The beta-diversity of the five different microbial mat communities were also compared using principal coordinate analysis (PCO) (Figure 3-3). This analysis revealed that the microbial communities profiles from different microbial mat morphologies were significantly different (PERMANOVA, p<0.05) from one another in terms of taxonomic composition; with the metagenomes clustering based on the microbial mat morphologies from which they are originated (Figure 3-3). Overall, the metagenomes of the two stromatolite-associated mats,

69

Chapter 3 blister and flocculent, clustered together; whereas a second and a third cluster were formed by the metagenomes of pustular mat (thrombolite-associated), and sediment-associated mats, respectively (Figure 3-3). PERMANOVA analysis confirmed that the taxonomic profile of the different microbial mat morphologies were significantly different (pair-wise test, p<0.05) among each other. The PCO plots also revealed that taxa composition within microbial mats from the lake littoral zone and the foreshore clustered separately, suggesting mat location could also explain taxonomic variation within the mats (PERMANOVA, p<0.05; Figure 3-3).

Together with mat morphology, lake location accounted for 53.4% of the taxonomic variation

(PCO1 = 36.7%; PCO2 = 16.7%) (Figure 3-3).

Figure 3-3. Beta diversity of metagenomes from microbial mat communities in hypersaline lakes at Rottnest

Island. Principal coordinate analysis (PCO) based on Bray-Curtis similarity matrices calculated using square root transformation of OTU abundance data from RefSeq annotation (level VII – species). Different colors represent different microbial mat morphologies – blister, flocculent, pustular, non-lithifying cohesive, non-lithifying loosely cohesive; and different shapes represent microbial mat location within the lake – littoral zone and foreshore.

Differential abundance analysis revealed the main phyla, classes, families, and genera that contributed to the variations in the microbial mat communities observed in the PCO analysis (Figure 3-4; Supplemental Figures 3-2 to 3-11). Comparing lithifying and non- lithifying microbial mats, differential abundance analysis revealed 12 phyla, 21 classes, 52

70

Chapter 3 families and 65 genera that contributed to the microbial community variation between the two groups (p<0.05). For visualization purposes, only taxa with a level of significance lower than

0.005 were presented in the log2-fold-change bar plots (Figure 3-4). Results revealed that lithifying mats were enriched with the phyla Bacteroidetes, Spirochaetes, Planctomycetes and

Proteobacteria – particularly Alpha-, Beta- and Deltaproteobacteria –, when compared to non- lithifying mats (p<0.005). Within the phylum Gammaproteobacteria, taxa from the genera

Stenotrophomonas and Pantoea were more abundant in lithifying mats; whereas taxa from the genus Vibrio was more abundant in non-lithifying mats (p<0.005). When comparing the differential abundance of Firmicutes between the two groups, non-lithifying mats were more enriched in Clostridiaceae and lithifying mats were more abundant in Bacillaceae (p<0.005).

Within the Actinobacteria, lithifying mats had a higher abundance of Salinispora and

Mycobacterium, whereas non-lithifying mats were more enriched in Corynebacterium

(p<0.005) (Figure 3-4).

71

Chapter 3

Figure 3-4. Differentially abundant OTUs (agglomerated by species level) between lithifying and non-lithifying microbial mats at (A) class, (B) family and (C) genus level. Only OTUs with a significance level of 0.005 are shown. A less than zero log2-fold-change indicates phyla that were more differently abundant in lithifying microbial mats, whereas a greater than zero log2-fold-change indicates phyla that were more differently abundant in non-lithifying mats. Each circle represents one species that was enriched in the two types of mats; not a single

OTU. Different colors represent different phyla.

Comparing the paired mat morphologies, differential abundance analysis revealed that lithifying microbial mats had significantly higher abundance (p<0.01) of Proteobacteria in general, when compared to non-lithifying mats (Supplemental Figures 3-4, 3-5, and 3-7 to 3-

10). For example, Gammaproteobacteria were significantly more abundant (p<0.01) in the three lithifying mats when compared to non-lithifying mats (Supplemental Figures 3-4, 5, and

3-7 to 3-10). In addition, blister, flocculent and pustular mats also had a higher abundance

72

Chapter 3

(p<0.01) of Betaproteobacteria compared to non-lithifying mats (Supplemental Figures 3-5, and 3-8 to 3-10). Alphaproteobacteria, in particular Rhizobiales from the Brucellaceae and

Rhizobiaceae families, were significantly (p<0.01) more abundant in blister and pustular mats compared to non-lithifying mats (Supplemental Figures 3-4, 3-9 and 3-10). The two stromatolite-associated mats – blister and flocculent – were more abundant in

Gammoproteobacteria, whereas Alpha and Betaproteobacteria were more abundant in the thrombolite-associated pustular mat (Supplemental Figures 3-3 and 3-6).

Although the differential abundance analysis between mat types revealed that the lithifying mats were more abundant (p<0.005) in Bacilli and the non-lithifying mats were more abundant (p<0.005) in Clostridia (Figure 3-4); the same analysis among different mats morphologies reveal that both non-lithifying mats had in general higher abundance (p<0.01) of those two Firmicutes classes – Bacilli and Clostridia –, when compared to the microbialite- associated mats (Supplemental Figures 3-5, and 3-8 to 3-10). In contrast, taxa from the

Bacteriodetes familes Cytophagaceae, Flavobacteriaceae and Porphyromonadaceae were significantly more abundant (p<0.01) in blister, flocculent and pustular mats, when compared to both non-lithifying mats (Supplemental Figures 3-4, 3-5, and 3-7 to 3-9). In general, the pustular mat had a significantly higher abundance (p<0.01) of Spirochaetes and Cyanobacteria

(Chroococcales and Oscillatoriales taxa) when compared to most of the other mats

(Supplemental Figures 3-3, 3-6 and 3-10). The non-lithifying loosely cohesive mat also had a higher abundance (p<0.01) of Oscillatoriales when compared to blister mat (Supplemental

Figure 3-5).

Archaea from the phyla Crenarchaeota, Euryarchaeota and Korarchaeota were also significantly abundant (p<0.01) in non-lithifying loosely cohesive mat compared to both stromatolite-associated mats – blister and flocculent (Supplemental Figures 3-5 and 3-8). Few

Eukarya phyla were reported as significantly different in pairwise comparisons between the

73

Chapter 3 different mats morphologies. Both stromatolite-associated mats, blister and flocculent, had a higher abundance (p<0.01) of Ascomycota and Basidiomycota, respectively, when compared to pustular and non-lithifying loosely cohesive mats (Supplemental Figures 3-3 and 3-8).

3.3.3 Microbial mat functional gene composition

Principal coordinate analysis (PCO) was also used to compare the functional composition of the five different microbial mat communities (Figure 3-5). Overall, the metagenomes of the non-lithifying mats clustered separately to the lithifying mats – blister, flocculent and pustular (Figure 3-5). The functional gene composition of non-lithifying mats was significantly different (p=0.001) from microbialite-associated mats. PCO analysis showed both stromatolite-associated mats (blister and flocculent) clustering separately to the pustular thrombolite-associated mat; as well as non-lithifying cohesive mat clustering separately from non-lithifying loosely cohesive mat (Figure 3-5). PERMANOVA confirmed significant variance in the functional composition between the two groups of microbialite-associated mats

(p<0.05) and between non-lithifying cohesive and non-lithifying loosely cohesive mats

(p=0.025). PCO and PERMANOVA analysis also suggested that lake location – littoral zone or foreshore – explained variations in the functional gene profiles of the microbial mat metagenomes (PERMANOVA, p<0.05; Figure 3-5). Across all the samples, microbial mat morphology and lake location accounted for 38.9% of the functional gene composition variation (PCO1 = 29.4%; PCO2 = 9.5%) (Figure 3-5).

74

Chapter 3

Figure 3-5. Functional gene composition among metagenomes of microbial mat communities from hypersaline lakes at Rottnest Island, at function level. Principal coordinate analysis (PCO) based on Bray-Curtis similarity matrices calculated using square root transformation of functional gene abundance data from SEED Subsystems annotation (level IV – function). Different colors represent different microbial mat morphologies – blister, flocculent, pustular, non-lithifying cohesive, non-lithifying loosely cohesive; and different shapes represent microbial mat location within the lake – littoral zone and foreshore.

Pair-wise two-sided White’s non-parametric t-tests between lithifying and non- lithifying mats and Tukey-Kramer post-hoc tests among the five microbial mat morphologies revealed the main functional genes that contributed to the variations in the microbial mat communities (p<0.05) (Figure 3-6; Supplemental Figure 3-12). Comparing lithifying and non-lithifying microbial mats, difference in relative abundance analysis revealed 161 SEED

Subsystems Level III functions that contributed to the microbial community variation between the two groups (p<0.05). In order to better present the extended error bar plots, only key pathways that influence carbonate precipitation in microbialites were selected, for example carbohydrates, cofactors, nitrogen metabolism, photosynthesis, respiration, and sulfur metabolism pathways (Figure 3-6). The lithifying microbial mat communities had a higher abundance of functional genes from carbohydrate associated pathways (p<0.05), whereas non- lithifying microbial mats communities were more abundant in functional genes related to

75

Chapter 3 nitrogen metabolism, photosynthesis, cofactors and respiration pathways (p<0.05) (Figure 3-

6). Sulfate reduction associated complexes genes were also significantly more abundant

(p<0.05) in microbial communities within non-lithifying mats compared to lithifying mats

(Figure 3-6).

Comparing the different mat morphologies, photosynthesis associated genes related to

Photosystem II and Phycobilisome were significantly abundant (p<0.01) in the non-lithifying mats when compared to stromatolite associated blister and flocculent mats (Supplemental

Figure 3-12). The pustular mat was also significantly more abundant (p<0.01) in Photosystem

II and Phycobilisome genes in comparison to blister and flocculent mats (Supplemental

Figure 3-12). Other photosynthesis pathway genes associated with the cofactors category, for example Chlorophyll biosynthesis, were more abundant (p<0.01) in non-lithifying as well as pustular mats, in contrast to blister and flocculent mats (Supplemental Figure 3-12). The non- lithifying loosely cohesive and non-lithifying cohesive mats were significantly more abundant

(p<0.01) in genes related to sulfate reduction associated complexes and inorganic sulfur assimilation, respectively; in comparison to all lithifying mats (Supplemental Figure 3-12).

Nitrogen metabolism genes related to nitrogen fixation and nitrate and nitrite ammonification were significantly more abundant (p<0.01) in pustular and non-lithifying loosely cohesive mats compared to blister and flocculent mat (Supplemental Figure 3-12). In terms of carbohydrates pathways, CO2 uptake carboxysome genes were more abundant (p<0.01) in both non-lithifying microbial mat communities, whereas the gene associated with biogenesis of c-type cytochromes (respiration pathway) were more abundant (p<0.01) in the three lithifying mats – blister, flocculent and pustular (Supplemental Figure 3-12).

76

Chapter 3

Figure 3-6. Differences in relative abundance of functions from SEED Subsystems annotation (level III – third lowest classification in SEED out of four levels). Extended error plots showing pairwise two-sided White’s non- parametric t-tests, with corrected p-values calculated using Benjamini-Hochberg FDR approach (p<0.05); error bars show 95% confidence intervals. Functions are highlighted according to SEED Subsystems pathways (level

I): black – carbohydrates; blue – cofactors, vitamins, prosthetic groups, pigments; green – nitrogen metabolism; purple – photosynthesis; red – respiration and; orange – sulfur metabolism.

77

Chapter 3

3.4 Discussion

The microbial mats from hypersaline lakes at Rottnest Island are currently under environmental pressure, mainly due to land and water-use activities on the Island. The Rottnest

Island Authority has been making significant efforts to develop strategies to protect these unique microbial communities. Advancing our understanding of the microbial mats taxonomy and functional gene composition will provide relevant information about their environmental needs and can be used by the Rottnest Island Authority to better manage these sensitive communities.

Bacterial taxa dominated all microbial mat metagenomes within the Rottnest Island hypersaline lakes, with a low abundance of Archaeal taxa. Similar results were observed for lithifying and non-lithifying mats from Highborne Cay in Bahamas (Khodadad and, Foster

2012), Lake Clifton and Shark Bay in Western Australia (Warden et al., 2016; Babilonia et al.,

2018) which were mainly composed of Bacteria with a low Archaeal abundance. Another metagenome study reported that the Shark Bay microbial mat communities had an abundance of Archaea from the Euryarchaeota phylum, specifically halophilic taxa (Ruvindy et al., 2016).

In the current work, although Archaeal relative abundance was low, the main class identified was the Halobacteria, which is in agreement with the prior work of Ruvindy et al. (2016). Due to their tolerance of high salt concentrations, Halobacteria are usually dominant in hypersaline environments (Najjari et al., 2015). Given that the Rottnest Island hypersaline lake systems has a salinity of ~200 PSU (Practical Salinity Unit), it is not surprising that we report halotolerant archaea within the microbial mat communities.

Proteobacteria, Bacteroidetes and Cyanobacteria were the most abundant phyla in the

Rottnest Island microbial mat ecosystem. This is consistent with findings from other studies that have identified the same major phyla in microbial mats from hypersaline environments including Lake Clifton and Shark Bay (Wong et al., 2015; Ruvindy et al., 2016; Warden et al.,

78

Chapter 3

2016). Within Proteobacteria and Bacteroidetes, Alphaproteobacteria (e.g. Rhizobiales and

Rhodobacterales), Gammaproteobacteria and Sphingobacteria (particularly

Sphingobacteriales) were the prevalent taxa in the microbial mats at Rottnest Island, which is consistent with other studies that have also reported the abundance of these taxa in different microbial mats (Bonilla-Rosso et al., 2012; Peimbert et al., 2012; Fernandez et al., 2016;

Gleeson et al., 2016; Warden et al., 2016; Naghoni et al., 2017). Within Cyanobacteria, the order Chroococcales were prevalent in the microbial mats at Rottnest Island. This cyanobacterial order has been commonly reported as abundant in a wide range of microbial mat ecosystems, including freshwater, marine, and hypersaline conditions, for example, Exuma

Sound, The Bahamas (Mobberley et al., 2013; Casaburi et al., 2016), Lake Clifton and Shark

Bay, Western Australia (Suosaari et al., 2016; Warden et al., 2016) and Cuatro Ciénegas Basin

(Breitbart et al., 2009).

Overall, the microbiome of lithifying microbial mats (blister, flocculent and pustular) and non-lithifying microbial mats (cohesive and loosely cohesive) collected from the hypersaline lakes at Rottnest Island harbored taxa and functional attributes that differed depending on the morphology of the mat collected and the surrounding environment. In comparison to non-lithifying mats, the lithifying mats were more abundant (p<0.01) in Alpha-

, Beta-, Delta- and Gammaproteobacteria, in particular the Alphaproteobacteria from the order

Rhizobiales and the Gammaproteobacteria genera Stenotrophomonas and Pantoea.

Contrasting results were observed in Lake Clifton and Shark Bay (Western Australia), where non-lithifying microbial mats were more abundant in Alpha-, Delta- and Gammaproteobacteria

(Warden et al., 2016; Babilonia et al., 2018).

Alpha-, Delta- and Gammaproteobacteria play an important role in sulfur cycling, which is known to be important in microbialite formation (Baumgartner et al., 2006; Dupraz et al., 2009; Dupraz et al., 2011). For example, sulfur reducing Deltaproteobacteria perform

79

Chapter 3

-2 anoxygenic respiration using sulfate (SO4 ) as a terminal electron acceptor, reducing it to sulfide (HS-). During this process carbonate concentration is increased, hence alkalinity also increases; and if local concentration of calcium (Ca2+) is sufficiently high, calcium carbonate precipitation is promoted (Visscher and Stolz, 2005; Baumgartner et al., 2006). Anoxygenic photosynthetic purple non-sulfur Alphaproteobacteria (e.g., Rhizobiales) and several

Gammaproteobacteria (e.g., Stenotrophomonas and Pantoea) can also metabolize various sulfur-containing compounds, for example, purple non-sulfur Alphaproteobacteria can utilize sulfide yield from sulfate reduction as an electron donor (Basak and Das, 2007; Tourna et al.,

2013; Llorens-Marès et al., 2015). In oxygenic conditions, sulfide is oxidized promoting calcium carbonate dissolution (Visscher and Stolz, 2005). However, in microbial mat communities, the concentration of oxygen is usually limited due to its consumption by aerobic heterotrophic bacteria, for example. In this case, sulfide oxidation is incomplete, having as products thiosulfates, polysulfides, and polythionates (van den Ende and van Gemerden, 1993).

During incomplete sulfide oxidation, hydroxide anions (OH-) are also yield, which motivates the system to precipitate calcium carbonate and reestablish pH equilibrium (Visscher and Stolz,

2005; Baumgartner et al., 2006). The presence in abundance of sulfur cycle-associated Alpha- and Gammaproteobacteria in the lithifying microbial mats from Rottnest Island suggests that this group of bacteria could be directly related to microbialite formation in the hypersaline lakes that harbor those mats.

Bacteroidetes – e.g. from Cytophagaceae and Flavobacteriaceae families – were significantly more abundant in lithifying microbial mats than in non-lithifying mats. Similar results were observed when lithifying and non-lithifying mats from Lake Clifton (Warden et al., 2016) and Shark Bay (Babilonia et al., 2018) were studied. In both ecosystems,

Bacterioidetes, in particular Flavobacteriales, were more abundant in microbialite-associated mats than in the non-lithifying mats. Shark Bay lithifying mats were also highly abundant in

80

Chapter 3

Cytophagales compared to the non-lithifying mats (Babilonia et al., 2018). Bacteroidetes are mainly heterotrophic organisms that are known to have a preference for metabolizing high molecular weight organic matter, including EPS and cyanobacterial derived metabolites and, therefore may be playing a key role in carbon cycling and carbonate precipitation (Cottrell and

Kirchman, 2000; O’Sullivan et al., 2006; Chen-Harris et al., 2013). Therefore, the presence in abundance of Bacteroidetes in the lithifying microbial mats of Rottnest Island could be indicative of their key role in degradation of organic matter and release of nutrients in these systems (e.g., phosphate and ammonia) making these nutrients available to other members of the microbial mat community, potentially leading to further carbonate mineralization processes

(Ruvindy et al., 2016). However, further investigation is required to determine the exact role of specific Bacteroidetes in influencing carbonate precipitation and microbialite formation.

Among the five mat morphologies studied, the thrombolite-associated pustular mat had a significantly higher abundance (p<0.01) of Cyanobacteria, particularly Chroococcales and

Oscillatoriales. Previous research at Lake Clifton has also shown that Cyanobacteria, including

Chroococcales and Oscillatoriales, were more abundant in thrombolite-associated microbial mats compared to sediment-associated mats (Warden et al., 2016). In hypersaline mats,

Cyanobacteria are the dominant primary producers and are responsible for the production of a thick EPS matrix that, due to the presence of negatively charged functional groups (e.g., R-

COO−) with the potential to bind cations (e.g., Ca2+), creates mineral nucleation points for precipitation (Dupraz and Visscher, 2005; Richert et al., 2005; Dupraz et al., 2009). Thereafter, heterotrophic organisms can degrade EPS liberating Ca2+ and inorganic carbon to the microbial mat microenvironment, increasing the carbonate mineral saturation index (Braissant et al.,

2007; White et al., 2015). Cyanobacteria can also facilitate calcium carbonate precipitation by increasing the concentration of carbonate ions through carbon dioxide uptake during photosynthesis process (Visscher and Stolz, 2005). In order to reestablish equilibrium after an

81

Chapter 3 excessive consumption of carbon dioxide during photosynthesis, the concentration of carbonate in solution increases (Baumgartner et al., 2006), hence increasing pH to values higher than 10 (Visscher and van Gemerden, 1991). The high alkalinity combined with high concentration of Ca2+ due to EPS degradation motivates calcium carbonate precipitation to take place (Visscher and Stolz, 2005; Baumgartner et al., 2006). During these processes,

Cyanobacteria can influence the shape and mineralogy of microbialites (Gérard et al., 2018).

The higher abundance of Cyanobacteria in the thrombolite-associated pustular mats, compared to the stromatolite-associated mats, suggests that this taxa may be influencing the formation of microbialites with different internal macrofabric in hypersaline lakes at Rottnest Island.

Spirochaetes were also significantly more abundant (p<0.05) in pustular mats compared to most of the other mat morphologies. Although it is not known whether taxa within this phylum are more abundant or not in lithifying mats, when compared to non-lithifying mats, other studies have described Spirochaetes as common members of hypersaline microbial mats

(Schneider et al., 2013; Wong et al., 2015; Gleeson et al., 2016). The role of Spirochaetes in microbialite formation has not been well explored, but it is known that they are able to depolymerize EPS, which liberates Ca2+ into the mat microenvironment making it available to bind to carbonate ions; hence promoting carbonate precipitation (Dupraz et al., 2011).

Flocculent and blister mats are both understood to be stromatolite-associated mats, whereas the pustular mat is associated with mottled thrombolites. The process of degrading EPS by

Spirochaetes coupled with EPS polymerization by Cyanobacteria in the pustular mat may change physiochemical properties of the mat matrix, thereby influencing the phenotype and mineralization of the precipitates.

Non-lithifying mats had a higher abundance (p<0.01) of Firmicutes, particularly Bacilli and Clostridia, when compared to lithifying mats. Similar results were observed in Highborne

Cay, Clifton Creek (Yukon, Canada) and Cuatro Ciénegas Basin, where the non-lithifying mats

82

Chapter 3 analyzed also had higher relative abundance of Bacilli and Clostridia, in comparison to microbialite-associated mats (Bonilla-Rosso et al., 2012; Khodadad and Foster, 2012; White et al., 2015). Bacilli and Clostridia taxa are known to influence carbonate precipitation via EPS degradation – which creates mineral nucleation points for precipitation –, and urease decomposition – which provides a source of CO2 to the environment – (Schneider et al., 2013;

Zhu and Dittrich, 2016). Different Archaeal taxa were also significantly abundant (p<0.01) in non-lithifying loosely cohesive mats compared to blister and flocculent mats. The role of

Archaea in microbialite formation is still poorly understood; however it is hypothesized that this taxa could be performing key interactions with other microbial mat microorganisms, such as sulfate-reducing bacteria. These interactions create perfect microniches of oxic and anoxic zones, which allows, for example, carbonate precipitation (Wong et al., 2017). At Rottnest

Island hypersaline lakes, the higher abundance of Firmicutes and Archaea in the non-lithifying microbial mats compared to the lithifying mats suggests that, although the process of carbonate precipitation is not apparently occurring, these mat types do have the potential to form microbialites but they are not performing this role.

In terms of functional gene composition, non-lithifying mats were also significantly different (p=0.001) from lithifying mats. Dissimilarities within abundances of key pathway genes that affect carbonate precipitation were observed among the different microbial mat morphologies. For example, both of the non-lithifying mats had a higher abundance (p<0.05) of photosynthesis associated genes than the blister and flocculent stromatolite-associated mats.

As mentioned previously, during photosynthesis process carbon dioxide uptake can lead to a local increase in alkalinity, which can promote carbonate precipitation (Visscher and Stolz,

2005; Baumgartner et al., 2006). Therefore, it was hypothesized that the lithifying mats from

Rottnest Island would have more photosynthesis associated genes than the non-lithifying mats; however again the opposite was observed. Similar results were detected for sulfate reduction

83

Chapter 3 associated genes, which were significantly more abundant (p<0.05) in the non-lithifying microbial mats compared to all lithifying mats. The process of sulfur reduction within microbial mats is also know to increase local alkalinity, hence potentially promoting carbonate precipitation (Dupraz et al., 2009; Dupraz et al., 2011). The greater abundance of sulfate reduction genes in non-lithifying mats was not expected as these mats are not likely to be active in carbonate lithification processes. In terms of carbohydrates pathways, the non-lithifying mats were shown to have higher abundance of CO2 uptake carboxysome genes, which are components of photosynthetic pathways known to be associated with carbonate precipitation

(Kamennaya et al., 2012; Klanchui et al., 2017). These results suggest that the non-lithifying mats do have the potential to form microbialites – further investigation is warranted to determine why they do not form microbialites.

3.5 Conclusions

This study is the first thorough investigation of the microbial communities within the microbial mats from hypersaline lakes at Rottnest Island. The results demonstrate that different microbial mats morphologies harbor distinct microbial communities, particularly when comparing lithifying mats and non-lithifying mats. The results have also shown that both mat types comprise taxonomic diversity and functional gene capacity to have the potential to precipitate carbonate, although the process of microbialite formation only occurs within lithifying mats. The high abundance of specific groups of bacteria, archaea and functional genes that are known to play important role in microbialite formation within the non-lithifying mats is intriguing as there is no evidence of microbialite accretion in those mats. It is possible that water based activities on the island could be influencing the hypersaline lakes ecosystem thus potentially impacting the health of non-lithifying microbial mats and inhibiting the process of microbialite formation by these mats. Further research is needed to identify the environmental factors that are potentially affecting the health of these microbial mats. In

84

Chapter 3 addition, determining they key components of both lithifying and non-lithifying microbial mats from hypersaline lakes at Rottnest Island will provide insights into their microbial community and ecology thus expanding our understanding of the processes of microbialite formation.

85

Chapter 3

Supplemental material

Supplemental Table 3-1. Microbial mat morphologies collected from eight sampling sites on five hypersaline lakes at Rottnest Island.

Sample ID Morphology Site Coordinate Occurrence Description B-HW Blister HW 31°59’47” S, Herschel lake Orange-pink-purple mat with blistery 115°31’53” E littoral zone surface growing on stromatolites with B-SN Blister SN 32°0’10” S, Serpentine lake discontinuous layers (depth = 1.0 to 115°31’33” E littoral zone ~50.0 cm; thickness = 2.0 to 8.0 mm) F-BG Flocculent BG 31°59’41” S, Baghdad lake Brown-green-purple mat with smooth- 115°31’25” E littoral zone flocculent texture growing on F-VC Flocculent VC 31°59’59” S, Vincent lake littoral stromatolites with continuous layers 115°30’52” E zone (depth = 1.0 to ~50.0 cm; thickness = 2.0 to 8.0 mm) P-BG Pustular BG 31°59’41” S, Baghdad lake 115°31’25” E foreshore P-HW Pustular HW 31°59’47” S, Herschel lake Black mat with pustular texture 115°31’53” E foreshore growing on mottled thrombolites P-SN Pustular SN 32°0’10” S, Serpentine lake (depth = 0.0 cm; thickness = 1.0 to 5.0 115°31’33” E foreshore mm) P-SW Pustular SW 32°0’25” S, Serpentine lake 115°31’13” E foreshore NC-GS Non-lithifying GS 31°59’50” S, Garden lake littoral Cohesive non-lithifying white-orange cohesive 115°32’11” E zone, when lake mat growing on sediment (depth = 0 water level is low cm ; thickness = 1.0 to 3.0 mm) NL-HE Non-lithifying HE 31°59’47” S, Herschel lake Loosely cohesive non-lithifying loosely cohesive 115°31’59” E foreshore white-blackish-green-purple mat growing on sediment (depth = 0 cm ; NL-GN Non-lithifying GN 31°59’44” S, Garden lake thickness = 1.0 to 10.0 mm) loosely cohesive 115°32’12” E foreshore

86

Chapter 3

Supplemental Table 3-2. Rottnest Island hypersaline lake water and microbial mat pore water chemistry.

Total Total Total Dissolved Salinity* Cations (mg L-1) Anions (mg L-1) Analyzed Sample pH Alkalinity Nitrogen Phosphorus Ammonium (PSU) 2+ 2+ + + - 2- 3- water ID Mg Ca Na K Cl SO4 PO4 (mg L-1) (mg L-1) (mg L-1) (mg L-1)

B-HW 7.80 338.46 6.77 0.15 0.11 240.60 9553 1143 78489 2873 131912 16160

Lake B-SN 7.79 288.00 8.93 0.18 0.17 221.23 8849 1170 73139 2753 120216 14695

water F-BG 8.06 256.00 6.54 0.18 0.23 196.34 7666 1639 63111 2391 106859 14290

F-VC 7.82 224.00 8.04 0.20 0.10 214.66 8732 1022 72082 2619 116061 13718

P-BG - 1621.62 638.88 50.90 53.68 56.95 2200 690 18000 940 30000 4900 120.00

P-HW - 800.50 873.84 24.33 2.47 79.55 3268 281 26518 1289 45144 2869 4.75

P-SN - 1501.50 568.60 30.39 56.24 66.15 2544 785 19627 932 39900 2201 14.70 Pore P-SW - 940.94 571.57 15.62 36.30 19.17 788 149 5984 334 11130 681 3.00 water NC-GS - 1041.04 243.14 0.78 18.18 214.01 8433 1140 68900 1965 126252 6871

NL-HE ------

NL-GN - 1541.54 393.06 5.55 43.09 91.69 3876 846 28200 999 54630 2884 0.20

*Salinity values are sum of total dissolved solids converted to PSU (Practical Salinity Unit); **DL = Detection limit of 0.10 mg L-1

87

Chapter 3

Supplemental Table 3-3. Summary of MG-RAST analysis statistics of microbial mats metagenomes from hypersaline lakes at Rottnest Island.

Non- Non- lithifying Blister Flocculent Pustular lithifying Total loosely cohesive cohesive bp Count 272682133 276397963 455566381 115308833 150195016 1270150326 Sequences 349165 341932 569517 145316 221508 1627438 Count Mean Upload Sequence 785 787 810 789 675 - Length Mean GC 63 61 63 62 59 - (%) bp Count 228499028 227426994 380690615 92983338 127463908 1057063883 Sequences 337102 332068 553783 142356 209674 1574983 count Mean Post QC Sequence 680 669 694 651 605 - Length Mean GC 63 61 63 62 60 - (%) Predicted Protein 388049 388979 636352 165645 246132 1825157 Features Processed Predicted rRNA 1154 1290 2032 562 923 5961 Features Identified Protein 239231 244494 421711 104213 160661 1170310 Features Alignment Identified rRNA 503 678 905 259 522 2867 Features Identified Annotation Functional 162383 100328 279788 55198 128071 725768 Categories Values were calculating by summing the statistical results from each different microbial mat morphology, including replicates: blister (n=6); flocculent (n=6), pustular (n=12); non-lithifying cohesive (n=3); non-lithifying loosely cohesive (n=6).

88

Chapter 3

Supplemental Figure 3-1. Taxonomic composition from metagenomes of microbial mat communities in hypersaline lakes at Rottnest Island. Multi-layered Krona charts representing hierarchical taxonomy of the average community composition of (A) blister, (B) flocculent, (C) pustular, (D) non-lithifying cohesive and (E) non- lithifying loosely cohesive microbial mats. Only taxa with relative abundances greater than 1% are reported. Taxa are plotted on the same color scale.

89

Chapter 3

Supplemental Figure 3-2. Differentially abundant OTUs (agglomerated by species level) between blister and flocculent microbial mats at class and family level. Only OTUs with a significance level of 0.01 are shown. A less than zero log2-fold-change indicates phyla that were more differently abundant in blister mat, whereas a greater than zero log2-fold-change indicates phyla that were more differently abundant in flocculent mat. Each circle represents one species that was enriched in the two types of mats; not a single OTU. Different colors represent different phyla.

Supplemental Figure 3-3. Differentially abundant OTUs (agglomerated by species level) between blister and pustular microbial mats at class and family level. Only OTUs with a significance level of 0.01 are shown. A less than zero log2-fold-change indicates phyla that were more differently abundant in blister mat, whereas a greater than zero log2-fold-change indicates phyla that were more differently abundant in pustular mat. Each circle represents one species that was enriched in the two types of mats; not a single OTU. Different colors represent different phyla. 90

Chapter 3

Supplemental Figure 3-4. Differentially abundant OTUs (agglomerated by species level) between blister and non-lithifying cohesive microbial mats at class and family level. Only OTUs with a significance level of 0.01 are shown. A less than zero log2-fold-change indicates phyla that were more differently abundant in blister mat, whereas a greater than zero log2-fold-change indicates phyla that were more differently abundant in non-lithifying cohesive mat. Each circle represents one species that was enriched in the two types of mats; not a single OTU.

Different colors represent different phyla.

Supplemental Figure 3-5. Differentially abundant OTUs (agglomerated by species level) between blister and non-lithifying loosely cohesive microbial mats at class and family level. Only OTUs with a significance level of

0.01 are shown. A less than zero log2-fold-change indicates phyla that were more differently abundant in blister mat, whereas a greater than zero log2-fold-change indicates phyla that were more differently abundant in non- lithifying loosely cohesive mat. Each circle represents one species that was enriched in the two types of mats; not a single OTU. Different colors represent different phyla.

91

Chapter 3

Supplemental Figure 3-6. Differentially abundant OTUs (agglomerated by species level) between flocculent and pustular microbial mats at class and family level. Only OTUs with a significance level of 0.01 are shown. A less than zero log2-fold-change indicates phyla that were more differently abundant in flocculent mat, whereas a greater than zero log2-fold-change indicates phyla that were more differently abundant in pustular mat. Each circle represents one species that was enriched in the two types of mats; not a single OTU. Different colors represent different phyla.

Supplemental Figure 3-7. Differentially abundant OTUs (agglomerated by species level) between flocculent and non-lithifying cohesive microbial mats at class and family level. Only OTUs with a significance level of 0.01 are shown. A less than zero log2-fold-change indicates phyla that were more differently abundant in flocculent mat, whereas a greater than zero log2-fold-change indicates phyla that were more differently abundant in non-lithifying cohesive mat. Each circle represents one species that was enriched in the two types of mats; not a single OTU.

Different colors represent different phyla.

92

Chapter 3

Supplemental Figure 3-8. Differentially abundant OTUs (agglomerated by species level) between flocculent and non-lithifying loosely cohesive microbial mats at class and family level. Only OTUs with a significance level of

0.01 are shown. A less than zero log2-fold-change indicates phyla that were more differently abundant in flocculent mat, whereas a greater than zero log2-fold-change indicates phyla that were more differently abundant in non-lithifying loosely cohesive mat. Each circle represents one species that was enriched in the two types of mats; not a single OTU. Different colors represent different phyla.

Supplemental Figure 3-9. Differentially abundant OTUs (agglomerated by species level) between pustular and non-lithifying cohesive microbial mats at class and family level. Only OTUs with a significance level of 0.01 are shown. A less than zero log2-fold-change indicates phyla that were more differently abundant in pustular mat, whereas a greater than zero log2-fold-change indicates phyla that were more differently abundant in non-lithifying cohesive mat. Each circle represents one species that was enriched in the two types of mats; not a single OTU.

Different colors represent different phyla.

93

Chapter 3

Supplemental Figure 3-10. Differentially abundant OTUs (agglomerated by species level) between pustular and non-lithifying loosely cohesive microbial mats at class and family level. Only OTUs with a significance level of

0.01 are shown. A less than zero log2-fold-change indicates phyla that were more differently abundant in pustular mat, whereas a greater than zero log2-fold-change indicates phyla that were more differently abundant in non- lithifying loosely cohesive mat. Each circle represents one species that was enriched in the two types of mats; not a single OTU. Different colors represent different phyla.

Supplemental Figure 3-11. Differentially abundant OTUs (agglomerated by species level) between non- lithifying cohesive and non-lithifying loosely cohesive microbial mats at class and family level. Only OTUs with a significance level of 0.01 are shown. A less than zero log2-fold-change indicates phyla that were more differently abundant in non-lithifying cohesive mat, whereas a greater than zero log2-fold-change indicates phyla that were more differently abundant in non-lithifying loosely cohesive mat. Each circle represents one species that was enriched in the two types of mats; not a single OTU. Different colors represent different phyla.

94

Chapter 3

Supplemental Figure 3-12. Mean proportion of functions from SEED Subsystems annotation (level I and III – first and third classifications in SEED out of four levels), among the five different microbial mat morphologies.

(A) Mean proportion of level I functions (± 1 SD) with Benjamini-Hochberg FDR corrected p-value <0.01; (B)

Mean proportion of top 30 most significantly different level III functions (± 1 SD) with Benjamini-Hochberg FDR corrected p-value <0.01. Only level III functions from the following level I key pathways are shown: black – carbohydrates; blue – cofactors, vitamins, prosthetic groups, pigments; green – nitrogen metabolism; purple – photosynthesis; red – respiration and; orange – sulfur metabolism.

95

Chapter 4

Chapter 4

Seasonal variability in taxonomic and functional capacity of lithifying and

non-lithifying microbial mats from hypersaline lakes at Rottnest Island

(Western Australia)

This chapter has been prepared for publication and it will be submitted to Frontiers in

Microbiology:

Mendes Monteiro, J., Vogwill, R., Bischoff, K., Gleeson, D. B., 2020 (in prep.). Seasonal variability in taxonomic and functional capacity of lithifying and non-lithifying microbial mats from hypersaline lakes at Rottnest Island (Western Australia). Prepared to be submitted to

Frontiers in Microbiology.

96

Chapter 4

4.1 Introduction

To better understand the processes involved in the lithifying process this chapter presents a temporal expansion of the research in Chapter 3, which is a characterization of the mat microbial communities and their functions at one point in time. The research in Chapter 3 was undertaken on five different microbial mat morphologies, including lithifying and non- lithifying mat types, present in the hypersaline lakes at Rottnest Island. This study showed that lithifying and non-lithifying mat types comprise different microbial communities and metabolic genes. In addition, it also showed that, although not actively precipitating carbonate, the non-lithifying mats harbor key taxa and function genes that are relevant to microbialite accretion (Mendes Monteiro et al., 2019). Other studies have also aimed to unravel the process of microbialite formation and have shown that both lithifying and non-lithifying mat types present distinct taxonomic and functional genes profiles, revealing new insight into the taxa and functional capabilities of microbial mats that can potentially form microbialites in relation to non-lithifying mats (Khodadad and Foster 2012; Warden et al., 2016).

Although these prior studies have characterized microbialite-forming communities in different environments, most of them are based on single sampling event (including Chapter 3 of this thesis); hence, there is very little information available that deepens our understanding of how microbial mat communities respond to seasonal or temporal variation. Examining the microbial diversity and functional genes from microbial mats collected temporally (across different seasons or time) is fundamental to providing insights into how changes in the environment (i.e., alkalinity, salinity and nutrient concentration variation), can affect the microbial mat structure, hence influencing carbonate precipitation and microbialite formation.

A recent study assessed the response of microbial mat communities from two ephemeral lakes to seasonal changes (Cabestrero et al., 2018). It was observed that changes in salinity and alkalinity, mainly due to seasonal water cycle, have changed the taxonomic and metabolic

97

Chapter 4 composition of the microbial mats, which have directly affected local mineral saturation indices, hence altering the nature of precipitated carbonates. For example, during wet season, it was observed that fluids were saturated with respect to calcite and aragonite in the microbial mats, whereas in dry season these two mineral carbonates were undersaturated (Cabestrero et al., 2018)

At Rottnest Island, the microbial mats from the hypersaline lakes are likely to be experiencing environmental changes, such as variation in salinity, as seen in Chapter 2, and nutrient concentration (John et al., 2009). For example, John et al. (2009) have reported an increase in total nitrogen and total phosphorus in some of the lakes in Rottnest Island; and, in

Chapter 2, the seasonal water chemistry analysis have showed that one of the lakes have presented a decreasing salinity trend. It is unknown how the microbial mat communities might respond to these changes. In the current study, the microbial communities of different microbial mats from four hypersaline lakes at Rottnest Island were seasonally sampled and characterized using metagenomics sequencing approaches.

4.2 Material and Methods

4.2.1 Study area and sampling sites

Microbial mat and water samples were seasonally collected in July (winter) and

November (early-summer) 2015, and in January (mid-summer), February (mid-summer) and

March (end-summer) 2016, from hypersaline lakes (at Rottnest Island, located off the coast of

Perth (Western Australia; 32°00'7.20" S, 115°31'1.20" E) (Figure 3-1A; Figure 3-1B). From the eight sampling sites in the five hypersaline lakes assessed in Chapter 3, five sites from four lakes (Lake Baghdad, Garden Lake, Herschel Lake, and Serpentine Lake) were selected to perform the seasonal analysis, based on the predominant mat morphology observed in those sites and their seasonality. At each site, microbial mat samples were collected according to their appearance and texture. For Lake Baghdad, Garden Lake and Serpentine Lake, microbial

98

Chapter 4 mat samples were collected from a single site each – site Baghdad (BG), site Garden North

(GN) and site Serpentine North (SN); respectively. For Herschel Lake, mat samples were collected from two sites – sites Herschel West (HW) and Herschel East (HE) (Figure 3-1C).

Herschel Lake was sampled at two separate sites due to the presence of the lithifying mats blister and pustular at site HW and the non-lithifying loosely cohesive mat at site HE throughout the seasonal sampling period (Figure 3-1D, F and H). For site Serpentine North

(SN), blister and pustular mats were also present throughout the sampling period, hence both were collected. At the other two sampling sites (Baghdad – BG and Garden North – GN) only one morphology of microbial mat was present during the seasonal period, hence only one morphology was collected from these three sites (Figure 3-1F and H; for full details of all mats collected, see Supplemental Table 3-1).

4.2.2 Microbial mat sampling, DNA extraction and high throughput sequencing

Seasonal microbial mat samples were collected in triplicate and DNA was extracted as described in Chapter 3. At least 20 ng/µL of DNA from each sample was used as a template for high throughput metagenome sequencing and sequenced at the Australian Genome

Research Facility (AGRF). Paired-end Illumina reads were trimmed and screened using the open source BBTools package (Bushnell 2017), then corrected using the high-performance tool for correcting sequence errors BFC version r181 (Li 2015). Reads with no mate pair were removed. The resulting reads were then assembled with SPAdes assembler version 3.11.1- check (Nurk et al., 2017), using a range of different k-mers. The entire filtered read set was mapped to the final assembly and coverage information generated using BBMap version 37.76

(Bushnell 2017), using default parameters.

The assembled reads were annotated using Metagenomic Rapid Annotations using

Subsystems Technology (MG-RAST) pipeline version 4.0.3 (Meyer et al., 2008). Before annotation, a quality control (QC) filtering was performed as described in Mendes Monteiro et

99

Chapter 4 al. (2019). Taxonomic and functional annotations were made based on sequence similarity searches against RefSeq and SEED Subsystems databases, respectively. Hierarchical taxonomic and functional abundance profiles were generated using Best Hit Classification, with a minimum alignment length of 15 bp, minimum e-value cutoff of 10-5, minimum abundance of 1 and a minimum percentage of identity cutoff of 60%, then downloaded for further analysis. Taxonomic and functional annotated data are available at MG-RAST under the study ID mgp84373 and mgp85454.

4.2.3 Visualization and statistical analysis

To better understand how seasonality influenced the microbial communities throughout the sampling period, the relative abundances of the Class taxa from the three different microbial mat morphologies were plotted as stacked bar charts in RStudio (version 1.2.1335) using the ggplot package (version 3.1.1). To statistically assess seasonal differences in microbial mat taxonomic diversity and functional gene composition, data were analyzed using Primer v7

(Anderson et al., 2008) with the PERMANOVA+ (permutational multivariate ANOVA) package (Clarke and Gorley 2015), DESeq2 package (version 1.24.0) (McMurdie and Holmes

2014) in RStudio (version 1.2.1335) and the Statistical Analysis of Metagenomic Profiles

(STAMP) package (version 2.1.3) (Parks et al., 2014).

MG-RAST annotations using RefSeq and SEED Subsystems databases were loaded into PRIMER v7 and a Bray-Curtis similarity resemblance matrix of both taxonomic (family level) and functional (level 4 – function level) profiles were created. No transformation was required to decrease any dominant contribution of abundant families and functions to Bray-

Curtis similarities. Principal coordinate analysis (PCO) and metric multidimensional scaling

(mMDS) on distance among centroids was performed based on the resemblance matrix constructed. Permutational multivariate analysis of variance (PERMANOVA) with 999 unrestricted permutations of raw data were used to test the statistical significance of the

100

Chapter 4 differences in the taxonomic and functional profiles among the paired seasons with each microbial mat morphology (namely blister, pustular and non-lithifying loosely cohesive) and between lithifying and non-lithifying mats in each season. Main and pair-wise tests were performed using type III sums of squares condition.

Differential abundance analysis was conducted to identify taxa more abundant among paired seasons within each microbial mat morphology, and between lithifying and non- lithifying microbial mats in each season, using the DESeq2 (version 1.24.0) (McMurdie and

Holmes 2014), Phyloseq (version 1.28.0) (McMurdie and Holmes 2013) and ggplot2 (version

3.1.1) packages in RStudio (version 1.2.1335). Differential abundance of OTUs among paired seasons within the mats, and between lithifying and non-lithifying mats, were performed on variance-stabilized data with a significance level set to 0.005. The script for this analysis is available at https://github.com/Microbial-Ecology/DESeq2-for-Metagenome.

MG-RAST annotations were loaded into STAMP (version 2.1.3) (Parks et al., 2014) to determine functions that were significantly different among paired seasons within each mat morphology and between lithifying and non-lithifying microbial mats in each season. Two group analyses of paired seasons were performed on in STAMP using two-sided White’s non- parametric t-tests, which were conducted applying a confidential interval (CI) of 95%. Two group statistics tables from STAMP containing mean proportions and p-values were used to plot mean proportion bar plots in RStudio (version 1.2.1335) using ggplot2 package (version

3.1.1). The script for this analysis is available at https://github.com/Microbial-

Ecology/STAMP-outputs-barplots.

101

Chapter 4

4.3 Results

4.3.1 Seasonal variation in the taxonomic composition of the microbial mats

Blister and pustular microbial mats, defined as lithifying mats, and loosely cohesive non-lithifying mats have been described in detail in Chapters 2 and 3. Briefly, the overall relative abundance of the microbial mats was comprised of Alphaproteobacteria (11%-24%),

Actinobacteria (8%-23%), Gammaproteobacteria (14%-22%), Clostridia (5%-11%), Bacilli

(4%-10%), Betaproteobacteria (5%-9%) and Cyanobacteria (2%-6%) (Figure 4-1). Although some taxa were relatively steady through the sampling seasons (i.e., Betaproteobacteria and

Deltaproteobacteria in the three mat morphologies), some changes in the relative abundance throughout the sampling period were observed. For example, in the blister mat, the relative abundance of Actinobacteria increased from 9% in July (winter) and November (early- summer) 2015 to 22% in January (mid-summer) 2016, and then decreased to 12% in March

(end-summer) 2016 (Figure 4-1). In the same mat morphology, the relative abundance of

Alphaproteobacteria decreased slightly from 18% in November 2015, to 13% in January and

February (mid-summer) 2016; whereas Gammaproteobacteria had a higher relative abundance in February 2015 – 21% –, when compared to the other sampling events – 16% in November

2015 and January 2016; 17% in July 2015 and March 2016 (Figure 4-1). The relative abundance of Bacilli also showed some seasonal change, increasing from 6% in November

2015 to 10% in March 2016. Clostridia relative abundance was reported as 11% in November

2015, which was higher than in July 2015 and January and February 2016, all registering relative abundance of 6% (Figure 4-1). For Cyanobacteria (unclassified class), the relative abundance in the blister mat was 6% in March 2016, compared to 2% for the other sampling events (Figure 4-1). For the pustular mat, the largest seasonal changes occurred in the classes

Actinobacteria and Alphaproteobacteria. In this mat, the relative abundance of Actinobacteria went from 9% in July 2015 to 13% in November 2015, whereas Alphaproteobacteria registered

102

Chapter 4 a relative abundance of 24% in July 2015 and 12% in January 2016 (Figure 4-1). For the non- lithifying loosely cohesive mat, seasonal changes in the metagenomes relative abundance were also observed but were not as evident as in the two lithifying mats. For example, the lowest and highest relative abundance of Actinobacteria in the non-lithifying mat was 9% in July and

November 2015 and January 2016 and 12% in March 2016, respectively, showing only a 3% difference; whereas the blister mat had lowest and highest relative abundances of the same class at 9% and 22%, respectively. Another example was the Alphaproteobacteria, which had highest and lowest relative abundances of 15% in November 2015 and 11% in January 2016 for the non-lithifying loosely cohesive mat, compared to 24% and 12% in the pustular mat

(Figure 4-1).

103

Chapter 4

Figure 4-1. Seasonal taxonomic composition from metagenomes of microbial mat communities in hypersaline lakes at Rottnest Island. Stacked bar charts representing the relative abundance of different Classes in blister, pustular and non-lithifying loosely cohesive mat in the different sampling events – July and November 2015, January, February and March 2016. Classes with relative abundance lower than 2% were represented as “Others”.

104

Chapter 4

The beta-diversity of the three different microbial mat communities at the five sampling events were compared using principal coordinate analysis (PCO) and metric multidimensional scaling (mMDS) on distance among centroids (Figure 4-2). The PCO analysis showed the metagenomes clustering based on the mat morphologies and a clear segregation between the two lithifying mats – blister and pustular – and the non-lithifying mat (Figure 4-2A). The first two principal coordinate axes explained a total of 67.9% of the variation among the data (PCO1

= 54.6%; PCO2 = 13.3%) (Figure 4-2A). The PERMANOVA analysis confirmed that the three mat morphologies had significant differences in their taxonomic profiles overall (p<0.005).

The variations in taxonomy from one season to another for each mat morphology is shown in the mMDS on distance among centroids (Figure 4-2B). Although the mMDS plot suggests that the taxonomic profile of all three mat morphologies is varying seasonally (Figure 4-2B),

PERMANOVA analysis revealed that significant variation from one season to another only exists for the blister and pustular mats (p<0.02); with the exception of July 2015 to November

2015 in the blister mat, and February 2016 to March 2016 (p>0.10) in the pustular mat.

105

Chapter 4

Figure 4-2. Beta diversity of metagenomes from microbial mat communities in hypersaline lakes at Rottnest Island. (A) Principal coordinate analysis (PCO) and (B) metric multidimensional scaling (mMDS) on difference among centroids based on Bray-Curtis similarity matrices calculated using OTU abundance data from RefSeq annotation

(level V – family). Different colors represent different microbial mat morphologies – blister, pustular, non-lithifying loosely cohesive; and different shapes represent sampling events – July and November 2015, January, February and March 2016.

106

Chapter 4

Differential abundance analysis revealed the main phyla, classes and families that contributed to the seasonal variations in the blister and pustular microbial mat communities observed in the mMDS analysis (Figures 4-3 and 4-4; Supplemental Figures 4-1 and 4-2).

Across different sampling events the blister and pustular mats had over 20 phyla, 30 classes and 60 families that contributed to their microbial community variation (p<0.05). For visualization purposes, only taxa with a level of significance lower than 0.005 were presented in the log-fold-change bar plots (Figures 4-3 and 4-4; Supplemental Figures 4-1 and 4-2).

The blister mats sampled in November 2015 had a significantly greater abundance (p<0.005) of Actinobacteria; Bacteroidetes from the Cytophagaceae family; Firmicutes from the

Clostridia class and Paenibacillaceae family; and the Euryarchaeota Methanococcaceae; when compared from the blister mats sampled in January 2016 which had a significantly greater abundance (p<0.005) of Cyanobacteria of the Prochlorococcaceae family (Figure 4-3A;

Supplemental Figure 4-1A). Differences in the Proteobacteria were also observed when comparing the blister mat microbial communities at these two sampling events; the blister mat sampled in had a significantly higher (p<0.005) relative abundance of Alpha-, Beta- and

Gammaproteobacteria, whereas in January 2016 the same mats had a significantly higher

(p<0.005) relative abundance of Deltaproteobacteria from the family Cystobacteraceae (Figure

4-3A; Supplemental Figure 4-1A). The blister mat sampled in January 2016 had a significantly higher relative abundance (p<0.005) of Cyanobacteria from order Oscillatoriales while in February the mat had a significantly higher abundance (p<0.005) of Firmicutes from

Clostridia and Bacilli classes (Figure 4-3B). Comparison between the two last sampling events

– February and March 2016 – revealed that Gammaproteobacteria, Methanosarcinaceae

(Euryarchaeota) and Gloeobacterales (Cyanobacteria) were significantly more abundant

(p<0.005) in February 2016, whereas Alphaproteobacteria were significantly more abundant

(p<0.005) in March 2016 (Figure 4-3B; Supplemental Figure 4-1C). Bacteroidetes from the

107

Chapter 4

Cytophagaceae family and Actinobacteria from the Nocardiopsaceae family were also significantly more abundant (p<0.005) in the blister mats sampled in February 2016; while

Bacteroidetes from Flavobacteriaceae family and Actinobacteria from Streptomycetaceae family were significantly more abundant in the blister mats sampled in March (Supplemental

Figure 4-1C).

108

Chapter 4

Figure 4-3. Differentially abundant OTUs (agglomerated by species level) in blister mat at class level. (A) November 2015 vs January 2016, (B) January 2016 vs February

2016, (C) February 16 vs March 2016. Only OTUs with a significance level of 0.005 are shown. A less than zero log2-fold-change indicates phyla and classes that were more differently abundant in (A) November 2015, (B) January 2016, (C) February 2016; whereas a greater than zero log2-fold-change indicates phyla and classes that were more differently abundant in (A) January 2016, (B) February 2016, (C) March 2016. Each circle represents one species that was enriched in the paired seasons analyses; not a single

OTU. Different colors represent different phyla.

109

Chapter 4

Seasonal differences in the pustular microbial mat communities were significant when comparing July 2015 versus November 2015; November 2015 versus January 2016; and

January 2016 versus February 2016. Differential abundance analysis revealed the main phyla, classes and families that contributed to this variation (p<0.005). For example, the first paired seasonal analysis showed that in July 2015 the pustular mat microbial communities had a significantly higher abundance (p<0.005) of Epsilonproteobacteria, whereas in November

2015 the pustular mat communities had a significantly higher abundance (p<0.005) of

Alphaproteobacteria and Deltaproteobacteria (Figure 4-4A). The pustular mat communities in

July 2015 also had a significantly higher abundance (p<0.005) of Chloroflexaceae,

Bacterioidetes, the Cyanobacterial family Nostocaceae and the Actinobacterial families

Corynebacteriaceae and Nakamurellaceae; in comparison to the pustular mat communities from November 2015, which had a significantly higher abundance (p<0.005) of

Chrysiogenetes and Firmicutes (Figure 4-4A; Supplemental Figure 4-2A). Differences in the archaeal community between these two seasons was also observed; in July 2015, the

Euryarchaeotal families Thermococcaceae and Archaeoglobaceae were significantly more abundant (p<0.005) in the pustular mat communities, whereas in November 2015

Methanomicrobiales, also from the Euryarchaeota was significantly more abundant (p<0.005) in the pustular mat communities (Figure 4-4A; Supplemental Figure 4-2A). Comparing the sampling events November 2015 and January 2016, the pustular mat communities had significantly higher abundance (p<0.005) of Alphaproteobacteria in November 2015, while in

January 2016, the pustular mat communities had significantly higher abundance (p<0.005) of

Delta- and Gammaproteobacteria, Actinobacteria, Flavobacteria and Clostridida (Figure 4-

4B). The last paired seasonal analysis, January 2016 and February 2016, revealed a significantly higher abundance of Gammaproteobacteria in the pustular mat from January

2016; whereas for February 2016, different phyla and classes contributed to the significant

110

Chapter 4 variation (p<0.005) from the previous sampling event, for example, Alpha- and

Betaproteobacteria, Firmicutes, Bacteroidetes and Crenarchaeota families Thermoproteaceae and Sulfolobaceae (Figure 4-4C; Supplemental Figure 4-2C). As mMDS analysis did not show significant changes among the non-lithifying metagenomes from the different seasons, no log2-fold-change plot was presented.

111

Chapter 4

Figure 4-4. Differentially abundant OTUs (agglomerated by species level) in pustular mat at class level. (A) July 2015 vs November 2015, (B) November 2015 vs January

2016, (C) January 2016 vs February 16. Only OTUs with a significance level of 0.005 are shown. A less than zero log2-fold-change indicates phyla and classes that were more differently abundant in (A) July 2015 (B) November 2015, (C) January 2016; whereas a greater than zero log2-fold-change indicates phyla and classes that were more differently abundant in (A) November 2015, (B) January 2016, (C) February 2016. Each circle represents one species that was enriched in the paired seasons analyses; not a single OTU.

Different colors represent different phyla. 112

Chapter 4

4.3.2 Seasonal variation in functional gene composition of the microbial mats

Principal coordinate analysis (PCO) and metric multidimensional scaling (mMDS) were also used to determine whether the functional gene composition of the three studied mat morphologies were different and whether they have changed throughout the sampling period

(Figure 4-5). Overall, the PCO plot showed that although blister and pustular mat functional gene profiles were not represented in two clearly segregated clusters, the two lithifying mats did cluster separately to the non-lithifying mats (Figure 4-5A). The first two principal coordinate axes explained a total of 51.3% of the variation among the data (PCO1 = 42.8%;

PCO2 = 8.5%) (Figure 4-5A). PERMANOVA confirmed that blister and pustular mat functional gene composition were significantly different from the non-lithifying mat at each sampling event (p<0.05), and between the two lithifying mats the variation was only significant when comparing their metagenomes collected in November 2015 and February and March

2016 (p<0.05). In July 2015 and January 2015, the blister and pustular mat functional gene profiles were not significantly different from one another (p>0.05). The mMDS on distance among centroids showed seasonal changes in the functional gene profile for each microbial mat morphology (Figure 4-5B), however PERMANOVA demonstrated that some of these changes were not significantly. For example, for the non-lithifying loosely cohesive mat the functional gene profiles only seasonally shifted from January 2016 to February to 2016

(p<0.02). The blister mat functional gene composition was significantly affected by season from November 2015 to January 2016, and then to February 2016 (p<0.05). The pustular mat was the mat morphology that varied most throughout the sampling period (p = 0.001), except from February 2016 to March 2016 (p>0.05).

113

Chapter 4

Figure 4-5. Functional gene profiles among metagenomes from microbial mat communities in hypersaline lakes at Rottnest Island. (A) Principal coordinate analysis (PCO) and (B) metric multidimensional scaling (mMDS) on difference among centroids based on Bray-Curtis similarity matrices calculated using functional gene abundance dada from SEED Subsystems annotation (level IV – function). Different colors represent different microbial mat morphologies – blister, pustular, non-lithifying loosely cohesive; and different shapes represent sampling events – July and November 2015, January, February and March 2016.

114

Chapter 4

Pair-wise two-sided White’s non-parametric t-tests between each paired season revealed the main functional genes that contributed to the seasonal variations in the microbial mat communities (p<0.10) that were observed in the mMDS analysis (Figures 4-6 to 4-8;

Supplemental Figure 4-5). In order to better present the extended error bar plots, only key pathways that influence carbonate precipitation in microbialites were selected, for example carbohydrates, cofactors, nitrogen metabolism, photosynthesis, respiration, and sulfur metabolism pathways (Figures 4-6 to 4-8; Supplemental Figure 4-5).

Comparing the blister mats of each paired season it was observed that there were some seasonal differences in the functional gene composition. For example, comparing the blister mats in November 2015 and January 2016, it was observed that the mats in November 2015 had a significantly higher abundance (p<0.10) of respiration associated genes (i.e., cytochrome oxidases and NiFe-hydrogenase) and the ammonia assimilation gene, which is associated with nitrogen metabolism; whereas the blister mats in January 2016 had a significantly higher abundance (p<0.10) of genes associated with photosynthesis (bacteriorhodopsin) and sulfur metabolism (taurine utilization) (Figure 4-6A). January 2016 blister mats had a significantly higher abundance (p<0.10) of photosynthesis and sulfur metabolism associated genes when compared to blister mats in February 2016 (Figure 4-6B).

For the pustular mat, the functional gene composition also showed some seasonal variation. For example, when comparing the pustular mats in July 2016 and November 2016, the first had significantly higher abundance (p<0.10) of nitrogen fixation genes, photosynthesis associated genes; whereas the pustular mats in November 2016 had significantly higher abundance (p<0.10) of genes associated with carbohydrate metabolism and sulfur associated gene DMSP breakdown (Figure 4-7A). When comparing the pustular mats in November 2016 and January 2016 however, genes associated with sulfur metabolism, and photosynthesis were significantly more abundant (p<0.10) in the pustular mat in January 2016; while nitrogen

115

Chapter 4 associated genes were significantly more abundant (p<0.10) in pustular mats in November

2015, when compared to January 2016 pustular mats (Figure 4-7B). Comparing the pustular mats in January 2016 and February 2016, the functional gene associated with photosynthesis, photosystem II, was more abundant in February 2016 mats. The same was observed for functional genes associated with nitrogen metabolism (i.e., nitrogen fixing and nitrate and nitrite ammonification), and CO2 uptake carboxysome genes (carbohydrate metabolism associated) and chlorophyll biosynthesis genes (co-factors associated); which were significantly more abundant (p<0.10) in pustular mats in February 2016, when compared to pustular mats in January 2016 (Figure 4-7C). Sulfur oxidation genes however, were significantly more abundant (p<0.10) in pustular mats from January 2016 than in pustular mats from February 2016 (Figure 4-7C).

The only paired seasons that had significant variation in functional gene composition within non-lithifying loosely cohesive mats were January 2016 compared to February 2016. It was observed that non-lithifying loosely cohesive mat metagenomes from January 2016 had significantly higher abundance (p<0.10) of TCA cycle genes (carbohydrate metabolism associated) and membrane bound hydrogenases (respiration associated); whereas February

2016 pustular mats had significantly higher abundance (p<0.10) of phycobilisome genes

(photosynthesis associated) and nitrate and nitrite ammonification genes (Figure 4.8).

116

Chapter 4

Figure 4-6. Mean proportion of significantly different (± 1 SD; p-value <0.10) functions from SEED Subsystems annotation (level III – third classification in SEED out of four levels) in blister mats. (A) November 2015 vs

January 2016, (B) January 2016 vs February 2016. Only level III functions from the following level I key pathways are shown: black – carbohydrates; blue – cofactors, vitamins, prosthetic groups, pigments; green – nitrogen metabolism; purple – photosynthesis; red – respiration and; orange – sulfur metabolism.

117

Chapter 4

Figure 4-7. Mean proportion of significantly different (± 1 SD; p-value <0.10) functions from SEED Subsystems annotation (level III – third classification in SEED out of four levels) in pustular mats. (A) July 2015 vs November

2015, (B) November 2015 vs January 2016, (B) January 2016 vs February 2016. Only level III functions from the following level I key pathways are shown: black – carbohydrates; blue – cofactors, vitamins, prosthetic groups, pigments; green – nitrogen metabolism; purple – photosynthesis; red – respiration and; orange – sulfur metabolism.

118

Chapter 4

Figure 4-8. Mean proportion of significantly different (± 1 SD; p-value <0.10) functions from SEED Subsystems annotation (level III – third classification in SEED out of four levels) in non-lithifying loosely cohesive mats –

January 2016 vs February 2016. Only level III functions from the following level I key pathways are shown: black

– carbohydrates; blue – cofactors, vitamins, prosthetic groups, pigments; green – nitrogen metabolism; purple – photosynthesis; red – respiration and; orange – sulfur metabolism.

119

Chapter 4

4.4 Discussion

Previous studies have characterized the taxonomic diversity and functional genes of microbial mats, aiming to expand scientific understanding of microbially-induced carbonate precipitation processes and microbialite accretion (Warden et al., 2016; White et al., 2016;

Babilonia et al., 2018). However, most of these studies do not consider the effects of seasonality and subsequent changes in environmental factors, on microbial mat communities. It is known that variation in salinity, nutrients, and water level can cause shifts in the microbial mat communities, and consequently may affect carbonate precipitation (Cabestrero et al., 2018; De

Anda et al., 2018). At Rottnest Island, the hypersaline lakes, in particular Garden Lake, seem to be disturbed by current land and water-use activities (John et al., 2009). In order to determine whether the microbial mats have been affected by seasonality and potential disturbance, the present study performed seasonal characterization of their microbial mat communities and functional genes.

As described previously in this thesis, microbial mats are complex consortia of different microorganisms from different functional groups. These functional groups are considerably dynamic in the mats and seasonal changes can lead to shifts in their richness and abundance

(Preisner et al., 2016). In Chapter 2, analysis of environmental data has shown that seasonal variation in water level, salinity, common anions/cations and other water quality components

(such as nutrients) exists. This variation seems to be primarily due to the rainfall/dilution dominated winter and evaporation/concentration dominated summer, but other local effects

(such as increased groundwater inflow and nutrients) have a role in water quality changes.

Therefore, shifts in the microbial communities of blister, pustular and non-lithifying loosely cohesive mats was expected. As also shown in Chapter 2 the mat communities themselves have

2- a role in driving water quality changes, particularly in the context of sulfur cycling and SO4 .

120

Chapter 4

Overall, the blister and pustular mat microbial communities significantly varied seasonally (p<0.02). These results may suggest that the biofunctional groups in the blister and pustular mat are dynamically responding to seasonal changes and adapting to the new conditions, with no inhibintion of carbonate formation. A recent study has shown that microbial mat communities that undergo small perturbation, such as seasonal changes and variable water levels, can return to their prior state (De Anda et al., 2018). This recovery can be related to the microbial interactions of different functional groups within the mats that could be providing a buffer against the environmental changes and influencing the community assembly and stability (Hunt and Ward, 2015; Konopka et al., 2015; De Anda et al., 2018). Contrastingly,

PERMANOVA did not show significant seasonal variation in the non-lithifying microbial mat community taxonomy (p>0.10). These non-lithifying mats are growing in sites HE and GN, in

Herschel and Garden Lake, respectively, which are the closest sites to the main settlement as well as the golf course on the island (Appendix-II; K. Bischoff, co-authored by J. Mendes-

Monteiro, currently in revision with the Journal Sedimentology). Irrigation and water uses activities might be affecting the mat development, not allowing it to establish a complex and healthy community with different biofunctional groups that would support the mat to adapt to seasonal changes. It was demonstrated in Chapter 2 that the water sample collected from site

GN was the freshest of all lakes; and in Chapter 3, it was shown that the non-lithifying loosely cohesive mat had the highest relative abundance of Cyanobacteria (34%, p<0.05) among all the mats studied (Mendes Monteiro et al., 2019). Cyanobacteria are the first colonizing organisms in microbial mats and this group is responsible for changing the mat microenvironment, hence allowing later colonization of more specialized bacteria with higher and specific environmental requirement (Boomer et al., 2009; Prieto-Barajas et al., 2018). Most likely, the decrease in salinity at Garden lake could enhance the development of Cyanobacteria in those mats and slow down the colonization by other more specific groups that also play

121

Chapter 4 important role in carbonate precipitation. During the present research, it was not possible to collect enough pore water volume seasonally from site HE, because it was undergoing drought during summer sampling events.

According to mMDS on distance among centroids analysis, a few number of different taxa contributed to the seasonal variation in the blister and pustular mats microbial communities. In the blister mat, for example, the microbial mat communities in November

2015 had higher abundance (p<0.005) of Bacteroidetes from the Cytophagaceae family, as well as Alpha-, Beta- and Gammaproteobacteria, and the Euryarchaeota family Methanococcaceae.

In January 2016 however, the microbial communities in the blister mat had a higher abundance

(p<0.005) of Cyanobacteria from Oscillatoriales and Prochlorococcaceae, and

Deltaproteobacteria from the family Cystobacteraceae. This shift suggests that due to warmer conditions and increased solar illumination, the growth of photosynthetic organisms such as

Cyanobacteria was enhanced in the blister mat, which might have led to an oxygen-rich microenvironment (Cabestrero et al., 2018). With the exception of the microaerophile

Anaeromyxobacter, all Cystobacteraceae are strict aerobes and spores producers (dos Santos et al., 2014). The development of the Cyanobacterial community in the blister mat in January

2016, and the new oxygen-rich microenvironment might have created suitable conditions for the growth of Cystobacteraceae.

In February 2016, a different Euryachaeota family, Methanosarcinaceae, was more abundant (p<0.005) in blister mat microbial communities, when compared to the blister mat from November 2015, in which the Euryarchaeota family Methanococcaceae was more abundant (p<0.005). Both families are methanogens (Oren, 2014a; Oren, 2014b), suggesting that methanogenesis is an important metabolism in the blister mat. However, they are capable of using different substrates during the methanogenesis process. Methanococcaceae are hydrogenotrophic methanogens, which use H2 for the reduction of CO2 or formate onto

122

Chapter 4 methane (Fenchel et al., 2012; Oren, 2014a), whereas most of the Methanosarcinaceae are acetoclastic methanogens, which use acetate as a substrate; or methylotrophic methanogens, which use methanol or methylamines as a substrate, with both types of methanogenesis reducing the substrate to CO2 and CH4 (Fenchel et al., 2012; Oren, 2014b). During hydrogenotrophic methanogenesis, CO2 is consumed, which increases local alkalinity and induces carbonate precipitation (Visscher and Stolz, 2005). Similarly to hydrogenotrophic methanogens, some groups of sulfur reducing bacteria (SRB) are able to oxidise H2 while fixing

CO2 (Vischer and Stolz, 2005). The presence of both hydrogenotrophic methanogens and SRB in microbial mats could indicate competition for H2, however both processes would still result in carbonate precipitation. To date, it is not clear how acetoclastic and methylotrophic methanogenesis affect the carbonate cycling in microbial mats, but it is thought that it could have different effects on the carbonate budget (Visscher and Stolz, 2005).

The Actinobacteria family Nocardiopsaceae and the Bacterioidetes family

Cytophagaceae were also more abundant (p<0.05) in February 2016. In March 2016, other

Actinobacteria and Bacteroidetes families were found to be significantly more abundant

(p<0.005) – Streptomycetaceae and Flavobacteriaceae, respectively. Both Actinobacteria and

Bacteroidetes are in general able to decompose different types of long chain organic molecules, which may liberate specific ions into the mat microenvironment such as Ca+ which would facilitate carbonate precipitation (Dupraz and Visscher, 2005; Chen-Harris et al., 2013; Lewin et al., 2016). Therefore, it is not clear whether shifts in the Actinobacteria and Bacteroidetes at family level in the blister mat would affect the capacity of this mat to develop or promote carbonate precipitation.

In the pustular mat, the family Chloroflexaceae was more abundant (p<0.005) in the

July 2015 pustular mat. The bacteria in this family are anoxygenic phototrophs and are commonly found in photosynthetic microbial mats (Baumgartner et al., 2009; Ley et al., 2006).

123

Chapter 4

Members from this group usually have a very versatile metabolism, being able to grow under anaerobic conditions, performing photoheterotrophic process in the light, and under aerobic conditions performing chemoheterotrophic process in the dark (Nierychlo et al., 2018; Kawai et al., 2019). For example, they are able to perform anoxygenic photoautotrophic process oxidizing H2S or fermentation of carbohydrates, and during both processes carbonate can precipitate (Visscher and Stolz, 2005; Klatt et al., 2007; Farias et al., 2017). The presence of

Chloroflexaceae in the pustular mat microbial communities, particularly in July 2015, indicates a carbonate precipitation potential in that sampling event.

In November 2015, Chrysiogenetes were significantly more abundant (p<0.005) in the pustular mat microbial communities, compared to the other sampling events. This phylum has currently only one species that exhibits anaerobic respiration and uses arsenate as the electron acceptor (Garrity et al, 2001). Arsenate reduction has not been vastly studied but it is believed that this process has played an important ecological role in microbial mats and microbialites before the oxygenation of the atmosphere (Visscher et al., 2014). Recent studies showed evidence of the arsenic cycle playing a potential role in carbonate precipitation in microbial mats of the Atacama Desert in northern Chile (Visscher et al., 2014). Another study of stromatolite mats from Socompa Lake (Andean Plateau), which has high arsenic concentrations, demonstrated that due this environmental stress, the microbial communities exhibited arsenic resistance mechanisms mainly based on arsenic reduction (Kurth et al., 2017).

In chapter 2, it was demonstrated that pore water collected from pustular mat had high concentration of arsenic, especially in November 2015 at Lake Baghdad (data not shown). The presence of Chrysiogenetes in the pustular mat microbial communities in November 2015 suggests that arsenic cycle might play a role in microbialite formation within the pustular mat.

The abundance of Archaea also showed seasonal variation in the pustular mat. In July

2015, the Euryarchaeota families Thermococcaceae and Archaeoglobaceae were more

124

Chapter 4 abundant (p<0.005) in the pustular mat microbial communities; whereas the Euryachaeota order Methanomicrobiales were more abundant (p<0.005) in November 2015 and the

Crenarchaeota families Thermoproteaceae and Sulfolobaceae were more abundant (p<0.005) in the pustular mat in February 2016. The Thermoccocaceae family is known to be able to reduce elemental sulfur to hydrogen sulfide (Poli et al., 2017), which can further be oxidized by aerobic or anaerobic sulfide oxidizing bacteria that promote dissolution and precipitation of carbonate, respectively (Visscher and Stolz, 2005). The Archaeoglobaceae family comprises organisms able to reduce sulfate, which is a process known to promote carbonate precipitation

(Zhu and Dittrich, 2016; Poli et al., 2017). Although the presence of Thermococcaceae could indicate that carbonate dissolution is taking place, carbonate precipitation could be also occurring in the pustular mat in July 2015 via sulfate reduction processes performed by

Archaeoglobaceae. Methanomicrobiales are able to perform hydrogenotrophic methanogenesis

(Browne et al., 2017), which as stated previously can induce carbonate precipitation. The higher abundance of Methanomicrobiales in the pustular mats in November 2015 suggests a shift in the dominant archaeal group that is actively accreting carbonate; however, it was not identified what environmental conditions could have enhanced the growth of this archaeal group. The same applies to the pustular mat community in February 2016, when the

Crenarchaeota families Thermoproteaceae and Sulfolobaceae, also involved in sulfur cycle

(Albers and Siebers, 2014; Itoh, 2014), were more abundant (p<0.005). In other words, although there is a clear shift in Archaeal community in the pustular mat, it does not seem to affect the potential of this mat with respect to carbonate accretion.

Seasonal variation was also observed in the functional gene composition within the microbial mats studied. The blister mat functional gene composition in November 2015, for example, was significantly abundant (p<0.10) in respiration associated genes for NiFe- hydrogenases. This enzyme is responsible for the activation of H2, during the process of

125

Chapter 4 methanogenesis, in which CO2 is reduced to CH4 (Rudolf et al., 2010). The Methanococcaceae family was one of the taxa that were abundant in November 2016 blister mat microbial communities. The presence of genes for NiFe-hydorgenases is another strong indicator that methanogenesis is taking place in that sampling event. In January 2016, the blister mats had a higher abundance (p<0.10) of bacteriorhodopsin genes associated with photosynthesis.

Bacteriorhodopsin is the major photosynthetic membrane protein that works as light-driven proton pump used by Archaea, particularly halobacteria, to harvest energy from light and transport protons out of the cell (Nango et al., 2016). Bacteriorhodopsin featured mainly in blister mats from January coupled with the fact that cyanobacteria are also abundant in mats from this same sampling event, suggests that photosynthesis metabolisms is an important process for blister mat in this season.

For the pustular mat, the functional gene composition also showed some seasonal variation. However, photosynthesis associated genes seemed to be present throughout the entire sampling period, except November 2015. The pustular mat is found at the foreshore of the lakes and is yearly exposed to sunlight, not being covered by lake water most of the time. Therefore, it was expected that photosynthesis would be abundantly expressed in this mat. Genes associated with sulfur metabolism (i.e., sulfur oxidation) were significantly more abundant

(p<0.10) in the pustular mat in January 2016. It is known that aerobic sulfide oxidation results in carbonate dissolution (Visscher and Stolz, 2004; Baumgartner et al., 2006). However, when the oxygen concentration is low in the mat microenvironment, it leads to incomplete sulfide oxidation, which if coupled to carbon fixation, stimulates carbonate precipitation instead of carbonate dissolution (Visscher and Stolz, 2005, Baumgartner et al., 2006). It is not clear what type of sulfide oxidation was taking place in the pustular mat in January 2016, but the potentially high photosynthetic activity suggest that may be the aerobic type.

126

Chapter 4

For the non-lithifying loosely cohesive mats, phycobilisome genes (photosynthesis associated) were more abundant in February 2016. Phycobilisomes are the light-harvesting systems in Cyanobacteria (Zhang et al., 2017). As described earlier in this chapter and in

Chapter 3, Cyanobacteria are abundant in the non-lithifying mat, in comparison to blister and pustular mat. Therefore the presence of this gene in abundance is not surprising, although it does suggests that photosynthesis might be more actively taking place in February 2016 than in the previous and following sampling periods, January and March 2016. Other functional genes that are important for carbonate precipitation did not contribute to seasonal variation in functional gene composition in non-lithifying mats, which is another indicative that these microbial mats might not comprise great populations of other more specialized bacteria with higher and specific environmental requirement.

4.5 Conclusion

The present study have assessed for the first time implications of seasonal variation onto lithifying and non-lithifying microbial mats from hypersaline lakes at Rottnest Island. The results demonstrate that the two different types of mats responded differently to seasonality.

The taxonomic diversity within the lithifying mats – blister and pustular – showed significant variation across seasons, however for the non-lithifying mats, seasonality did not imply in significant variation of the microbial community composition. These results suggest that while lithifying mats are able to shift their microbial community in response to seasonality, the non- lithifying mats are lacking the same capacity. The reason for that seems to be salinity decrease due to irrigation activities near the lakes where non-lithifying mats are growing. The water freshening in those lakes seems to be enhancing the development of Cyanobacteria, not allowing other biofunctional groups to colonize the non-lithifying mats. The seasonal variation observed in the microbial mat communities, as well as in the functional gene composition of lithifying mats does not seem to be inhibiting microbialite formation because across the

127

Chapter 4 different seasons it was detected biofunctional groups and genetic information that are involved in important metabolic processes that leads to carbonate accretion, such as methanogenesis, photosynthesis and sufur metabolism. Further research is needed to identify the how the different biofunctional groups present in the microbial mats in different seasons are affecting carbonate accretion and whether it is affecting the microbialites mineralogy. In addition, ongoing monitoring of the non-lithifying mats are necessary to provide even further insights into the effects of salinity variation in those mats. This would support the Rottnest Island

Authority to better manage the microbial mats ecosystem and enhance our understanding of the influence of microbial community into microbialite formation processes.

128

Chapter 4

Supplemental material

Supplemental Figure 4-1. Differentially abundant OTUs (agglomerated by species level) in blister mat at family level. (A) November 2015 vs January 2016, (B) January 2016 vs February 2016, (C) February 16 vs March 2016. Only OTUs with a significance level of 0.005 are shown. A less than zero log2-fold-change indicates phyla and families that were more differently abundant in (A) November 2015, (B) January 2016, (C) February 2016; whereas a greater than zero log2-fold-change indicates phyla and families that were more differently abundant in (A) January 2016, (B) February 2016, (C) March 2016. Each circle represents one species that was enriched in the paired seasons analyses; not a single OTU. Different colors represent different phyla.

129

Chapter 4

Supplemental Figure 4-2. Differentially abundant OTUs (agglomerated by species level) in pustular mat at family level. (A) July 2015 vs November 2015, (B) November

2015 vs January 2016, (C) January 2016 vs February 16. Only OTUs with a significance level of 0.005 are shown. A less than zero log2-fold-change indicates phyla and families that were more differently abundant in (A) July 2015 (B) November 2015, (C) January 2016; whereas a greater than zero log2-fold-change indicates phyla and families that were more differently abundant in (A) November 2015, (B) January 2016, (C) February 2016. Each circle represents one species that was enriched in the paired seasons analyses; not a single OTU. Different colors represent different phyla.

130

Chapter 4

Supplemental Figure 4-3. Differentially abundant OTUs (agglomerated by species level) between lithifying and non-lithifying microbial mats at class level. (A) July 2015, (B) November 2015, (C) January 2016, (D) February

16 (E) March 2016. Only OTUs with a significance level of 0.005 are shown. A less than zero log2-fold-change indicates phyla and classes that were more differently abundant in lithifying mats; whereas a greater than zero log2-fold-change indicates phyla and classes that were more differently abundant in non-lithifying mats. Each circle represents one species that was enriched in the paired seasons analyses; not a single OTU. Different colors represent different phyla.

131

Chapter 4

Supplemental Figure 4-4. Differentially abundant OTUs (agglomerated by species level) between lithifying and non-lithifying microbial mats at family level. (A) July 2015, (B) November 2015, (C) January 2016, (D) February

16 (E) March 2016. Only OTUs with a significance level of 0.005 are shown. A less than zero log2-fold-change indicates phyla and families that were more differently abundant in lithifying mats; whereas a greater than zero log2-fold-change indicates phyla and families that were more differently abundant in non-lithifying mats. Each circle represents one species that was enriched in the paired seasons analyses; not a single OTU. Different colors represent different phyla.

132

Chapter 4

Supplemental Figure 4-5. Mean proportion of significantly different (± 1 SD; p-value <0.005) functions from

SEED Subsystems annotation (level III – third classification in SEED out of four levels) in lithifying mats vs non- lithifying, at each seasonal event. (A) July 2015 (B) November 2015, (C) January 2016, (D) February 2016, (E)

March 2016. Only level III functions from the following level I key pathways are shown: black – carbohydrates; blue – cofactors, vitamins, prosthetic groups, pigments; green – nitrogen metabolism; purple – photosynthesis; red – respiration and; orange – sulfur metabolism.

133

Chapter 5

Chapter 5

Discussion and Conclusion

5.1 Discussion

Modern microbial mats are considered an important resource for understanding the physiological processes that shaped planetary biology (Chilton et al., 2012). Moreover, they are an ideal setting to study microbially-induced precipitation and microbialite formation

(Inskeep et al., 2013; Cabestrero and Sanz-Montero, 2018; Prieto-Barajas et al., 2018). Several studies have investigated the link between microbial mat communities and carbonate precipitation (Dupraz et al., 2009; Pace et al., 2016; Wilmeth et al., 2018), however there is still not an explicit understanding of how the communities accrete carbonate and what environmental conditions limit microbialite formation and microbial mat persistence. This thesis significantly contributes to advancing scientific knowledge on the dynamic nature of microbial mats communities and their role in microbialite accretion. In addition, it provides relevant information regarding interactions between microbial communities and environmental factors in the hypersaline lakes at Rottnest Island, which is fundamental for the conservation of these sensitive ecosystems by the Rottnest Island Authority.

Limited research has been conducted on the microbial mats hydroecology at Rottnest

Island and little is known about their responses to environmental disturbance and/or seasonality. It is known however, that environmental changes, such as altered hydroperiod, altered groundwater regime, greater nutrient loads, acidification, changes in turbidity and salinity may affect the microbial mats by causing shifts in the dominant microbe communities

(Timms, 2005). To explore the process in which microbial mats influence carbonate

134

Chapter 5 precipitation and at the same time provide insights into the microbial ecology of the mats in the hypersaline lakes at Rottnest Island, the microbial communities of lithifying and non- lithifying microbial mats were compared in order to identify differences in taxa diversity and functional gene composition between the two mat types. In addition, the evolution and adaption of the microbial mats in response to seasonal and environmental factors changes were also assessed by analysing the communities taxonomic and functional variations during winter and summer along with lake and pore water chemistry evaluation.

Across all the hypersaline lakes analysed, Garden and Herschel Lakes appear to be experiencing the most significant environmental disturbance, for example decreasing salinity, due to increased groundwater inflow from increased irrigation of a neighbouring golf course

(Rottnest Island Authority, 2014). Field observations during the current research revealed that these two lakes are the only ecosystems harbouring non-lithifying loosely cohesive mats, with decreasing salinity – 140 g/L in March 2015 (Gillen, 2015, unpub.) and 69 g/L in March 2015

(Chapter 2) – influencing mat development and hydroecology. Analysis of surface water levels has shown that Garden Lake (the lake most impacted by irrigation activities on the island) is experiencing a rising water level trend, which reaffirms that irrigation activities are increasing groundwater discharge into this lake. Water analysis showed that Garden Lake had alkalinity results over the minimum required for microbialite formation (75.03 mg/L, suggested by Zeyen et al., 2017), and appropriate local concentrations of bivalent cations (e.g., Ca2+, Mg2+) in addition to active sulfur reduction, however lithifying microbial mats and microbialites are absent in this lake. Most likely decreases in salinity is affecting the microbial communities in

Garden Lake mats, not allowing the development of a diverse community with the different biofunctional groups that are necessary for microbialite formation. Previous studies have shown that changes in salinity can alter the microbial mats dominant taxa and their functional genes composition in San Salvador Island (Bahamas) and Lake Clifton (Western Australia)

135

Chapter 5

(Smith et al., 2010; Gleeson et al., 2016; Preisner et al., 2016); which in turn can affect the way microbial mats precipitate carbonates and even the mineralogy of the microbialites formed

(Cabestrero et al., 2018). Although Herschel Lake did not show a similar trend in rising water level, and lithifying mats are present in this lake, the non-lithifying loosely cohesive mat was also observed growing on the eastern side of Herschel, which is located in the vicinity of

Garden Lake. Therefore, irrigation activities could also be the reason for the development of non-lithifying loosely cohesive mat in the eastern side of Herschel Lake.

In Rottnest Island, differences in salinity among the lakes appears to be a key factor causing differences in microbial community taxa and functional genes composition between lithifying and non-lithifying microbial mats. Comparisons between lithifying and non- lithifying mats revealed that the lithifying mats had a higher abundance of Alpha- Delta- and

Gammaproteobacteria, Bacteroidetes, and Spirochaetes, which are all known to have potential roles in microbialite formation (Baumgartner et al., 2006; Dupraz et al., 2009; Dupraz et al.,

2011; Chen-Harris et al., 2013). In addition, although it has been shown in Chapter 3 that the overall relative abundance of Cyanobacteria in the non-lithifying mat communities collected in February 2016, was higher than when compared to the lithifying mat communities, some

Cyanobacterial families, in particular from Chroococcales order, were significantly more abundant (p<0.05) in blister and pustular mat when compared to non-lithifying mats

(Supplemental Figure 3-5 and 3-10). Bischoff et al. (2020), using light microscopy and scanning electron microscopy and energy dispersive spectroscopy (SEM-EDS), have demonstrated that lithifying mats from Seprentine Lake, for example, are also enriched with

Chroococcales and suggest that this Cyanobacterial order is most likely inducing mineral precipitation, more specifically Mg-silicates formation. It is known that Mg-silicate minerals require a minimum pH of 8.70 (Tosca and Masterson, 2014), which is higher than all pH measumrents for all lakes (Chapter 2). It is also know that Cyanobacteria can increase local pH

136

Chapter 5 due to carbon dioxide uptake during photosynthetic activities (Visscher and Stolz, 2005;

Baumgartner et al., 2006) further enhancing the likelihood of microbialites formation. In the same study, Bischoff et al. (2020) has suggested that Chroococcales are increasing local pH to values higher than 8.70 via photosynthesis, hence facilitating the precipitation of Mg-silicates

(Bischoff et al., 2020).

Alpha-, Delta- and Gammaproteobacterial taxa include sufate reducing (SRB) and sulfide oxidizing bacteria (SOB), which are common members of microbial mats (Warden et al., 2016; Babilonia et al., 2018). Within microbial mats, SRB and SOB stablish a syntrophic relationship, wherein the SRB anaerobically reduce sulfate to sulphide and the SOB oxidize the sulfide produced by the SRB. During this process local alkalinity within the mat increases due to elevated concentration of carbonate ions in the microbial mat microenvironment

(Visscher and Stolz, 2005; Baumgartner et al., 2006). Most Bacteroidetes and Spirochaetes, two other highly abundant groups in the lithifying mats, are organoheterotrophic able degrade high molecular weight compounds, for example EPS, realising bivalent ions, such as Ca2+,

2- which are then free to bind onto CO3 and form calcium carbonate (Cottrell and Kirchman,

2000; O’Sullivan et al., 2006; Chen-Harris et al., 2013, Wong et al., 2015). All these different biofunctional groups, that were, according to the differential abundance analysis, more abundant in the lithifying mats than in the non-lithifying mats, are potentially contributing to the process of microbialite formation in the hypersaline lakes at Rottnest Island. In addition, the lower abundance of these important taxa that influence microbialite formation in the non- lithifying mats may be a consequence of the lower salinity trend experienced in Garden Lake.

In terms of functional gene composition however, important genes related to the same metabolic pathways, sulfate reduction and photosynthesis, were more abundant in non- lithifying mats when compared to lithifying mats. These results suggest that although non- lithifying mats are not actively forming microbialites and do not harbour a diverse microbial

137

Chapter 5 community with different biofunctional groups in abundance as described for the lithifying mats, they do have the functional capacity required for carbonate precipitation and microbialite formation.

In the present study, metagenomic analysis also revealed that lithifying and non- lithifying mats were differently affected by seasonality. The lithifying mats showed significant variation across seasons, which was not observed for the non-lithifying mats. These results suggest that the lithifying mats are dynamically responding to seasonal changes, whereas the non-lithifying mats are not. The interactions among different metabolic groups of microorganisms within the lithifying mats provide a buffering system against seasonal changes and support the microbial mat community stability (Hunt and Ward, 2015; Konopka et al.,

2015; De Anda et al., 2018). In the non-lithifying mats the same is not happening most likely because irrigation activities are impacting the growth of the non-lithifying mats, hence they are not able to develop a biofunctionally diverse mat, which in turn does not allow the mat to adequately respond to seasonal changes and possibly drives its composition to a non-lithifying mat type.

Although an overall seasonal variation in the taxonomic diversity was observed in the lithifying mats such changes might not result in carbonate precipitation ceasing in the mats.

For example, the archaeal community showed significant seasonal variation, from July 2015 to November 2015. In July 2015, the Euryarchaeota family Achaeoglobaceae was more abundant in pustular mat, whereas in November 2015, the Euryarchaeota order

Methanomicrobiales was more abundant. Although these two groups are from different taxonomic groups, both are able to induce carbonate precipitation; the first one via sulfate reducing process, and the second via hydrogenotrophic methanogenesis (Visscher and Stolz;

2005; Zhu and Dittrich, 2016; Browne et al., 2017; Poli et al., 2017). A similar scenario was observed in blister mat, where Actinobacteria and Bacterioidetes communities also showed

138

Chapter 5 significant seasonal variation, from February 2016 to March 2016. In February the

Actinobacteria family Nocardiopsaceae and the Bacterioidetes family Cytophagaceae were also more abundant, whereas in March 2016, other Actinobacteria and Bacteroidetes families were found to be significantly more abundant – Streptomycetaceae and Flavobacteriaceae, respectively. Both Actinobacteria and Bacteroidetes are in general able to decompose complex organic molecules, and during this process bivalent ions (e.g., Ca2+) that were previously attached to the organic matrix are released into the mat microenvironment, which can facilitate calcium carbonate precipitation (Dupraz and Visscher, 2005; Chen-Harris et al., 2013; Lewin et al., 2016). Therefore, shifts in the Actinobacteria and Bacteroidetes at family level in the blister mat would probably not affect the capacity of this mat to accrete carbonate.

Through the seasonal analysis of the microbial mat communities it was also possible to identify the dominant biofunctional groups and metabolic pathways within the different lithifying mats that are potentially affecting microbialite formation. For example, in November

2015 (early-summer) and February 2016 (mid-summer) methanogenic archaea were dominant in the blister mat microbial communities. In addition, genes of the enzyme NiFe-hydrogenase, which is involved in methanogenesis process, were also dominant in the functional gene composition of the blister mat, suggesting that methanogenesis is an important metabolism that might be directly influencing carbonate accretion in the blister mat. For the pustular mat, which grows at the foreshore of the lakes, photosynthesis associated genes were dominant at most of the sampling events, suggesting that due to constant sunlight exposure, photosynthesis is highly active in this mats and could be influencing microbialite formation in this mat. The effect of photosynthesis may be direct through forming carbonate via metabolic activity or even indirect through production of EPS, increasing the buffering capacity of the mat to environmental change. The information about the dominant communities and functional genes in these mats

139

Chapter 5 can provide a baseline in many cases to support future work regarding the process of microbialite formation and microbial mats persistence in the lakes at Rottnest Island.

5.2 Conclusion

Overall the findings presented in this thesis enhance our knowledge regarding the microbial ecology and hydrology of the microbial mats in the hypersaline lakes at Rottnest

Island, which have never been thoroughly assessed before. The results showed a strong indication that the microbial mats in Garden Lake are being affected by a decrease in water salinity likely where irrigation activities occur adjacent to the lakes, and therefore are not accreting carbonate. However, it was demonstrated that the non-lithifying mats do harbour taxonomic diversity and functional genes that have the potential to form microbialites. With respect to lithifying mats, photosynthesis, sulfur cycling and methanogenesis are likely to be the metabolic processes that are promoting carbonate and other minerals precipitation and seasonal analysis revealed that this process could be happening throughout different seasons.

Further analyses are required, to refine our understanding of the genes that are being expressed and the chemical metabolic process occurring in different seasons. Ongoing monitoring of both lithifying and non-lithifying microbial mats will provide even further insights into the microbial ecology of these mats and their response to seasonal and environmental changes, which will support the Rottnest Island Authority to better manage this sensitive ecosystem and expand scientific knowledge regarding microbialite accretion.

140

References

References

Albers, S. V., and B. Siebers. 2014. The family Sulfolobaceae, In: Rosenberg E., DeLong E.

F., Lory S., Stackebrant E., Thompson F. (eds.) The Prokaryotes. Springer, Berlin,

Heidelberg, pp. 323–346. doi:10.1007/978-3-642-38954-2_329

Anderson, M. J., Gorley, R. N., and Clarke, K. R. 2008. PERMANOVA+ for PRIMER: Guide

to Software and Statistical Methods, In: Plymouth, UK.

Awramik, S. M. 1976. Chapter 6.3 Gunflint stromatolites: microfossil distribution in relation

to stromatolite morphology. Dev. Sedimentol. 20: 311–320. doi:10.1016/S0070-

4571(08)71141-5

Babilonia, J., A. Conesa, G. Casaburi, C. Pereira, A. S. Louyakis, R. P. Reid, and J. S. Foster.

2018. Comparative metagenomics provides insight into the ecosystem functioning of the

Shark Bay Stromatolites, Western Australia. Front. Microbiol. 9: 1–16.

doi:10.3389/fmicb.2018.01359

Basak, N., and D. Das. 2007. The prospect of purple non-sulfur (PNS) photosynthetic bacteria

for hydrogen production: The present state of the art. World J. Microbiol. Biotechnol. 23:

31–42. doi:10.1007/s11274-006-9190-9

Baumgartner, L. K., J. R. Spear, D. H. Buckley, N. R. Pace, R. P. Reid, C. Dupraz, and P. T.

Visscher. 2009. Microbial diversity in modern marine stromatolites, Highborne Cay,

Bahamas. Environ. Microbiol. 11: 2710–2719. doi:10.1111/j.1462-2920.2009.01998.x

Baumgartner, L. K., R. P. Reid, C. Dupraz, A. W. Decho, D. H. Buckley, J. R. Spear, K. M.

Przekop, and P. T. Visscher. 2006. Sulfate reducing bacteria in microbial mats: Changing

paradigms, new discoveries. Sediment. Geol. 185: 131–145.

doi:10.1016/j.sedgeo.2005.12.008

Bischoff, K., E. Sirantoine, M. E. J. Wilson, A. D. George, J. Mendes Monteiro, and M.

Saunders. 2020. Spherulitic microbialites from modern hypersaline lakes, Rottnest Island,

141

References

Western Australia. Geobiology 00:1–17. doi:10.1111/gbi.12400

Bolhuis, H., M. S. Cretoiu, and L. J. Stal. 2014. Molecular ecology of microbial mats. FEMS

Microbiol. Ecol. 90: 335–350. doi:10.1111/1574-6941.12408

Bonilla-Rosso, G., M. Peimbert, L. D. Alcaraz, I. Hernández, L. E. Eguiarte, G. Olmedo-

Alvarez, and V. Souza. 2012. Comparative Metagenomics of Two Microbial Mats at

Cuatro Ciénegas Basin II: Community Structure and Composition in Oligotrophic

Environments. Astrobiology 12: 659–673. doi:10.1089/ast.2011.0724

Boomer, S. M., K. L. Noll, G. G. Geesey, and B. E. Dutton. 2009. Formation of multilayered

photosynthetic biofilms in an alkaline thermal spring in Yellowstone national Park,

Wyoming. Appl. Environ. Microbiol. 75(8): 2464–2475. doi:10.1128/AEM.01802-08

Braissant, O., A. W. Decho, C. Dupraz, C. Glunk, K. M. Przekop, and P. T. Visscher. 2007.

Exopolymeric substances of sulfate-reducing bacteria: Interactions with calcium at

alkaline pH and implication for formation of carbonate minerals. Geobiology 5: 401–411.

doi:10.1111/j.1472-4669.2007.00117.x

Breitbart, M., A. Hoare, A. Nitti, and others. 2009. Metagenomic and stable isotopic analyses

of modern freshwater microbialites in Cuatro Ciénegas, Mexico. Environ. Microbiol. 11:

16–34 doi:10.1111/j.1462-2920.2008.01725.x

Browne, P., H. Tamaki, N. Kyrpides, and others. 2017. Genomic composition and dynamics

among Methanomicrobiales predict adaptation to contrasting environments. ISME J.

11(1): 87–99. doi:10.1038/ismej.2016.104

Bundeleva, I. A., L. S. Shirokova, P. Bénézeth, O. S. Pokrovsky, E. I. Kompantseva, and S.

Balor. 2012. Calcium carbonate precipitation by anoxygenic phototrophic bacteria. Chem.

Geol. 291: 116–131. doi:10.1016/j.chemgeo.2011.10.003

Burne, R. V., and L. S. Moore. 1987. Microbialites: Organosedimentary Deposits of Benthic

Microbial Communities. Palaios 2: 241. doi:10.2307/3514674

142

References

Bushnell, B. 2017. BBTools software package. URL http//sourceforge. net/projects/bbmap.

Cabestrero, M. E. Sanz-Montero, L. Arregui, S. Serrano, and P. T. Visscher. 2018. Seasonal

Variability of Mineral Formation in Microbial Mats Subjected to Drying and Wetting

Cycles in Alkaline and Hypersaline Sedimentary Environments. Aquat. Geochemistry.

24(1): 79–105. doi:10.1007/s10498-018-9333-2

Cabestrero, Ó., and M. E. Sanz-Montero. 2018. Brine evolution in two inland evaporative

environments: influence of microbial mats in mineral precipitation. J. Paleolimnol. 59(2):

139–157. doi:10.1007/s10933-016-9908-0

Canfield, D. E., and D. J. Des Marais. 1991. Aerobic sulfate reduction in microbial mats.

Science 251: 1471–1473. doi:10.1126/science.11538266

Cantrell, S. A., and L. Duval-Pérez. 2013. Microbial mats: An ecological niche for fungi. Front.

Microbiol. Front. Microbiol. 3(424): 1–7. doi:10.3389/fmicb.2012.00424.

Casaburi, G., A. A. Duscher, R. P. Reid, and J. S. Foster. 2016. Characterization of the

stromatolite microbiome from Little Darby Island, The Bahamas using predictive and

whole shotgun metagenomic analysis. Environ. Microbiol. 18: 1452–1469.

doi:10.1111/1462-2920.13094

Chen-Harris, H., M. K. Borucki, C. Torres, T. R. Slezak, and J. E. Allen. 2013. Ultra-deep

mutant spectrum profiling: Improving sequencing accuracy using overlapping read pairs.

BMC Genomics 14: 96. doi:10.1186/1471-2164-14-96

Chilton, A., B. Neilan, and B. Burns. 2012. Report on Characterisation of freshwater

thrombolite microbial communities from Lake Richmond via barcoded pyrosequencing:

Prepared for the Department of Environment and Conservation, The University of New

South Wales.

Clarke, K. R., and Gorley, R. N. 2015. Primer v7: User Manual/Tutorial, In: Plymouth, UK.

Cottrell, M. T., and D. L. Kirchman. 2000. Natural assemblages of marine proteobacteria and

143

References

members of the Cytophaga-flavobacter cluster consuming low- and high-molecular-

weight dissolved organic matter. Appl. Environ. Microbiol. 66: 1692–1697.

doi:10.1128/AEM.66.4.1692-1697.2000

Cox, M. P., D. A. Peterson, and P. J. Biggs. 2010. SolexaQA: At-a-glance quality assessment

of Illumina second-generation sequencing data. BMC Bioinformatics 11: 485.

doi:10.1186/1471-2105-11-485

Crosbie R.S., D. Morrow, R. G. Cresswell, F. W. Leaney, S. Lamontagne, and M. Lefournour.

2012. New insights into the chemical and isotopic composition of rainfall across Australia.

CSIRO. 76p. doi:10.5072/83/5849a0d6c685a

D’Agostino, P. M., J. N. Woodhouse, H. T. Liew, L. Sehnal, R. Pickford, H. L. Wong, B. P.

Burns, and B. A. Neilan. 2019. Bioinformatic, phylogenetic and chemical analysis of the

UV-absorbing compounds scytonemin and mycosporine-like amino acids from the

microbial mat communities of Shark Bay, Australia. Environ. Microbiol. 21(2): 702–715.

doi:10.1111/1462-2920.14517

De Anda, V., I. Zapata-Peñasco, J. Blaz, A. C. Poot-Hernández, B. Contreras-Moreira, M.

González-Laffitte, N. Gámez-Tamariz, M. Hernández-Rosales, L. E. Eguiarte, and V.

Souza. 2018. Understanding the mechanisms behind the response to environmental

perturbation in microbial mats: A metagenomic-network based approach. Front.

Microbiol. 9: 2606. doi:10.3389/fmicb.2018.02606

Des Marais, D. J. 2003. Biogeochemistry of hypersaline microbial mats illustrates the

dynamics of modern microbial ecosystems and the early evolution of the biosphere.

Biological Bulletin. 204(2): 160–167. doi:10.2307/1543552 dos Santos, D. F. K., C. M. Kyaw, T. A. De Campos, R. N. G. Miller, E. F. Noronha, M. M.

D. C. Bustamante, and R. Kruger. 2013. The family Cystobacteraceae, In: Rosenberg E.,

DeLong E. F., Lory S., Stackebrant E., Thompson F. (eds.) The Prokaryotes. Springer,

144

References

Berlin, Heidelberg, pp. 19–40. doi:10.1007/978-3-642-39044-9_304

Dupraz, C., A. Fowler, C. Tobias, and P.T. Visscher. 2013. Stromatolitic knobs in Storr’s Lake

(San Salvador, Bahamas): A model system for formation and alteration of laminae.

Geobiology. 11(6): 527–548. doi:10.1111/gbi.12063

Dupraz, C., and P. T. Visscher. 2005. Microbial lithification in marine stromatolites and

hypersaline mats. Trends Microbiol. 13: 429–438. doi:10.1016/j.tim.2005.07.008

Dupraz, C., R. P. Reid, and P. T. Visscher. 2011. Microbialites, modern. In: Reitner J., Theil

V. (eds) Encyclopedia of Geobiology. Encycl. Earth Sci. Ser. Springer, Dordrecht, pp.

617–635. doi:10.1007/978-1-4020-9212-1_195

Dupraz, C., R. P. Reid, O. Braissant, A. W. Decho, R. S. Norman, and P. T. Visscher. 2009.

Processes of carbonate precipitation in modern microbial mats. Earth-Science Rev. 96:

141–162. doi:10.1016/j.earscirev.2008.10.005

Edgcomb, V. P., J. M. Bernhard, R. E. Summons, W. Orsi, D. Beaudoin, and P. T. Visscher.

2014. Active eukaryotes in microbialites from Highborne Cay, Bahamas, and Hamelin

Pool (Shark Bay), Australia. ISME J. 8(2): 418-429. doi:10.1038/ismej.2013.130

Farías, M. E., M. C. Rasuk, K. L. Gallagher, M. Contreras, D. Kurth, A. B. Fernandez, D. Poire,

F. Novoa, and P. T. Visscher. 2017. Prokaryotic diversity and biogeochemical

characteristics of benthic microbial ecosystems at La Brava, a hypersaline lake at Salar de

Atacama, Chile. PLoS One. 12(11):1–25. doi:10.1371/journal.pone.0186867

Fenchel, T., G. M., King, and T. H. Blackburn. 2012. Bacterial Metabolism, p. 1–34. In T.

Fenchel, G. M., King and T. H. Blackburn (eds) Bacterial Biogeochemistry (3rd edition).

Academic Press. doi:10.1016/B978-0-12-415836-8.00001-3

Fernandez, A. B., M. C. Rasuk, P. T. Visscher, and others. 2016. Microbial diversity in

sediment ecosystems (evaporites domes, microbial mats, and crusts) of Hypersaline

Laguna Tebenquiche, Salar de Atacama, Chile. Front. Microbiol. 7: 1–18.

145

References

doi:10.3389/fmicb.2016.01284

Foster, J. S., S. J. Green, S. R. Ahrendt, S. Golubic, R. P. Reid, K. L. Hetherington, and L.

Bebout. 2009. Molecular and morphological characterization of cyanobacterial diversity

in the stromatolites of highborne cay, bahamas. ISME J. 3: 573–587.

doi:10.1038/ismej.2008.129

Franks, J., and J. F. Stolz. 2009. Flat laminated microbial mat communities. Earth-Science Rev.

96: 163–172. doi:10.1016/j.earscirev.2008.10.004

Gallagher, K. L., T. J. Kading, O. Braissant, C. Dupraz, and P. T. Visscher. 2012. Inside the

alkalinity engine: The role of electron donors in the organomineralization potential of

sulfate-reducing bacteria. Geobiology 10: 518–530. doi:10.1111/j.1472-

4669.2012.00342.x

Garrity, G. M., J. G. Holt, J. M. Macy, T. Krafft, L. I. Sly, and G. I. Chrysiogenes. 2001.

Phylum BV. Chrysiogenetes phy. nov., In: Boone D. R., Castenholz R. W., Garrity G. M.

(eds) Bergey’s Manual® of Systematic Bacteriology. Springer, New York, NY, pp. 421–

425. doi:10.1007/978-0-387-21609-6_22

Gérard, E., S. D. De Goeyse, M. Hugoni, and others. 2018. Key role of Alphaproteobacteria

and Cyanobacteria in the formation of stromatolites of Lake Dziani Dzaha (Mayotte,

Western Indian Ocean). Front. Microbiol. 9:796. doi:10.3389/fmicb.2018.00796

Gerdes, G. 2010. What are Microbial Mats?, p. 3–25. In J. Seckbach and A. Oren [eds.],

Microbial Mats: Modern and Ancient Microorganisms in Stratified Systems.

Dordrecht:Springer. doi:10.1007/978-90-481-3799-2_1

Gillen, M. 2015. Hydrology and geochemistry of saline lakes on Rottnest Island, Western

Australia. (Unpublished Master’s Thesis). The University of Western Australia, Perth,

Australia.

Gleeson, D. B., D. Wacey, I. Waite, A. G. O’Donnell, and M. R. Kilburn. 2016. Biodiversity

146

References

of Living, Non-marine, Thrombolites of Lake Clifton, Western Australia. Geomicrobiol.

J. 33: 850–859. doi:10.1080/01490451.2015.1118168

Glunk, C., C. Dupraz, O. Braissant, K. L. Gallagher, E. P. Verrecchia, and P. T. Visscher. 2011.

Microbially mediated carbonate precipitation in a hypersaline lake, Big Pond (Eleuthera,

Bahamas). Sedimentology 58: 720–736. doi:10.1111/j.1365-3091.2010.01180.x

Golubic, S. 1976. Chapter 4.1 Organisms that Build Stromatolites, p. 113–126. In M. R. Walter

[eds.], Stromatolites. Dev. Sedimentol. doi:10.1016/S0070-4571(08)71132-4

Gomez-Alvarez, V., T. K. Teal, and T. M. Schmidt. 2009. Systematic artifacts in metagenomes

from complex microbial communities. ISME J. 3: 1314–1317. doi:10.1038/ismej.2009.72

Green, S. J., and L. L. Jahnke. 2010. Molecular Investigations and Experimental Manipulations

of Microbial Mats: A View to Paleomicrobial Ecosystems, p. 183–206. In J. Seckbach

and A. Oren [eds.], Microbial Mats: Modern and Ancient Microorganisms in Stratified

Systems. Dordrecht:Springer. doi:10.1007/978-90-481-3799-2_9

Grey, K., and N. J. Planavsky. 2009. Microbialites of Lake Thetis, Cervantes, Western

Australia — a field guide. Geological Survey of Western Australia, 21p. Retrieved from

http://dmpbookshop.eruditetechnologies.com.au/product/Microbialites-of-Lake-Thetis-

Cervantes-Western-Australia-a-field-guide.do

Gutiérrez-Preciado, A., A. Saghaï, D. Moreira, Y. Zivanovic, P. Deschamps, and P. López-

García. 2018. Functional shifts in microbial mats recapitulate early Earth metabolic

transitions. Nat. Ecol. Evol. 2(11): 1700–1708. doi:10.1038/s41559-018-0683-3

Hamilton, T. L., D. A. Bryant, and J. L. Macalady. 2016. The role of biology in planetary

evolution: Cyanobacterial primary production in low-oxygen Proterozoic oceans.

Environ. Microbiol. 18(2): 325–340. doi:10.1111/1462-2920.13118

Havemann, S. A., and J. S. Foster. 2008. Comparative characterization of the microbial

diversities of an artificial microbialite model and a natural stromatolite. Appl. Environ.

147

References

Microbiol. 74: 7410–7421. doi:10.1128/AEM.01710-08

Hubert, H. L., E. C. Rankey, and C. Omelon. 2018. Organic Matter, Textures, and Pore

Attributes of Hypersaline Lacustrine Microbial Deposits (Holocene, Bahamas). J.

Sediment. Res. 88(7): 872–849. doi:10.2110/jsr.2018.42

Hunt, D. E., and C. S. Ward. 2015. A network-based approach to disturbance transmission

through microbial interactions. Front. Microbiol. 6: 1182. doi:10.3389/fmicb.2015.01182

Inskeep, W. P., Z. J. Jay, S. G. Tringe, M. J. Herrgård, and D. B. Rusch. 2013. The YNP

metagenome project: Environmental parameters responsible for microbial distribution in

the yellowstone geothermal ecosystem. Front. Microbiol. 4(67): 1–15.

doi:10.3389/fmicb.2013.00067

Itoh, T. 2014. The family Thermoproteaceae, In: Rosenberg E., DeLong E. F., Lory S.,

Stackebrant E., Thompson F. (eds.) The Prokaryotes. Springer, Berlin, Heidelberg, pp.

389–401. doi:10.1007/978-3-642-38954-2_330

Jiang, H. B., H. M. Cheng, K. S. Gao, and B. S. Qiu. 2013. Inactivation of Ca2+/H+ exchanger

in Synechocystis sp. strain PCC 6803 promotes cyanobacterial calcification by

upregulating CO2-concentrating mechanisms. Appl. Environ. Microbiol. 79(13): 4048–

4055. doi:10.1128/AEM.00681-13

John, J., M. Hay, and J. Paton. 2009. Cyanobacteria in benthic microbial communities in

coastal salt lakes in Western Australia. Arch. Hydrobiol. Suppl. Algol. Stud. 130: 125–

135. doi:10.1127/1864-1318/2009/0130-0125

Jungblut, A. D., I. Hawes, T. J. Mackey, M. Krusor, P. T. Doran, D. Y. Sumner, J. A. Eisen,

C. Hillman, and A. K. Goroncy. 2016. Microbial mat communities along an oxygen

gradient in a perennially ice-covered Antarctic lake. Appl. Environ. Microbiol. 82(2):

620–630. doi:10.1128/AEM.02699-15

Kamennaya, N., C. Ajo-Franklin, T. Northen, and C. Jansson. 2012. Cyanobacteria as

148

References

Biocatalysts for Carbonate Mineralization. Minerals. 2(4): 338–364.

doi:10.3390/min2040338

Kawai, S., N. Kamiya, K. Matsuura, and S. Haruta. 2019. Symbiotic growth of a thermophilic

sulfide-oxidizing photoautotroph and an elemental sulfur-disproportionating

chemolithoautotroph and cooperative dissimilatory oxidation of sulfide to sulfate. Front.

Microbiol. 10: 2250. doi:10.3389/fmicb.2019.01150

Kennard, J. M., and N. P. James. 1986. Thrombolites and stromatolites: two distinct types of

microbial structures. Palaios 492–503.

Khodadad, C. L. M., and J. S. Foster. 2012. Metagenomic and metabolic profiling of

nonlithifying and lithifying stromatolitic mats of highborne cay, the Bahamas. PLoS One

7. doi:10.1371/journal.pone.0038229

Klanchui, A., S. Cheevadhanarak, P. Prommeenate, and A. Meechai. 2017. Exploring

Components of the CO2-Concentrating Mechanism in Alkaliphilic Cyanobacteria

Through Genome-Based Analysis. Comput. Struct. Biotechnol. J. 15: 340–350.

doi:10.1016/j.csbj.2017.05.001

Klatt, C. G., D. A. Bryant, and D. M. Ward. 2007. Comparative genomics provides evidence

for the 3-hydroxypropionate autotrophic pathway in filamentous anoxygenic phototrophic

bacteria and in hot spring microbial mats. Environ. Microbiol. 9(8): 2067–2078.

doi:10.1111/j.1462-2920.2007.01323.x

Konopka, A., S. Lindemann, and J. Fredrickson. 2015. Dynamics in microbial communities:

Unraveling mechanisms to identify principles. ISME J. 9: 1488–1495.

doi:10.1038/ismej.2014.251

Kurth, D., A. Amadio, O. F. Ordoñez, V. H. Albarracín, W. Gärtner, and M. E. Farías. 2017.

Arsenic metabolism in high altitude modern stromatolites revealed by metagenomic

analysis. Sci. Rep. 7: 1024. doi:10.1038/s41598-017-00896-0

149

References

Langmead, B., C. Trapnell, M. Pop, and S. L. Salzberg. 2009. Ultrafast and memory-efficient

alignment of short DNA sequences to the human genome. Genome Biol. 10.

doi:10.1186/gb-2009-10-3-r25

Lewin, G. R., C. Carlos, M. G. Chevrette, H. A. Horn, B. R. McDonald, R. J. Stankey, B. G.

Fox, and C. R. Currie. 2016. Evolution and Ecology of Actinobacteria and Their

Bioenergy Applications. Annu. Rev. Microbiol. 70: 235–254. doi:10.1146/annurev-

micro-102215-095748

Li, H. 2015. BFC: Correcting Illumina sequencing errors. Bioinformatics 31: 2885–2887.

doi:10.1093/bioinformatics/btv290

Llorens-Marès, T., S. Yooseph, J. Goll, J. Hoffman, M. Vila-Costa, C. M. Borrego, C. L.

Dupont, and E. O. Casamayor. 2015. Connecting biodiversity and potential functional role

in modern euxinic environments by microbial metagenomics. ISME J. 9: 1648–1661.

doi:10.1038/ismej.2014.254

Louyakis, A. S., J. M. Mobberley, B. E. Vitek, P. T. Visscher, P. D. Hagan, R. P. Reid, R.

Kozdon, I J. Orland, J. W. Valley, N. J. Planavsky, G. Casaburi, and J. S. Foster. 2017. A

Study of the Microbial Spatial Heterogeneity of Bahamian Thrombolites Using

Molecular, Biochemical, and Stable Isotope Analyses. Astrobiology. 17(5): 413–430.

doi:10.1089/ast.2016.1563

Margulis, L., E. S. Barghoorn, D. Ashendorf, S. Banerjee, D. Chase, S. Francis, S. Giovannoni,

and J. Stolz. 1980. The microbial community in the layered sediments at Laguna Figueroa,

Baja California, Mexico: Does it have analogues? Precambrian Res. 11: 93–

123. doi:10.1016/0301-9268(80)90042-X

McMurdie, P. J., and Holmes, S. 2013. Phyloseq: an R package for reproducible interactive

analysis and graphics of microbiome census data. PLoS One. 8(4): 1–11.

doi:10.1371/journal.pone.0061217.

150

References

McMurdie, P. J., and Holmes, S. 2014. Waste not, want not: why rarefying microbiome data is

inadmissible. PLoS Comput. Biol. 10(4): 1–12. doi:10.1371/journal.pcbi.1003531

Mendes Monteiro, J., R. Vogwill, K. Bischoff, and D. B. Gleeson. 2019. Comparative

metagenomics of microbial mats from hypersaline lakes at Rottnest Island (WA,

Australia), advancing our understanding of the effect of mat community and functional

genes on microbialite accretion. Limnol. Oceanogr. 00:1-17. doi:10.1002/lno.11323

Meyer, F., D. Paarmann, M. D’Souza, and others. 2008. The metagenomics RAST server–a

public resource for the automatic phylogenetic and functional analysis of metagenomes.

BMC Bioinformatics 9: 386.

Mlewski, E. C., C. Pisapia, F. Gomez, F. Gomez, L. Lecourt, E. S. Rueda, K. Benzerara, B.

Ménez, S. Borensztajn, F. Jamme, M. Réfrégiers, and E. Gérard. 2018. Characterization

of pustular mats and related Rivularia-rich laminations in oncoids from the Laguna Negra

lake (Argentina). Front. Microbiol. 9: 1–23. doi:10.3389/fmicb.2018.00996

Mobberley, J. M., C. L. M. Khodadad, and J. S. Foster. 2013. Metabolic potential of lithifying

cyanobacteria-dominated thrombolitic mats. Photosynth. Res. 118: 125–140.

doi:10.1007/s11120-013-9890-6

Naghoni, A., G. Emtiazi, M. A. Amoozegar, M. S. Cretoiu, L. J. Stal, Z. Etemadifar, S. A.

Shahzadeh Fazeli, and H. Bolhuis. 2017. Microbial diversity in the hypersaline Lake

Meyghan, Iran. Sci. Rep. 7: 1–13. doi:10.1038/s41598-017-11585-3

Najjari, A., M. S. Elshahed, A. Cherif, and N. H. Youssef. 2015. Patterns and Determinants of

Halophilic Archaea (Class Halobacteria) Diversity in Tunisian Endorheic Salt Lakes and

Sebkhet Systems. Appl. Environ. Microbiol. 81: 4432–4441. doi:10.1128/aem.01097-15

Nango, E., A. Royant, M. Kubo, and others. 2016. A three-dimensionalmovie of structural

changes in bacteriorhodopsin. Science. 354(6319): 1552–1557.

doi:10.1126/science.aah3497

151

References

Nierychlo, M., A. Miłobȩdzka, F. Petriglieri, B. McIlroy, P. H. Nielsen, and S. J. McIlroy.

2019. The morphology and metabolic potential of the Chloroflexi in full-scale activated

sludge wastewater treatment plants. FEMS Microbiol. Ecol. 95(2): 1–11.

doi:10.1093/femsec/fiy228

Nishida, A., V. Thiel, M. Nakagawa, S. Ayukawa, and M. Yamamura. 2018. Effect of light

wavelength on hot spring microbial mat biodiversity. PLoS One. 13(1): 1–30.

doi:10.1371/journal.pone.0191650

Noffke, N., D. Christian, D. Wacey, and R. M. Hazen. 2013. Microbially Induced Sedimentary

Structures Recording an Ancient Ecosystem in the ca. 3.48 Billion-Year-Old Dresser

Formation, Pilbara, Western Australia. Astrobiology. 13(12): 1103-1124.

doi:10.1089/ast.2013.1030

Nurk, S., D. Meleshko, A. Korobeynikov, and P. A. Pevzner. 2017. MetaSPAdes: A new

versatile metagenomic assembler. Genome Res. 27: 824–834. doi:10.1101/gr.213959.116

Nutman, A. P., V. C. Bennett, C. R. L. Friend, M. J. Van Kranendonk, and A. R. Chivas. 2016.

Rapid emergence of life shown by discovery of 3,700-million-year-old microbial

structures. Nature. 537(7621): 535–538. doi:10.1038/nature19355

O’Sullivan, L. A., J. Rinna, G. Humphreys, A. J. Weightman, and J. C. Fry. 2006. Culturable

phylogenetic diversity of the phylum “Bacteroidetes” from river epilithon and coastal

water and description of novel members of the family Flavobacteriaceae: Epilithonimonas

tenax gen. nov., sp. nov. and Persicivirga xylanidelens gen. nov., sp. nov. Int. J. Syst.

Evol. Microbiol. 56: 169–180. doi:10.1099/ijs.0.63941-0

Olson, S. L., C. T. Reinhard, and T. W. Lyons. 2016. Cyanobacterial diazotrophy and Earth’s

delayed oxygenation. Front. Microbiol. 7: 1–6. doi:10.3389/fmicb.2016.01526

Ondov, B. D., Bergman, N. H., and Phillippy, A. M. 2011. Interactive metagenomic

visualization in a Web browser. BMC bioinformatics. 12:385 doi:10.1186/1471-2105-12-

152

References

385

Oren, A. 2014a. The family Methanosarcinaceae, In: Rosenberg E., DeLong E. F., Lory S.,

Stackebrant E., Thompson F. (eds.) The Prokaryotes. Springer, Berlin, Heidelberg, pp.

259–281. doi:10.1007/978-3-642-38954-2_408

Oren, A. 2014b. The family Methanococcaceae, In: Rosenberg E., DeLong E. F., Lory S.,

Stackebrant E., Thompson F. (eds.) The Prokaryotes. Springer, Berlin, Heidelberg, pp.

215–224. doi:10.1007/978-3-642-38954-2_410

Pace, A., R. Bourillot, A. Bouton, E. Vennin, S. Galaup, I. Bundeleva, P. Patrier, C. Dupraz,

C. Thomazo, P. Sansjofre, Y. Yokoyama, M. Franceschi, Y. Anguy, L. Pigot, A. Virgone,

and P. T. Visscher. 2016. Microbial and diagenetic steps leading to the mineralisation of

Great Salt Lake microbialites. Sci. Rep. 6: 1–12. doi:10.1038/srep31495

Paerl, H. W., and A. C. Yannarell. 2010. Environmental Dynamics, Community Structure and

Function in a Hypersaline Microbial Mat, p. 421–442. In J. Seckbach and A. Oren [eds.],

Microbial Mats: Modern and Ancient Microorganisms in Stratified Systems.

Dordrecht:Springer. doi:10.1007/978-90-481-3799-2_22

Pagès, A., K. Grice, D. T. Welsh, P. T. Teasdale, M. J. Van Kranendonk, and P. Greenwood.

2015. Lipid Biomarker and Isotopic Study of Community Distribution and Biomarker

Preservation in a Laminated Microbial Mat from Shark Bay, Western Australia. Microb.

Ecol. 70(2): 459–472. doi:10.1007/s00248-015-0598-3

Pajares, S., V. Souza, and L. E. Eguiarte. 2015. Multivariate and phylogenetic analyses

assessing the response of bacterial mat communities from an ancient oligotrophic aquatic

ecosystem to different scenarios of long-term environmental disturbance. PLoS One.

10(3): 1–18. doi:10.1371/journal.pone.0119741

Parkhurst, D. L., and C. A. J. Appelo. 2013. Description of Input and Examples for PHREEQC

Version 3 — A Computer Program for Speciation , Batch-Reaction , One-Dimensional

153

References

Transport , and Inverse Geochemical Calculations. U.S. Geol. Surv. Tech. Methods, B. 6,

chapter A43. doi:10.1016/0029-6554(94)90020-5

Parks, D. H., G. W. Tyson, P. Hugenholtz, and R. G. Beiko. 2014. STAMP: Statistical analysis

of taxonomic and functional profiles. Bioinformatics 30: 3123–3124.

doi:10.1093/bioinformatics/btu494

Peimbert, M., L. D. Alcaraz, G. Bonilla-Rosso, G. Olmedo-Alvarez, F. García-Oliva, L.

Segovia, L. E. Eguiarte, and V. Souza. 2012. Comparative Metagenomics of Two

Microbial Mats at Cuatro Ciénegas Basin I: Ancient Lessons on How to Cope with an

Environment Under Severe Nutrient Stress. Astrobiology 12: 648–658.

doi:10.1089/ast.2011.0694

Pepe-Ranney, C., W. M. Berelson, F. A. Corsetti, M. Treants, and J. R. Spear. 2012.

Cyanobacterial construction of hot spring siliceous stromatolites in Yellowstone National

Park. Environ. Microbiol. 14: 1182–1197. doi:10.1111/j.1462-2920.2012.02698.x

Poli, A., I. Finori, I. Romano, A. Gioiello, L. Lama, and B. Nicolaus. 2017. Microbial Diversity

in Extreme Marine Habitats and Their Biomolecules. Microorganisms. 5(2): E25.

doi:10.3390/microorganisms5020025

Ponce-Soto, G. Y., E. Aguirre-von-Wobeser, L. E. Eguiarte, J. J. Elser, Z. M. P. Lee, and V.

Souza. 2015. Enrichment experiment changes microbial interactions in an ultra-

oligotrophic environment. Front. Microbiol. 6:1-11. doi:10.3389/fmicb.2015.00246

Preisner, E. C., E. B. Fichot, and R. S. Norman. 2016. Microbial mat compositional and

functional sensitivity to environmental disturbance. Front. Microbiol. 7: 1–16.

doi:10.3389/fmicb.2016.01632

Preisner, E. C., E. B. Fichot, and R. S. Norman. 2016. Microbial mat compositional and

functional sensitivity to environmental disturbance. Front. Microbiol. 7: 1–16.

doi:10.3389/fmicb.2016.01632

154

References

Prieto-Barajas, C. M., E. Valencia-Cantero, and G. Santoyo. 2018. Microbial mat ecosystems:

Structure types, functional diversity, and biotechnological application. Electron. J.

Biotechnol. 31: 48–56. doi:10.1016/j.ejbt.2017.11.001

Reitner, J. 2011. Microbial mats, In Reitner, J. and Thiel, V. (Eds). Encyclopedia Geobiology.

Part of the Series of Encyclopedia of Earth Sciences. Spinger, Heidelberg, pp. 606–608.

Revsbech, N. P., E. Trampe, M. Lichtenberg, D. M. Ward, and M. Kühl. 2016. In situ hydrogen

dynamics in a hot spring microbial mat during a diel cycle. Appl. Environ. Microbiol.

82(14): 4209–4217. doi:10.1128/AEM.00710-16

Richert, L., S. Golubic, R. Le Guédès, J. Ratiskol, C. Payri, and J. Guezennec. 2005.

Characterization of exopolysaccharides produced by cyanobacteria isolated from

Polynesian microbial mats. Curr. Microbiol. 51: 379–384.doi:10.1007/s00284-005-0069-

z

Riding, R. 2006. Microbial carbonate abundance compared with fluctuations in metazoan

diversity over geological time. Sediment. Geol. 185(3-4): 229–238.

doi:10.1016/j.sedgeo.2005.12.015

Riding, R. 2011. Microbialites, stromatolites, and thrombolites, p. 635–654. In J. Reitner and

V. Thiel [eds.], Encyclopedia of Geobiology: Encyclopedia of Earth Science Series.

Heidelberg:Springer. doi:10.1007/978-1-4020-9212-1_196

Rottnest Island Authority. 2014. Terrestrial Management Plan. Rottnest Island Authority.

URL: http://ria.wa.gov.au/policy-and-reports/management-strategies. Access date:

01/08/2020.

Russell, J. A., A. L. Brady, Z. Cardman, G. F. Slater, D. S. S. Lim, and J. F. Biddle. 2014.

Prokaryote populations of extant microbialites along a depth gradient in Pavilion Lake,

British Columbia, Canada. Geobiology. 12(3): 250–264. doi:10.1111/gbi.12082

Ruvindy, R., R. A. White, B. A. Neilan, and B. P. Burns. 2016. Unravelling core microbial

155

References

metabolisms in the hypersaline microbial mats of Shark Bay using high-throughput

metagenomics. ISME J. 10: 183–196. doi:10.1038/ismej.2015.87

Sanz-Montero, M. E., Ó. Cabestrero, and M. Sánchez-Román. 2019. Microbial Mg-rich

Carbonates in an Extreme Alkaline Lake (Las Eras, Central Spain). Front. Microbiol. 10:

1–15. doi:10.3389/fmicb.2019.00148

Schneider, D., G. Arp, A. Reimer, J. Reitner, and R. Daniel. 2013. Phylogenetic Analysis of a

Microbialite-Forming Microbial Mat from a Hypersaline Lake of the Kiritimati Atoll,

Central Pacific. PLoS One 8: e66662. doi:10.1371/journal.pone.0066662

Shadrin, N. V., and E. V. Anufriieva. 2019. Ecosystems of hypersaline waters: Structure and

trophic relations. Zh. Obshch. Biol. 79(6): 418–427. doi:10.1134/S0044459618060076

Sharma, A., M. Schmidt, B. Kiesel, N. K. Mahato, L. Cralle, Y. Singh, H. H. Richnow, J. A.

Gilbert, W. Arnold, and R. Lal. 2018. Bacterial and Archaeal Viruses of Himalayan Hot

Springs at Manikaran Modulate Host Genomes. Front. Microbiol. 9: 1–15.

doi:10.3389/fmicb.2018.03095

Shen, Y., V. Thiel, J. P. Duda, and J. Reitner. 2018. Tracing the fate of steroids through a

hypersaline microbial mat (Kiritimati, Kiribati/Central Pacific). Geobiology. 16(3): 307–

318. doi:10.1111/gbi.12279

Smith, M. D., S. E. Goater, E. S. Reichwaldt, B. Knott, and A. Ghadouani. 2010. Effects of

recent increases in salinity and nutrient concentrations on the microbialite community of

Lake Clifton (Western Australia): Are the thrombolites at risk? Hydrobiologia. 649: 207–

216. doi:10.1007/s10750-010-0246-3

Stolz, J. F. 2000. Structure of microbial mats and biofilms. In: Riding, R. E., Awramik, S. M.

(eds.) Microbial Sediments. Springer-Verlag, Berlin, Heidelberg, pp.1-8. 10.1007/978-3-

662-04036-2_1

Suosaari, E. P., R. P. Reid, P. E. Playford, J. S. Foster, J. F. Stolz, G. Casaburi, P. D. Hagan,

156

References

V. Chirayath, I. G. Macintyre, N. J. Planavsky and G. P. Eberli. 2016. New multi-scale

perspectives on the stromatolites of Shark Bay, Western Australia. Sci. Rep. 6: 20557.

doi:10.1038/srep20557

Suosaari, E. P., S. M. Awramik, R. P. Reid, J. F. Stolz, and K. Grey. 2018. Living Dendrolitic

Microbial Mats in Hamelin Pool, Shark Bay, Western Australia. Geosciences. 8(6): 1-17.

doi:10.3390/geosciences8060212

Thomazo, C., E. Couradeau, and F. Garcia-Pichel. 2018. Possible nitrogen fertilization of the

early Earth Ocean by microbial continental ecosystems. Nat. Commun. 9: 1–8.

doi:10.1038/s41467-018-04995-y

Timms, B. V. 2005. Salt lakes in Australia: Present problems and prognosis for the future.

Hydrobiologia 552: 1–15. doi:10.1007/s10750-005-1501-x

Tosca, N. J. ., and A. L. Masterson. 2014. Chemical controls on incipient Mg-silicate

crystallization at 25°C: Implications for early and late diagenesis. Clay Miner. 49: 165–

194. doi:10.1180/claymin.2014.049.2.03

Tourna, M., P. Maclean, L. Condron, M. O’Callaghan, and S. A. Wakelin. 2014. Links between

sulphur oxidation and sulphur-oxidising bacteria abundance and diversity in soil

microcosms based on soxB functional gene analysis. FEMS Microbiol. Ecol. 88: 538–

549. doi:10.1111/1574-6941.12323 van den Ende, F. P., and H. van Gemerden. 1993. Sulfide oxidation under oxygen limitation

by a thiobacillus thioparus isolated from a marine microbial mat. FEMS Microbiol. Ecol.

13: 69–78. doi:10.1111/j.1574-6941.1993.tb00052.x van Gemerden, H. 1993. Microbial mats: A joint venture. Mar. Geol. Elsevier Sci. Publ. B.V

113: 3–25. doi:10.1016/0025-3227(93)90146-M

Visscher, P, M. E Farías, M. A. Contreras, F. Novoa-Cortez, C. Rasuk, M. M. Patterson, P.

Philippot, M. C. Sforna, K. L. Gallagher, and C. Dupraz. 2014. The role of arsenic cycling

157

References

in carbonate precipitation in microbialites through time [abstract]. In: The Geology

Society of America Annual Meeting, 2014 October 19-22, British Columbia, Canada,

paper no 281-15.

Visscher, P. T., and H. van Gemerden. 1991. Production and consumption of

dimethylsulfoniopropionate in marine microbial mats. Appl. Environ. Microbiol. 57:

3237–3242.

Visscher, P. T., and J. F. Stolz. 2005. Microbial mats as bioreactors: Populations, processes,

and products. Palaeogeogr. Palaeoclimatol. Palaeoecol. 219: 87–100.

doi:10.1016/j.palaeo.2004.10.016

Vogwill, R., Gleeson, D., Bischoff, K., and Monteiro, J. M. 2019. Microbialite monitoring

program and criteria. (Unpublished Report). Hydro Geo Enviro Pty Ltd, Perth, Australia.

Wacey, D., L. Urosevic, M. Saunders, and A. D. George. 2018. Mineralisation of filamentous

cyanobacteria in Lake Thetis stromatolites, Western Australia. Geobiology. 16(2): 203–

215. doi:10.1111/gbi.12272

Warden, J. G., G. Casaburi, C. R. Omelon, P. C. Bennett, D. O. Breecker, and J. S. Foster.

2016. Characterization of microbial mat microbiomes in the modern thrombolite

ecosystem of lake clifton, western australia using shotgun metagenomics. Front.

Microbiol. 7: 1–14. doi:10.3389/fmicb.2016.01064

White, R. A., A. M. Chan, G. S. Gavelis, B. S. Leander, A. L. Brady, G. F. Slater, D. S. S. Lim,

and C. A. Suttle. 2016. Metagenomic analysis suggests modern freshwater microbialites

harbor a distinct core microbial community. Front. Microbiol. 6: 1–14.

doi:10.3389/fmicb.2015.01531

White, R. A., H. L. Wong, R. Ruvindy, B. A. Neilan, and B. P. Burns. 2018. Viral communities

of Shark Bay modern stromatolites. Front. Microbiol. 9: 1-15.

doi:10.3389/fmicb.2018.01223

158

References

White, R. A., I. M. Power, G. M. Dipple, G. Southam, and C. A. Suttle. 2015. Metagenomic

analysis reveals that modern microbialites and polar microbial mats have similar

taxonomic and functional potential. Front. Microbiol. 6: 966.

doi:10.3389/fmicb.2015.00966

Wilmeth, D. T., H. A. Johnson, B. W. Stamps, W. M. Berelson, B. S. Stevenson, H. S. Nunn,

S. L. Grim, M. L. Dillon, O. Paradis, F. A. Corsetti, and J. R. Spear. 2018. Environmental

and biological influences on carbonate precipitation within hot spring microbial mats in

Little Hot Creek, CA. Front. Microbiol. 9: 1–13. doi:10.3389/fmicb.2018.01464

Winston, R. B. 2000. Graphical User Interface for MODFLOW, Version 4: U.S. Geological

Survey Open-File Report 00-315, 27 p.

Wong, H. L., D. L. Smith, P. T. Visscher, and B. P. Burns. 2015. Niche differentiation of

bacterial communities at a millimeter scale in Shark Bay microbial mats. Sci. Rep. 5: 1–

17. doi:10.1038/srep15607

Wong, H. L., P. T. Visscher, R. A. White, D. L. Smith, M. M. Patterson, and B. P. Burns. 2017.

Dynamics of archaea at fine spatial scales in Shark Bay mat microbiomes. Sci. Rep. 7: 1–

12. doi:10.1038/srep46160

Wong, H. L., R. A. White, P. T. Visscher, J. C. Charlesworth, X. Vázquez-Campos, and B. P.

Burns. 2018. Disentangling the drivers of functional complexity at the metagenomic level

in Shark Bay microbial mat microbiomes. ISME J. 12:2619-2639. doi:10.1038/s41396-

018-0208-8

Wong, H., A. Ahmed-Cox, and B. Burns. 2016. Molecular Ecology of Hypersaline Microbial

Mats: Current Insights and New Directions. Microorganisms 4: 6.

doi:10.3390/microorganisms4010006

Zeyen, N., D. Daval, P. Lopez-Garcia, D. Moreira, J. Gaillardet, and K. Benzerara. 2017.

Geochemical Conditions Allowing the Formation of Modern Lacustrine Microbialites.

159

References

Procedia Earth Planet. Sci. 17: 380–383. doi:10.1016/j.proeps.2016.12.096

Zhang, J., J. Ma, D. Liu, S. Qin, S. Sun, J. Zhao, and S. F. Sui. 2017. Structure of phycobilisome

from the red alga Griffithsia pacifica. Nature. 551(7678): 57–63.

doi:10.1038/nature24278

Zhu, T., and M. Dittrich. 2016. Carbonate Precipitation through Microbial Activities in Natural

Environment, and Their Potential in Biotechnology: A Review. Front. Bioeng.

Biotechnol. 4: 1–21. doi:10.3389/fbioe.2016.00004

160

Appendices

Appendices

The following appendices include the publications that have arisen from this thesis, including co-authorship:

I. Mendes Monteiro, J., Vogwill, R., Bischoff, K., Gleeson, D. B., 2019. Comparative metagenomics of microbial mats from hypersaline lakes at Rottnest Island (WA, Australia), advancing our understanding of the effect of mat community and functional genes on microbialite accretion. Limnology and Oceanography, 00, pp.1-17. doi:10.1002/lno.11323

II. Bischoff, K., Wilson, M. E. J., Mendes Monteiro, J., Vogwill, R., George, A. D., 2019

(under review). Sedimentology of coastal carbonate lakes. Sedimentology, Manuscript ID

SED-2019-OM-036.R1.

A1

Appendices

Appendix-I

A2

Appendices

A3

Appendices

A4

Appendices

A5

Appendices

A6

Appendices

A7

Appendices

A8

Appendices

A9

Appendices

A10

Appendices

A11

Appendices

A12

Appendices

A13

Appendices

A14

Appendices

A15

Appendices

A16

Appendices

A17

Appendices

A18

Appendices

Appendix-II

A19

Appendices

1 Title of the article: Sedimentology of hypersaline coastal carbonate lakes 2 Autohr’s complete name: KARL BISCHOFF1*, MOYRA E.J. WILSON1, JULIANA 3 MENDES MONTEIRO2, RYAN VOGWILL3,, ANNETTE D. GEORGE1 4 Authors’ Institutional affiliations: 1School of Earth Sciences, The University of Western 5 Australia, 35 Stirling Highway, Perth 6009, Australia; 2 School of Agriculture and 6 Environment, The University of Western Australia, 35 Stirling Highway, Perth 6009, 7 Australia; 3Hydrogeoenviro, 16 Retford Place, Carine 6020, Australia 8 *Corresponding author 9 E-mail address: [email protected] 10 Key words: Rottnest Island, lacustrine, carbonate microbialites (thrombolites and 11 stromatolites), Holocene, sedimentary facies, microbial mats

12 ABSTRACT

13 Carbonate lakes are a common feature of semi-arid coastlines, yet there are few studies 14 comprehensively describing their modern sedimentology. Here, we investigate the sedimentary 15 environments and deposits of five endorheic and hypersaline carbonate lakes on Rottnest 16 Island, Western Australia, using field observations and petrographic analyses of samples 17 collected along seven transects. Lake types comprise shallow unstratified lakes (ephemeral to 18 <2.5 m depth) with continuous low-angle margins and monomictic deep lakes (>2.5 m depth) 19 with a slope break between littoral and basinal environments. Lake sediments are mostly 20 bioclastic and of three types. (1) Reworked calcareous sandstone bedrock (Tamala Limestone) 21 is mostly present in onshore areas. (2) Biogenic marine grains from a minor sea-level highstand 22 ~2–3 kyr ago are unevenly distributed, and locally include articulated bivalves and pristine 23 foraminifera, suggesting little detrital sedimentation since partial in situ deposition during the 24 lagoonal phase of the lakes. (3) Peloidal sediment composed of invertebrate faecal pellets and 25 micritised skeletal grains. Ubiquitous microbial mats in the lakes are spatially associated with 26 texturally diverse microbial carbonate and gypsum. Flocculent mats dominated by filamentous 27 cyanobacteria characterize the shallow lakes, and include lithifying mats spatially associated 28 with stromatolites, and non-lithifying mats containing thick intervals of extracellular polymeric 29 substances, locally containing spherulites and composite ooids. Oncoids form in onshore areas 30 of the deep lakes with heavy groundwater discharge. Texturally diverse microbial mats in the 31 deep lakes are coccoid-dominant, bathymetrically controlled and spatially associated with 32 thrombolites in littoral areas. Gypsum precipitation beyond the littoral–basinal slope break

A20

Appendices

33 suggests groundwater-driven variations in water chemistry influence carbonate precipitation in 34 the nearshore. Although compositionally and texturally similar to other modern carbonate lake 35 facies, biogenic carbonate in hypersaline coastal lakes is mostly relict marine, and carbonate 36 precipitates in littoral areas show greater textural diversity, reflecting steeper environmental 37 gradients.

38 INTRODUCTION

39 Carbonate lakes formed during Holocene eustatic sea-level rise, shoreline migration 40 and drainage blockage behind coastal sand bodies are a common feature of semi-arid coastlines 41 (Brooke, 2001; Cohen, 2003). Despite their importance as ecological refuges and recorders of 42 past environmental and/or sea-level change, sedimentological descriptions of carbonate lakes 43 in coastal settings are rare in the literature. In regions such as western and southern Australia, 44 where these lakes are a common coastal feature, published work has tended to focus on 45 microbialites forming at the lake margins (e.g., Grey et al., 1990; Moore et al., 1994; Rosen et 46 al., 1996; Coshell et al., 1998; Wright et al., 2005; Perri et al., 2012; Burne et al., 2014; Wacey 47 et al., 2018). As a result, the diversity and spatial distribution of lacustrine environments and 48 their facies in such settings remain largely undocumented. Discovery of hydrocarbon reservoirs 49 associated with lacustrine microbialites in offshore Brazil (Carminatti et al., 2008) and Angola 50 (Saller et al., 2016) further highlights the need for modern analogue data from these settings 51 (Awramik et al., 2012). Here we contribute new information on the exceptional variability 52 present in five modern coastal hypersaline lakes on Rottnest Island, Western Australia (Fig. 1). 53 Modern lake sedimentology has been described from glacial (Eyles et al., 1988), fluvial 54 (Bohacs, 2000), tectonically active (Cohen et al., 1987; Lambiase, 1990), volcanic (Last et al., 55 1990), playa and continental saline lake settings (Eugster et al., 1978; De Deckker, 1988; Last 56 et al., 1997). Coastal lakes comprise ~2.3% of the present day total global lake area (Meybeck, 57 1995) and are concentrated in semi-arid regions such as southern Australia (e.g., De Dekker, 58 1988; Chagas et al., 2016) and the eastern margin of southern Africa (e.g., Hill, 1975; Hobday, 59 1979). Subdued topography in these regions, coupled with the tendency for coastal lakes to 60 occupy hydrologically closed (endorheic; e.g., karstic or inter-dune) basins, typically limits 61 clastic input. The bedrock commonly comprises late Quaternary marine and/or coastal aeolian 62 deposits (Brooke, 2001; Hearty et al., 2008) typically resulting in carbonate-dominated settings 63 (e.g., Adlam, 2014; Bouton et al., 2016). Lake types that may share sedimentary features with 64 coastal carbonate lakes include: perennial continental-saline lakes that produce well-laminated 65 evaporite, carbonate, and siliciclastic facies (e.g., Last, 1990); and freshwater-karstic and maar

A21

Appendices

66 lakes that commonly contain indurated-carbonate hardgrounds and microbialites in littoral 67 areas (e.g., Arp et al., 1999; Morellón et al., 2009; Last et al., 2010). Rottnest Island contains 68 a range of unstratified ephemeral lakes and seasonally stratified perennial lakes. Although the 69 catchment geology is predominantly bioclastic carbonate material, these lake systems also 70 contain a significant microbial component, evaporites and minor siliciclastic sediments, 71 providing a unique opportunity to examine the variability in coastal carbonate lakes. 72 Coastal carbonate lakes may differ from other lake types in their microbial influence 73 and microbialite diversity, coupled with low clastic sedimentation rates and presence of marine 74 sediments (e.g., Wright et al., 2005; Spadafora et al., 2010). Variable lake morphology and 75 physiochemical conditions can result in diverse microbial mat communities, producing 76 microbialites with a wide range of external morphologies and internal fabrics. Microbialites in 77 coastal lakes predominantly have a clotted internal structure (e.g., Dupraz et al., 2004; Moore 78 et al., 1994). Modern lacustrine microbialites form primarily by in situ mineral precipitation 79 (e.g., Coshell et al., 1998; Laval et al., 2000; Gischler et al., 2008; Chidsey et al., 2015; Pace 80 et al., 2018), in contrast to modern marine examples formed primarily by grain trapping 81 processes (e.g., Planavsky et al., 2009; Jahnert et al., 2012). Modern lacustrine microbialites 82 may therefore provide the most suitable analogues for ancient Phanerozoic (e.g., Theisen et al., 83 2016; Kirkham et al., 2018) and Precambrian microbialites (e.g., Turner et al., 2000; Van 84 Kranendonk, 2006) also formed by in situ mineral precipitation. Although laminated 85 microbialites have been reported in Holocene lakes (e.g., Perri et al., 2012; Gomez et al., 2014) 86 they are uncommon, and new reports of modern examples may provide valuable insight into 87 the genesis of Precambrian stromatolites, due to the formation of both principally by 88 autochthonous mineral precipitation (Riding, 2000). Diverse microbial mat communities in 89 seven of nine hypersaline carbonate lakes on Rottnest Island are actively forming microbialites. 90 With the exception of brief descriptions of microbialite macro-morphologies and skeletal 91 limestones in Playford (1977, 1988) and microbial mat cyanobacterial and diatom assemblages 92 in John (2009), no detailed work has been published on the microbialites or lacustrine 93 sedimentology of Rottnest Island. 94 The general aim of this paper is to develop improved models for coastal carbonate lake 95 systems, to better understand the controlling influences on facies development. The specific 96 objectives are to: document the range and spatial distribution of lacustrine depositional 97 environments on Rottnest Island; describe their sedimentary and microbial facies; and discuss 98 controls on facies distributions and development. Here we present some of the first high spatial

A22

Appendices

99 resolution information on the lateral variability present within these dynamic and 100 paleoenvironmentally-important depositional settings.

101 GEOLOGICAL SETTING

102 Rottnest Island is situated 18 km west of the coast in Western Australia, in a tectonically 103 quiescent portion of the Perth Basin (Cawood and Nemchin, 2000; Fig. 1A;). The island 104 bedrock is composed of aeolian grainstones (of the Tamala Limestone), deposited as coastal 105 dune ridges during the last Pleistocene interglacial highstand (Hearty, 2003). Tamala 106 Limestone is locally intercalated with a coral framestone–bafflestone (Rottnest Limestone) on 107 the southern coast of the island (Stirling et al., 1995). Eustatic sea-level rise separated the island 108 from the mainland about 6.5 kyr ago (Eisenhauer et al., 1993; Lambeck et al., 2014) and the 109 subaqueous element of the connecting dune system now forms a ridge ~10 m deep (Fig. 1A). 110 The eastern end of the island has nine hypersaline lakes that are the focus of this study 111 (Fig. 1B; Table 1), and several lower salinity, ephemeral and perennial interdunal swamps 112 (Edward et al., 1959; Gouramanis et al., 2012). The lakes have elongate to ovoid shapes, and 113 are broadly aligned and deepen from east to west. Interpretation of the lakes as karst features 114 based on lake shape and the underlying carbonate geology by Playford (1988) has since been 115 questioned in comparable settings (Wyrwoll et al., 2006). Such shapes may have a growth 116 rather than purely karstic origin, with one possible mechanism being the biotic-self 117 organisation of mollusc beds in tidal flats stabilising the substrate (Schlager et al., 2015). 118 Limestones and skeletal sand containing a shallow marine and stenohaline fossil fauna 119 (Herschell Limestone; Fig. 1B) are found along the lake margins. The Herschell Limestone 120 overlies the Tamala Limestone and is overlain by microbialites (Playford, 1988). Radiocarbon 121 dates on Katelysia spp. bivalves and serpulid tubes at wave-cut notches in the Tamala 122 Limestone constrain the timing of two changes in relative sea level: an upper notch 2.4 m above 123 present mean sea level between 4.8 and 5.9 kyr ago, containing Katelysia rhytiphora; and a 124 lower notch ~1 m above present between 2.2 and 3.2 kyr ago, containing Katelysia scalarina. 125 Some of the lakes have been affected by human activity. The southern margin of 126 Government House Lake, the northern margin of Lake Herschell and the former isthmus 127 between the two (now a causeway) have been physically altered to build roads (Fig. 1B). 128 Garden Lake, located near the main settlement and golf course, has been affected by 129 eutrophication due to fertilizer runoff and by wastewater disposal/irrigation (John et al., 2009). 130 Pearse Lakes were used for salt production until 1952 (Playford et al., 1977). Care was taken 131 to work on undisturbed areas of the lakes during this study.

A23

Appendices

132 CLIMATE

133 Rottnest Island has a Mediterranean-type climate characterized by hot-dry summers and 134 cool-wet winters. The island has an average rainfall of 569 mm/year, of which 422 mm/year 135 falls between May and September (BOM, 2017; data from 1983–2015). Annual evapo- 136 transpiration rates far exceed precipitation, for example in 2016 when 1628 mm was recorded 137 (BOM, 2017). Strong southeasterly afternoon sea breezes, typically >30 km per hour, prevail 138 during summer. In contrast, more variable lighter winds, predominantly from the southwest 139 and northeast, characterize the winter months (BOM, 2017; data from 1965–2015).

140 HYDROLOGY AND HYDROCHEMISTRY

141 Lake levels are driven by rainfall recharge and evaporation cycles (Bryan et al., 2017) 142 and vary seasonally between around -1.1 (late autumn, i.e., April–May) to -0.5 m (late winter, 143 i.e., July–August) below the Australian Height Datum (Table 1). The shallow lakes (<2.5 m 144 autumn depth) have gently dipping sides, whereas the deep lakes (>2.5 m depth autumn depth) 145 have a sigmoidal slope break at ~1 m water depth. 146 Late summer salinity in the lakes can exceed 150 g L-1 and hypersaline lake water has 147 no measurable impact on the composition of a fresh groundwater lens approximately 1 km to 148 the west (Bryan et al., 2016). This coupled with seasonal change in lake levels of up to ~1 m 149 together suggest seawater intrusion is minimal. The deeper lakes (Government House Lake, 150 Lake Serpentine, Lake Herschell) are temperature and salinity stratified during winter and 151 spring: brines below the thermocline, located at water depths between 2–3 m, are up to 10°C 152 warmer than the surface (Bunn et al., 1984). Stratification is caused by rainfall-derived 153 freshwater and discharging groundwater spreading over denser hypersaline water during 154 autumn–winter, and is subsequently destroyed during summer periods by reduced lake water 155 levels and increased wind and wave action.

156 METHODS

157 Surficial deposits (upper 10–30 cm) were collected along seven transects into or across 158 five lakes during two visits to the island in early April and May of 2017 (Fig. 1B). Transect 159 locations were chosen to capture the sedimentological and modern physiochemical variability 160 within and between lakes. Samples were collected at visible changes in lithology, sediment 161 colour and/or composition, as well as microbial mat colour and/or surface morphology. 162 Submerged samples were collected with a trowel in water depths of <0.8 m and at deeper depths 163 with a Van Ween grab sampler lowered on a rope. In water depths greater than a meter where

A24

Appendices

164 these features could not be discerned, samples were collected from a boat at ~50 cm changes 165 in water depth. Both sampling methods resulted in very good recovery and no overt disturbance 166 in shallow (<1.2 m) waters where lithifying mats and microbialites are found. Recovery of non- 167 lithifying mats in water depths of >1.2 m was also good and involved minimal disturbance. 168 The physical characteristics (internal organisation, texture and surface morphology) of the 169 microbial mats at the mesoscale (cm–mm scale), and their relationship with sediments and 170 lithified samples, were recorded in the field. Representative microbial mat samples were 171 collected in sterile 80 ml vials and refrigerated at ~2° C for further examination. 172 Grain-size distributions for the <2 mm sieved sediment fraction were quantified with a 173 Malvern Mastersizer Hydro 2000G® laser particle-size analyser.. Samples were weighed and 174 treated with hydrogen peroxide to remove organic matter and weighed again to obtain total 175 organic carbon (TOC) content. Samples were then treated with a sodium hexametaphosphate 176 solution to disaggregate grains and run through the analyser. The >2 mm size fractions were 177 sieved and weighed, with the weights of all fractions normalised to yield overall grain-size 178 distributions. 179 Prior to embedding sediment samples in epoxy resin for thin section production, locally 180 abundant organic matter, notably extracellular polymeric substances (EPS) present in microbial 181 mats, was removed by decanting the sediment in de-ionised water 3–6 times. Quantitative and 182 textural data on sediment grains were obtained by petrographic examination of 43 thin sections 183 under a Nikon Eclipse Ci POL optical microscope. Between 250 and 300 grains were counted 184 in each sample using photomicrographs and the recursive point counting method in J- 185 Microvision® software (Roduit, 2008). Components >2 mm were identified and semi- 186 quantified through visual inspection. Foraminiferal genera in sediment samples were 187 investigated by removing the <100 µm fraction and examining the sediment under a light 188 microscope. Microbial mats were examined within two days of field collection, by placing a 189 small amount of microbial mat into paraffin on a thin section, compressing with a coverslip 190 and identifying the dominant cyanobacterial and diatom genera under a Nikon Eclipse Ci POL 191 optical microscope. 192 Mesoscale internal fabrics in lithified samples, including the microbialites, were 193 initially described from saw-cut sections. Microscale (mm–µm scale) features were 194 investigated with 26 resin-embedded petrographic thin sections (oriented perpendicular to the 195 depositional surface) using Leica DM2500M and Nikon Eclipse Ci POL optical microscopes. 196 Microbialite microfabric components and their elemental composition were further 197 investigated using broken fragments coated in 6 nm of platinum, on an FEI Verios XHR

A25

Appendices

198 scanning electron microscope (SEM), equipped with an Oxford Instruments X-Max 80 energy 199 dispersive X-ray spectroscopy (EDS) system and Oxford Instruments AZtec 3.0 nanoanalysis 200 software. Working distance ranged from 4.5–5.5 mm and the accelerating voltage was 5–15 201 keV. 202 Bulk mineralogy of sediment and lithified samples was determined by powder X-ray 203 diffraction using a Panalytical Empyrean® diffractometer. Measurements were made between 204 2-theta incidence angles of 5o–80 o. 205 Radiocarbon ages of two microbialite fabrics were obtained by Accelerator Mass 206 Spectrometry (AMS). Approximately 3 mg of diagnostic microbial carbonate material (e.g., 207 mesoclots, laminae) were subsampled with a dental drill 0.06 mm in diameter and treated with 208 hydrogen peroxide to remove organic carbon. Radiocarbon ages were converted to calendar 209 years using Calib 6.0 software. No marine reservoir correction was made because obtained 210 radiocarbon ages were significantly younger than the likely age of the lagoon–lake transition 211 (Playford, 1988). 212 DEFINITIONS 213 Lithified limestone facies are named following Dunham’s (1962) textural classification 214 scheme, modified by Embry and Klovan (1971). Unlithified sediment facies names are based 215 on modal grain size with a component prefix where the latter comprises at least 8% of the 216 sediment. Microbial mats are classified on the basis of surface morphology and internal 217 structure using existing names in the literature where possible (e.g., Golubic et al., 2000; 218 Jahnert and Collins, 2012). Microbial mats are, however, named from marine systems; no 219 systematic nomenclature for microbial mats in lacustrine settings exists. We use the term 220 microbialite in the sense of Burne and Moore (1987) to describe in situ carbonate precipitates 221 formed by the interaction of benthic microbial mats with detrital and chemical sediments. There 222 are two types of microbialite on Rottnest Island, those with: a laminated fabric (stromatolite, 223 adjective: stromatolitic); and a massive to clotted fabric (thrombolite, adjective: thrombolitic). 224 Designation of microbialite types follows Kennard and James (1986) and Burne and Moore 225 (1987).

226 LAKE ENVIRONMENTS AND FACIES

227 Using the approach of Platt and Wright (2009), four lacustrine sedimentary 228 environments are established according to the frequency with which they are submerged, 229 namely backshore, foreshore, littoral and basinal areas (Fig. 2). The backshore encompasses a 230 perennially emergent, sub-horizontal vegetated terrace 3–40 m wide, locally with a berm-

A26

Appendices

231 gravel ridge (0.4–3 m wide and 4–12 cm high) representing the high water mark of the lakes. 232 Backshore sediment occurs in shallow, elongate shore-parallel depressions (1–5 cm deep, 0.3– 233 1 m wide and up to 3 m long) between feather spear grass (Austrostipa flavescens), beaded 234 samphire (Sarcocornia quinqueflora) and scrubby samphire (Tecticornia halocnemoides; 235 Rippey et al., 2003). The foreshore environment includes areas that are submerged with an 236 annual to sub-annual frequency (4–30 m wide), and are typically associated with areas of 237 groundwater discharge (3–15 m wide). Seasonally to perennially submerged areas proximal to 238 the shoreline, containing diverse microbial mat communities to ~1.3 m autumn water depth, 239 define the littoral environment. Finally, perennially submerged areas characterized by reduced 240 microbial mat diversity (i.e., surface morphology, colour and internal texture) beyond the 241 littoral zone comprise the basinal environment. 242 Two lake types can be distinguished based on autumn water depth and lake-margin 243 morphology. Shallow lakes (Fig. 2A, B) range from ephemeral to perennial, are less than 2.5 244 m deep and have continuous, gently inclined slopes of 1.2°–4°. These include Lake Sirius (Fig. 245 S1), Pink Lake (Fig. 3) and Lake Baghdad (Figs. 4, 5). Deep lakes, by comparison, have 246 stepped or ‘bench’ style margins (sensu Murphy et al., 1980) featuring a slope break with a 247 maximum slope angle of 8°–15° between littoral and basinal environments. The deep lakes 248 investigated here are Lake Serpentine (Figs. 6, S2) and Government House Lake (Fig. S3). 249 Within these lake environments deposits are placed into skeletal limestone, bioclastic sediment, 250 spherulite and coated grain, microbial mat, microbialite and gypsum facies groups in which 23 251 facies are collectively recognised and summarised in Table 2.

252 Skeletal limestone facies group

253 Skeletal grainstone (Tamala Limestone bedrock)

254 Areas of topographic relief including most of the bedrock on Rottnest Island are 255 composed of skeletal grainstone (Playford et al., 1977; Hearty and O’Leary, 2008; Fig. 1). 256 Large-scale cross-bedding is common and includes eroded “platforms” around the lake margins 257 (Fig. 7A). Constituent grains comprise moderately sorted, fine to medium sand-sized abraded 258 fragments of molluscs, coralline algae, echinoderm and foraminifera, and up to ~20% quartz 259 (Fig. 7B). Skeletal grainstones with small serpulid fragments (~50 µm thick and up to 1 mm 260 long) form 2–5 cm thick layers beneath microbialites in eastern Lake Baghdad. 261 Skeletal floatstone–rudstone (Herschell Limestone) 262 This facies is up to 2.5 m thick, predominantly found in foreshore areas, and consists 263 of friable rudstone to strongly cemented floatstone with a micritic matrix (Playford, 1988; Figs.

A27

Appendices

264 1, 9C, D). Constituent grains comprise a diverse shallow-marine and mollusc-dominated fossil 265 fauna, including echinoderms, foraminifera, corals, polychaete tubes, chitons, scaphopods and 266 arthropods. The locally extinct bivalve Katelysia scalarina, and bivalves Brachidontes sp. and 267 Irus irus are most common. Locally in foreshore–littoral areas, these bivalves are articulated 268 and in life-positions (Fig. 7C). Intact fossils of large symbiont-bearing benthic foraminifera 269 Amphisorus are present around Government House Lake. The ratio of rotalid to miliolid 270 foraminifera is typically >3:1, consistent with foraminiferal assemblages in marine 271 environments at similar latitudes to Rottnest Island (e.g., Mossadegh et al., 2009). According 272 to radiocarbon dates of Katelysia bivalves, the Herschell Limestone was deposited during the 273 two mid–late Holocene highstands when sea levels were up to 2.4 m higher and the lakes were 274 connected to the sea (Playford, 1988).

275 Bioclastic sediment facies group

276 Unlithified sediment with an aragonite- or calcite-dominated mineral composition is 277 present in all shallow lake environments and in backshore and foreshore areas of the deep lakes 278 (Fig. 8). Although quartz sand locally comprises >10% of samples, siliciclastic content 279 typically makes up <5% of most sample as confirmed by X-ray diffraction (Table 2). Seagrass 280 root and rhizome fibres of Posidonia australis (Kendrick, G, pers. comm., 2017), are locally 281 abundant in shallow lake deposits and in the western foreshore areas of Lake Serpentine. 282 Additionally, rare semi-translucent and carbonate-encrusted Posidonia leaf blades are found in 283 Pink Lake. 284 Skeletal grain constituents are similar to those of the Herschell Limestone and include 285 bivalves, gastropods, benthic foraminifera, articulated and crustose coralline algae, calcareous 286 worm tubes, echinoid fragments, bryozoans, sponge spicules, ostracods, scaphopods, and crab 287 claw fragments (Fig. 8A–E, G, H). Numerous fragmented shells, mostly of Katelysia, typify 288 backshore areas, whereas bivalves in foreshore and littoral areas are commonly intact, locally 289 articulated and in life positions. In areas of groundwater discharge in western Lake Serpentine 290 and northern Government House Lake, mollusc valves and bioclastic and lithic granules form 291 the nuclei of oncoids (Fig. 2C). In contrast, bivalve fragments within the shallow lakes are 292 commonly abraded, bored and partially micritised. 293 Angular bioclasts (>2 mm) comprise a large percentage of the gravel fraction in most 294 foreshore areas and are mostly skeletal grainstone–packstones or rudstones resembling the 295 Tamala Limestone or Herschell Limestone (Fig. 8E, F). A third sub-rounded to angular clast

A28

Appendices

296 type is composed of heterogeneous wackestone, locally with irregular laminae and/or peloidal 297 aggregates. 298 Peloids are ubiquitous in the lake sediments. Three types can be distinguished (Fig. 8I). 299 (1) Rounded to sub-rounded, typically homogenous grains 50–250 µm in diameter are common 300 in foreshore–littoral areas of the shallow lakes. The size, shape and homogenous internal 301 structure are consistent with the faecal pellets of gastropods and polychaete worms (Flügel, 302 2010), the shells and agglutinated tubes of which are locally common in the lakes (Fig. 8H). 303 (2) Variously shaped grains ranging in size from 50 µm to 1 mm, often preserving some internal 304 structure, notably of coralline algae, are interpreted as abraded and partly to wholly micritised 305 skeletal grains. (3) The smallest peloids comprise rounded to sub-rounded homogenous grains 306 <5–15 µm in diameter, are usually clustered into aggregates up to ~300 µm wide and are 307 tentatively interpreted as colonies of coccoid cyanobacteria (e.g., Arp et al., 1999). 308 Sediment grain sizes in backshore and foreshore areas are sandy with large bimodal to 309 trimodal gravel fractions. The gravel fractions are predominantly composed of whole or 310 fragmented molluscs and lithic grains up to 6 cm in diameter (Fig. 8A). Clay- to silt-sized 311 fractions are notably absent from the deep lakes and western sides of the shallow lakes. In 312 contrast, muds comprise up to 20% of the sediment on eastern foreshore and littoral areas of 313 the shallow lakes.

314 Spherulite and coated grain facies group

315 Aggregates of fused and coalescent spherulites with a mostly radial crystal orientation 316 (Verrecchia et al., 1995), as well as composite ooids comprising peloidal aggregates sharing 317 multiple cortical laminae, are locally present in all shallow lakes. These allochems are, 318 however, especially common in western Lake Baghdad, where they occur within thick intervals 319 of dark grey and gelatinous organic matrix composed of EPS (Arp et al., 1999; Wingender et 320 al., 2012) in flocculent microbial mats (Figs. 4, 9A–D, 10B). Allochems transition laterally 321 within five meters from aggregates of spherulites ~100 µm in diameter, to composite ooids of 322 poorly sorted peloids (100–300 µm in diameter), encased within alternating, continuous 323 micritic and sparitic laminae, forming irregularly shaped clumps up to 1 mm wide. 324 Elliptical oncoids, 4–12 mm in diameter, along with irregular coatings over Katelysia 325 valves, are found in foreshore areas of Lake Serpentine and Government House Lake where 326 there is groundwater discharge (Figs 2C, 9E, F). The cortex is typically laminated, whereas the 327 grain nucleus comprises structureless micrite and/or shell fragments. Oncoid laminae are 328 concentric and semi-continuous alternating bands of pale finely crystalline carbonate and

A29

Appendices

329 micrite. Shell fragments and peloids locally disrupt laminae. Irregularly shaped voids, 330 approximating the shape of bivalve fragments, are common in oncoid nucleii.

331 Microbial mat facies group

332 Benthic microbial mats are ubiquitous on the floors of the Rottnest Island lakes and in 333 some foreshore areas where there is groundwater discharge (Table 2; Fig. 10A–J). On the basis

334 of surface morphology and/or internal structure five mat types are recognised, namely 335 flocculent, pustular, blister, nodular and perforate mats (Golubic et al., 2000; Jahnert and 336 Collins, 2012). The EPS in microbial mats varies substantially in internal colour and texture, 337 commonly containing an irregularly laminated upper part and a more massive lower part 338 varying significantly in thickness. To show this thickness variability, figure depictions of the 339 mats distinguish between upper and lower parts of the mats, however they comprise the same 340 microbial mat (e.g., Fig. 10B). In the deep lakes the microbialites transition systemically from 341 foreshore areas into the lakes (Fig. 6). All microbial mats are dominated by similar 342 cyanobacterial and diatom genera, which vary in abundance between mats (John et al., 2009). 343 Filamentous cyanobacteria include Oscillatoria, Microcoleus, Spirulina and Schizothrix. In 344 contrast, coccoid cyanobacteria are predominantly Gloeocapsa, and usually occur with 345 Aphanothece, which instead have an oval to cylindrical shape. Sulphate reducing bacteria are 346 abundant in the microbial mats on Rottnest Island as revealed by meta-genomic analyses 347 (Monteiro et al., in press). 348 Flocculent mats are restricted to the shallow lakes (Figs. 3–5, 10A–D). This mat has a 349 1–5 mm thick, smooth, non-cohesive yellow-green layer , over a <1–2 mm thick purple layer. 350 On the western side of ephemeral Pink Lake the upper layer of the flocculent mat contains pink 351 granular halite crystals and commonly has mollusc shells embedded in the surface. Locally, 352 this mat contains thick accumulations (up to ~25 cm thickness) of gelatinous EPS, as seen in 353 Pink Lake and western Lake Baghdad. Filamentous cyanobacteria predominate and are mostly 354 Schizothrix, whereas diatoms are mostly Navicula. 355 Pustular mat encrusts gravel in foreshore areas with heavy groundwater discharge, and 356 forms laterally extensive mats in littoral areas to autumn water depths of ~30 cm (Figs. 5, 6, 357 10E, F). This mat has a black pustular or crenulated surface made up of individual pustules 1– 358 8 mm wide and high. Layering is typically absent to weak and irregular beneath the surface 359 layer. However, semi-continuous patches of purple colouration are locally visible, likely 360 indicating some spatial differentiation of microorganisms. Cyanobacteria are predominantly of 361 coccoidal Gloeocapsa.

A30

Appendices

362 Blister mat predominates in littoral areas of the deep lakes in 20–50 cm water depth 363 (Figs. 6, 10G, H). This mat is characterized by bulbous to elongate undulations 2–8 cm in 364 diameter and ~1 cm high. From top to bottom, laminae comprise a pink–orange layer 1–3 mm 365 thick, a green layer 0.2–1 mm thick, and a purple layer 0.5–2 mm thick. Cyanobacteria are 366 predominantly Gloeocapsa, followed by Aphanothece and Schizothrix. In addition, diatom 367 frustules of Navicula and Mastogloia were observed. The nodular mat is most common in water 368 depths between between 0.7–1.3 m depths (Figs. 6, 10I, J). The mat is laminated and has an 369 orange surface layer 1 mm thick. Beneath the surface layer there is a green layer 0.3–1 mm 370 thick and, locally, a purple layer thickening in depressions up to 2 mm thickness. Organic 371 nodules 1–2 mm wide and 1–4 mm high aggregate into clumps on local relief up to 10 mm 372 wide and 12 mm high. Cyanobacteria mostly comprise Gloeocapsa, whereas diatom frustules 373 of both Navicula and Mastogloia were observed. 374 The perforate mat dominates the deep water areas of the large perennial lakes in 375 association with gypsum crusts (Figs. 6, 10K, J). The surface mat layer is emerald green, up to 376 5 mm thick, unlaminated and cohesive with a stippled or perforated surface. Surface 377 morphology is undulate to domical and locally columnar. Cyanobacteria are predominantly 378 Gloeocapsa and Aphanothece.

379 Microbialite facies group

380 Macrostructure

381 Well-lithified microbialites are extensive in foreshore and littoral environments to ~1.3 382 m autumn water depth, where they form tabular accumulations <30 cm thick and 1–40 m wide 383 (Figs. 2A, C, 3, 5, 6). At the macroscale three types of microbialite accumulation are defined 384 by their surface morphology, namely sheets, domes and cones. Sheets have planar to slightly 385 undulate surfaces and two types of internal organisation. The first comprises 1–3 planar and 386 stratiform layers 1–4 cm thick, containing variable amounts of shallow marine fossils, 387 characteristic of the shallow lakes. The second contains a single very well-cemented planar to 388 undulate layer 3–8 cm thick, typical of the shallow littoral environment of the deep lakes to 389 ~50 cm autumn water depth. Domes characterize foreshore areas of the deep lakes and occur 390 locally around the shallow lakes. Individual domes are sub-circular to ovoid in shape, 10–120 391 cm wide and <20 cm high, and comprise either several concentric and stratiform layers 1–3 cm 392 thick, or are massive and contain abundant bivalve fossils. Intact domes are hollow, uniformly 393 jointed and joints infilled by carbonate cement. Eroded domes only preserve the outermost 394 layer, suggesting they were also hollow. Cones, in contrast to sheets and domes, are restricted

A31

Appendices

395 to perennially submerged areas of the deep lakes and are 5–15 cm wide and high. On Rottnest 396 Island, the same microbialite external morphology may contain either stromatolitic (internally 397 laminated) and/or thrombolitic (internally unlaminated to clotted) (Burne et al., 1987) internal 398 fabrics.

399 Thrombolite facies

400 Two types of thrombolite facies were distinguished by the composition and 401 morphology of isolated masses (termed mesoclots) within the matrix (Shapiro et al., 2000): 402 thrombolites with discrete mesoclots (Type 1) and those with a more diffuse or mottled fabric 403 (Type 2). At the mesoscale, Type 1 thrombolites contain conspicuous and striking pale brown 404 mesoclots in a darker brown matrix, containing abundant and irregular pore space (Fig. 11). At 405 the microscale, mesoclots are revealed to be composed of fused peloidal aggregates within 406 micrite and locally envelop gastropod shells, while the matrix comprises peloidal and/or 407 skeletal grainstone (Fig. 11B, C). Mesoclots and detrital grains within the matrix are rimmed 408 with isopachous aragonite. Mesoclots range from 0.5–4 mm in diameter and vary in shape from 409 amoeboid to polylobate. Type 1 thrombolites are restricted to sheets in the shallow lakes and 410 were emergent at the time of sampling, making any microbial mat affinity uncertain (Fig. 3). 411 At the mesoscale, Type 2 thrombolites are characterized by diffuse dark mottled areas 412 in a pale brown matrix (Fig. 12A). At the microscale, mottled areas are revealed to comprise 413 circular to ovoid features ~5 μm wide, aggregating into contiguous and irregular clumps up to 414 ~3 mm in diameter, separated by acicular aragonite cement infilling interstices (Fig. 12B). SEM 415 observations of thrombolites sampled near the interface with the blister microbial mat, reveal 416 mineralised coccoid cells with similar dimensions to Gloeocapsa (Fig. 12C). SEM-EDS 417 measurements of thrombolite indicate that mineralised cells and the matrix have an elemental 418 composition consistent with aragonite (Fig. 12D). Type 2 thrombolites occur within sheets and 419 cones in littoral areas of the deep lakes, and are spatially associated with the blister and nodular 420 microbial mats, both characterized by abundant coccoidal cyanobacteria.

421 Stromatolite facies

422 Following Monty (1976), two types of stromatolite facies are distinguished by the 423 continuity and arrangement of laminae, namely: stromatolites with laterally discontinuous and 424 irregular laminae (Type 1); and those with continuous and concentrically arranged laminae 425 (Type 2). Type 1 stromatolites are present as discrete pale brown layers ~5–20 mm thick with 426 negligible porosity, either overlying skeletal floatstones or as a layer within mottled

A32

Appendices

427 thrombolites (Fig. 12A). At the microscale under a light microscope, laminae are composed of 428 wavy, discontinuous and tapering laminae of mottled sparite 20–150 μm thick (Fig. 12E). 429 Locally, irregular voids truncate laminae. Skeletal grains and peloids are variably present in 430 this fabric. At higher magnification, the mottled areas within laminae are revealed to comprise 431 rod-shaped molds and/or body fossils that are 3–4 μm long and <1 μm wide (Fig. 12F). Such 432 fossil dimensions are consistent with the bacillus bacterial morphotype (Jones et al., 2001). 433 Type 1 stromatolites were not observed directly in contact with a microbial mat, thus microbial 434 mat affinities are uncertain. 435 At the mesoscale, Type 2 stromatolites contain indistinct laminae within 10–20 mm 436 thick planar to undulate crusts (Fig. 13A). At the microscale under a light microscope, the 437 stromatolites comprise laterally continuous and alternating pale sparite laminae 10–80 μm 438 thick, alternating with dark 5–50 μm thick laminae of dark micrite (Fig. 13B). Together stacks 439 of these laminae couplets form small-scale domes 3–12 mm wide and 1–6 mm high. Under 440 SEM, crystalline carbonate laminae are revealed to contain abundant sub-circular to elongate 441 voids ~3–25 μm wide (Fig. 13C). SEM analyses of stromatolites sampled near the microbialite- 442 microbial mat interface show the presence of filament-like features (Fig. 13D). At higher 443 magnification under SEM, voids are revealed to be segmented, with width and segment spacing 444 dimensions similar to those present in the cyanobacterium Schizothrix (Fig. 13E). This fabric 445 is restricted to microbialites with a sheet morphology in the shallow lakes, particularly Lake 446 Baghdad, where it is spatially associated with the flocculent microbial mat, characterized by 447 abundant filamentous cyanobacteria.

448 Gypsum facies group

449 Authigenic gypsum deposits have formed in littoral areas of Lake Baghdad below 0.7 450 m water depth, in basinal areas of Lake Baghdad and the deep lakes (Figs, 5, 6). Gypsum 451 deposits take two forms. In the first, unconsolidated anhedral–lenticular granules occur within 452 transitional nodular to perforate mats between 1.3 and 2.3 m water depth. Although only 453 weakly aggregated into surficial crusts, this gypsum type is characteristic of Lake Baghdad 454 deposits. Below ~2.3 m water depth, sets of centimetre-thick gypsum crusts comprising 455 texturally similar lenticular granules form domes and columns 5–10 cm wide and high (Fig. 456 10L). The underside of the gypsum crusts have a granular ‘sugary’ texture. Needle-like 457 prismatic crystals ~100 μm wide and 1 mm long occur at the bottom of the crusts, characteristic 458 of crystal formation in supersaturated solutions (Magee, 1991).

459 Radiocarbon dating

A33

Appendices

460 To constrain the age of the microbialites two AMS radiocarbon ages were obtained 461 from a Type 1 stromatolite and Type 2 thrombolite, both from the same microbialite sample 462 (Fig. 12A). The thrombolite give an age of 1006 ± 16 calendar years, whereas the Type 1 463 stromatolite gives an age of 794 ± 18 calendar years. These ages are necessarily older than the 464 actual age of the microbialites because of the hard water effect (e.g., Gischler et al. 2008). This 465 results from the interaction of groundwater with Tamala Limestone bedrock (>70 kyr) that is 466 completely depleted in 14C, discharging into the lakes and diluting the 14C in the lake water, 467 from which the microbial carbonate precipitated. The dates are still useful, however, because 468 they give a maximum age for the microbialites.

469 Summary of lake types and their facies distribution

470 Two lake types (shallow and deep) on Rottnest Island, defined by depth and lake margin 471 morphology, contain distinctive surficial sedimentary deposits. Bioclastic sediment is largely 472 restricted to onshore environments of the deep lakes (Fig. 6), in contrast to the shallow lakes 473 where it is present at the depositional surface in all environments (Figs. 3–5). Bivalve sand 474 typifies backshore–foreshore areas and is dominated by locally extinct bivalve Katelysia 475 scalarina. In contrast, gastropod and peloidal sand are more common in littoral areas of the 476 shallow lakes, where they contain Hydrococcus spp., and Batillaria spp. and up to ~10% 477 miliolid foraminifera, including Quinquokkulina spp. and Triloculina sp. Gastropod sand is 478 notable for containing abundant Posidonia australis seagrass root and rhizome fibres, in 479 addition to rare leaf blades. Oncoids are largely confined to foreshore areas of the deep lakes 480 with heavy groundwater discharge, whereas spherulites and ooids appear to be restricted to 481 littoral areas of the shallow lakes where they are found within thick EPS intervals. 482 Microbial mats are spatially associated with particular microbialite fabrics and macro- 483 morphologies. Flocculent mats are confined to the shallow lakes and are spatially associated 484 with sheets containing stromatolites and gypsum below ~0.5 m in eastern areas, whereas 485 flocculent mats in western areas contain thick (up to ~25 cm)accumulations of EPS and locally 486 contain spherulites and ooids. Pustular mats locally form in the shallow lakes and are dominant 487 in foreshore areas of the deep lakes, particularly with groundwater discharge, forming in 488 association with oncoids, and microbialite sheet and domical macromorphologies containing 489 skeletal floatstones and mottled thrombolites. In contrast, blister and nodular mats are only 490 present in the deep lakes where they are spatially associated with stromatolites and thrombolites 491 within sheets, domes and cones. The perforate mat is only observed in basinal areas of the deep

A34

Appendices

492 lakes where it forms in association with gypsum crusts below the thermocline at ~2.5 m water 493 depth.

494 DISCUSSION

495 Carbonate lake facies models

496 The spatial distribution of lacustrine facies in the Rottnest Island lakes can be divided 497 into two groups based on environment, namely littoral (including foreshore areas) and basinal 498 facies, in keeping with published facies models for coastal and non-coastal carbonate lake 499 systems (Platt and Wright, 2009; Table 3). The unifying feature of carbonate lakes is a 500 limestone- and/or volcanic-dominated catchment resulting in mildly to strongly alkaline lake 501 waters. Most carbonate production, however, is concentrated in littoral environments where 502 light and oxygen are most abundant. Although bioclastic sediment locally dominates shoreline 503 areas in hypersaline coastal carbonate lakes, as on Rottnest Island and in Lake Marion in South 504 Australia (Perri et al., 2012), it is marine and records deposition during sea-level fluctuations 505 2–3 kyr ago (Playford, 1988). In contrast, biogenic carbonate facies in hyposaline coastal lakes 506 and non-coastal carbonate lakes are dominated by lacustrine ostracods and molluscs (De 507 Deckker, 1988; Moore et al., 1994) which in fresh to brackish lakes are typically found with 508 aquatic plants and charophytes (Cohen and Thouin, 1987; Morellón et al., 2009). Charophyte 509 meadows are a characteristic feature of low energy conditions in hyposaline lakes, where they 510 baffle, trap and stabilise mud. Hypersaline conditions on Rottnest Island preclude most fauna 511 and flora, including charophytes, with a few notable exceptions (e.g., Artemia in the water 512 column, halo-tolerant nematodes in microbial mats). High carbonate productivity in littoral

513 areas of hyposaline lakes has been attributed to CO2 uptake by respiring aquatic plants and 514 charophytes (Liu et al., 2008), promoting abiotic carbonate encrustation of algae and stems, 515 which then rapidly break down in situ to carbonate muds (Murphy et al., 1980) partly 516 explaining the paucity of micritic clays locally in the Rottnest Island lakes. In addition, the low 517 clay content in Rottnest Island facies may partly reflect higher energy conditions during the 518 lagoonal phase of the lakes winnowing out the clay fraction. 519 One striking aspect of carbonate lakes is the diversity of microbial features in littoral 520 environments. On Rottnest Island, carbonate supersaturated groundwater discharges into lake 521 water with high calcium concentrations, contributing to the formation of spherulites, ooids, 522 oncoids and microbialites in foreshore and littoral areas. Extensive aragonitic ooid sands have 523 formed in Great Salt Lake, as a result of strong seasonal runoff and groundwater influx, coupled 524 with much larger catchment areas and lake size enabling sufficient wind and wave action to

A35

Appendices

525 cause ooid saltation (Sandberg, 1975; Post, 1977). In Lake Tanganyika, high energy conditions 526 and development of ooid shoals and gastropod concentrations in the nearshore hinders the 527 formation of microbialites, which instead form at depths of >20 m (Cohen and Thouin, 1987). 528 In contrast, spherulites and ooids on Rottnest Island are concentrated in low energy areas and 529 appear to have formed with limited to no agitation. 530 Microbialite macro-morphology is influenced by water depth and accommodation 531 space availability. Microbialite accumulations in the shallow and ephemeral lakes mostly 532 comprise tabular sheets (Fig. 14). In contrast, microbialites with surface relief, notably conical 533 thrombolites, occur in perennially submerged areas of the deep lakes, and are similar in overall 534 morphology to those in perennial Lake Clifton (Moore and Burne, 1994; Fig. 15). 535 Stromatolites and thrombolites have both formed in the Rottnest Island lakes. Detailed 536 interpretation of their formation and diagenesis is beyond the scope of this paper and will be 537 published elsewhere. Thrombolites and the mesoclots that are their diagnostic components can 538 be primary features that did not result from metazoan disruption of pre-existing laminae 539 (Kennard and James, 1986), as shown by formation of thrombolites in hypersaline Great Salt 540 Lake (Pace et al., 2016) and on Rottnest Island. All reported microbialite occurrences in 541 hyposaline lakes appear to be thrombolites that co-exist with a grazing fauna. Hyposaline 542 localities include include Lake Clifton where conical thrombolites contain burrowing isopods 543 (Moore and Burne, 1994), domical thrombolites from Lake Tanganyika that are heavily grazed 544 by two species of gastropod (Cohen and Thouin, 1987) and dendritic thrombolites in Pavilion 545 Lake containing midges (Laval et al., 2000). Stromatolites, in contrast, rarely co-exist with a 546 grazing fauna, and as a result are more likely to be present in hypersaline lakes where grazing 547 pressure is reduced. Modern hypersaline localities where stromatolites have formed include 548 Lake Thetis (Grey et al., 1990), the Coorong Region of South Australia (Wright et al., 2005), 549 Cayo Coco island near Cuba (Pace et al., 2018), lakes in the Great Plains area of Canada (Last 550 et al., 2010) and on Rottnest Island. This suggests that hypersalinity is a necessary but not 551 sufficient condition to form stromatolites by in situ mineral precipitation, and that microbial 552 assemblage is equally important in determining the type of fabric that forms (Shiraishi et al., 553 2017). 554 In contrast to the extensive lateral variability in littoral areas, lake facies in deeper, open 555 water, can largely be predicted by the physiochemistry of the lake water, lake margin geometry 556 and detrital influx. Evaporite and/or laminite facies are commonly dominant, depending on the 557 degree of chemical stratification, anoxia and salinity (Last and Vance, 1997; Morellón et al.,

A36

Appendices

558 2009; Petrash et al., 2012). Gypsum dominates the basinal areas of the Rottnest lakes, 559 indicating that the water column is fully oxidising and detrital input is minimal.

560 Controls on facies variability

561 Bioclastic sediment and skeletal limestone mineral composition

562 The carbonate mineralogy of the Rottnest Island lake sediments reflects the proportion 563 of detrital grains and in situ precipitates, coupled with the extent of subaerial exposure since 564 deposition (Table 2). Detrital grains are mostly skeletal and were originally composed of either 565 aragonite (most molluscs, including Katelysia) or high-Mg calcite (HMC) (foraminifera, 566 coralline algae, echinoderms and some molluscs). These two minerals are only stable sub- 567 aqueously in marine or marine-influenced conditions (including coastal lakes), and rapidly alter 568 to low-Mg calcite (LMC) during subaerial exposure (Bathurst, 1972; Kindler et al., 1996). The 569 Tamala Limestone skeletal grainstones comprising Rottnest Island’s bedrock have been 570 heavily altered from their original mineralogy and are now mostly LMC. In contrast, skeletal 571 floatstones–rudstones (Herschell Limestone) have undergone less alteration and retain more of 572 their original aragonite and HMC composition. Similarly, bioclastic sediment contains both 573 LMC and HMC, suggesting in the case of LMC derivation from the Tamala Limestone, and in 574 the case of HMC, derivation from, or deposition roughly contemporaneous with, the Herschell 575 Limestone, as seen also at Lake Thetis and Lake Marion (Borch et al., 1977; Grey et al., 1990). 576 These contrast with peloidal sands mostly composed of aragonite and minor HMC, suggesting 577 little subaerial exposure.

578 Authigenic mineral composition

579 Mineral composition of lacustrine microbialites, spherulites and coated grains trends 580 with lake water chemistry (Chagas et al., 2016). In the present study, this must be qualified by 581 the observation that lake conditions may have differed in the past when the precipitates were 582 forming. Microbialite ages obtained by AMS 14C dating of ~800–1000 years, however, must 583 be older than the true microbialite ages due to the hard water effect, suggesting lake conditions 584 have not changed substantially since the microbialites formed. Microbialites, spherulites and 585 coated grains on Rottnest Island are mostly composed of aragonite, as revealed by XRD and 586 confirmed by SEM-EDS analyses of microbialite samples. The ionic composition of coastal 587 lakes reflects their seawater derivation and concentration through groundwater runoff and 588 evaporation. Thus, lake water on Rottnest Island is Na-Cl dominant, hypersaline, and contains 2+ 2- 589 much higher Mg and SO4 concentrations than seawater (Table 1). Numerous empirical and

A37

Appendices

590 experimental studies indicate that precipitation of calcite versus aragonite is directly related to 591 the Mg/Ca ratio (Fyfe et al., 1965; Fernandez-Diaz et al., 1996). This is because the presence 592 of Mg2+ ions inhibits calcite nucleation while having no effect on aragonite precipitation, which 593 does not incorporate Mg2+ into its crystal lattice. Removal of Ca2+ by aragonite and calcite 594 precipitation increases Mg/Ca ratios until calcite precipitation is no longer possible. Aragonite 595 appears to be favoured when Mg/Ca is >2, and calcite when Mg/Ca is ≤2 (Sandberg, 1975; 596 Given et al., 1985). Consequently, modern microbialites in marine settings, such as the 597 Bahamas (Feldmann et al., 1998) and Hamelin Pool, Western Australia (Logan, 1961), are 598 composed of aragonite. With the exception of Pink Lake (described below), the Rottnest Island 599 lakes have Mg/Ca ratios ranging from 5.5 to 12.1 (Table 1), resulting in the preferential 600 precipitation of aragonite. These values are similar to several other West Australian coastal 601 lakes, including Lake Thetis and Lake Clifton, both of which contain well developed 602 thrombolites composed of aragonite in littoral areas (Grey et al., 1990; Moore and Burne, 2- 603 1994). Additionally, high concentrations of dissolved SO4 , as are present in the Rottnest 604 Island lakes, may decrease the Mg/Ca ratio in which aragonite precipitation is favoured over 605 calcite (Bots et al., 2011). 606 Pink Lake represents a conspicuous departure from Mg/Ca ratios approximating 607 seawater, of >80 during the summer of 2015–2016 (Table 1). Empirical and experimental data 608 from the Coorong Region indicate that bacterial sulphate reduction can overcome kinetic 2- 2+ 2+ 609 constraints on dolomite formation, by removing SO4 , releasing Mg and Ca ions and 610 contributing to Mg/Ca ratios of up to 329, but typically 30–60 (Wright, 1999; Wright and 2- 611 Wacey, 2005). Pink Lake differs markedly in its SO4 concentration, which is 2–10 times 612 higher than the lakes in the Coorong (Wright, 1999). This suggests either that sulfate-reducing 2- 613 bacteria have been relatively inactive in Pink Lake, and/or high concentrations of SO4 614 inhibited the formation of dolomite (Warren, 2000).

615 Environmental change

616 Shallow marine fossils at the present day depositional surface, locally including intact 617 echinoids, pristine foraminifera and articulated bivalves in life position, are indicative of very 618 low sedimentation rates and low energy conditions (Fig. 16). High percentages of miliolid 619 foraminifera, present in the shallow lakes and particularly Pink Lake, and their occurrence with 620 Hydrococcus spp., Batillaria spp., are typical of elevated salinities, observable today in the 621 salinity gradient within restricted marine conditions at Hamelin Pool, Western Australia 622 (Mossadegh et al., 2009; Suosaari et al., 2016). It should be noted, however, that this miliolid

A38

Appendices

623 and gastropod assemblage may also occur in stenohaline conditions. In addition, miliolids have 624 a much wider salinity tolerance range than their accompanying biogenic carbonate-forming 625 organisms, thus may have been deposited later and mixed in with the other skeletal grains (e.g., 626 Burne et al., 1982). Nonetheless, such faunal variations are consistent with progressive 627 severance of the lagoons from the sea and a shift to increasingly restricted conditions (e.g., 628 Bufarale et al., 2015) and/or to uneven patterns of water circulation and salinity (Playford et 629 al., 1977). 630 The seagrass Posidonia australis grows profusely in dense meadow-like stands on the 631 seafloor in southern and western Australia, including today in the Indian Ocean around Rottnest 632 Island. It has massive rhizomes, a deep root system and broad, strip-like leaves (Jernakoff et 633 al., 1996). The variable abundance of Posidonia debris in lake sediments suggests that patchy 634 meadows flourished when the lakes were connected to the sea (Fig. 16). In dense Posidonia 635 meadows, the prolific growth of seagrass and dense rhizome mats can inhibit active bivalve 636 infauna (James et al., 2007). This may partly account for the variable distribution of gastropod 637 and bivalve dominated facies. In addition, Posidonia leaves would have acted as a sediment 638 baffle inhibiting transport of silts and clays onshore (Ginsburg et al., 1958). Interestingly, large 639 benthic foraminifera, such as Amphisorus, which occur in close association with seagrasses 640 including Posidonia, are conspicuously absent from the shallow lake sediments, but present in 641 foreshore areas of Government House Lake where Posidonia was not observed. This may be 642 due to a number of factors, including the poor preservation potential of Posidonia in subaerially 643 exposed areas where Amphisorus are observed, asynchronous deposition, preferential post 644 depositional transport of Posidonia fragments to the west, meteoric dissolution of Amphisorus 645 tests, and/or local differences in the paleoenvironment.

646 Lake physiography and microbial carbonate distribution

647 Sedimentation rates partly reflect topography and the size of lake catchment areas for 648 both ground- and surface-water, which appears to influence the distribution of microbialites, 649 spherulites and coated grains. The lakes are located in the northeast corner of the island 650 resulting in a much larger catchment area to the west, particularly for the western lakes, 651 coincident with stronger observed groundwater discharge, thicker lake sediment intervals and 652 a paucity of microbialites (Figs. 4, 5). Additionally, currents created by strong prevailing 653 southeasterly winds during summer may ‘push’ or re-suspect sediment towards western shores, 654 and the absence of microbialites in the west may partly reflect greater sediment fluxes (Fig. 655 14). Microbial mats in these areas are well-developed, forming EPS accumulations greater than

A39

Appendices

656 25 cm thick (Fig. 4). EPS can both inhibit and promote CaCO3 precipitation (Dupraz et al. 657 2004). Cyanobacteria produce most EPS (Decho et al., 2004), which can chelate large amounts 2+ 658 of dissolved Ca ions, inhibiting precipitation of CaCO3 (Kawaguchi et al., 2002). 659 Alternatively, degradation of EPS by heterotrophic bacteria can release calcium ions,

660 promoting CaCO3 precipitation (Decho et al., 2005; Braissant et al., 2009). Included in 661 heterotrophic bacteria are sulfate-reducing bacteria, which have been implicated in the

662 precipitation of CaCO3 in microbialites in a number of studies (e.g., Braissant et al., 2009; 663 Decho, 2010). Accumulation of substantial EPS in western areas of the shallow lakes, 664 therefore, may reflect both greater sediment fluxes and a higher ratio of phototrophic to 665 heterotrophic bacteria. 666 Littoral areas of the deep lakes and eastern sides of the shallow lakes contain 667 microbialites (Figs. 3, 5, 6). Location of lithifying microbial mats proximal to the shore, the 668 groundwater-fed character of the lakes and seasonal stratification in the deep lakes, suggests 669 variations in lake water chemistry between littoral and basinal areas may be an important 670 control on the distribution of microbialites. Bathymetric control on microbialites has been 671 reported in other lakes, notably Lake Clifton, in which coalescent reef-like structures pass into 672 domical forms with increasing depth (Moore and Burne, 1994), similar to the transition from 673 undulate to conical forms in deep lake littoral areas on Rottnest Island. 674 Oncoids are locally developed in higher energy foreshore areas of the deep lakes with heavy 675 groundwater discharge, and at the time of sampling were encrusted with pustular microbial mat 676 characterized by abundant coccoidal cyanobacteria (Figs. 6, 9E, F). The oncoids are mostly 677 composed of micrite around a core of skeletal debris and/or lithic fragments, within an envelope 678 of irregular sparitic laminae. Alterations in cortical laminae thickness in oncoids from other 679 terrestrial environments have been attributed to changes in the thickness of enveloping 680 biofilms, precipitating carbonate minerals (Gerdes et al., 1994) and/or trapping sediment grains 681 (Jones, 2011). 682 In contrast, aggregates of fused and coalescent spherulites, and peloid-nucleated 683 composite ooids sharing cortical laminae are restricted to littoral areas of the shallow lakes

684 (Figs. 4, 9A–D). Field and experimental studies have shown that agitation and CaCO3 685 supersaturation state are important controls on spherulite and ooid formation and texture. 686 Spherulites and radial ooids precipitate in quieter settings with no/less agitation, whereas 687 concentrically laminated (tangential) ooids form in higher energy environments commonly 688 involving saltation (Sandberg, 1975; Davies et al., 1976; Strasser, 1986). Spherulites owe their 689 radial crystal orientation to formation in situ and may subsequently form ooid nuclei given

A40

Appendices

690 grain salation (Davies et al., 1978; Verrecchia et al., 1995). Extensive ooid sands in Great Salt 691 Lake vary from radial to tangential crystal orientations, partly attributable to seasonal changes 692 in lake water physiochemistry (Sandberg, 1975). In contrast, tangential ooid sands in the 693 Bahama’s, rather than changing texturally based on these two parameters, in quieter conditions 694 beyond the surf zone are cemented into clumps (grapestones) by micrite between ooids (O'reilly 695 et al., 2017). The composite ooids on Rottnest Island differ from these examples, however, by 696 their aggregation into irregular and friable clumps by shared cortical laminae between peloids, 697 also suggesting largely in situ formation (Strasser, 1986). 698 A number of studies have suggested a biotic role in the formation of spherulites and 699 ooids, emphasising the role of microbes in carbonate precipitation (Verrecchia et al., 1995; 700 Davaud et al., 2001; Kahle, 2007), or the presence of a common microbial community between 701 disparate locations (Summons et al., 2013; O'reilly et al., 2017). Location of the Rottnest Island 702 spherulites and ooids within thick EPS intervals in the perennial shallow lakes may have 703 provided a viscous medium to precipitate highly friable aggregates in situ. In addition, 704 accumulation of thick EPS layers and their subsequent degradation, liberating chelated Ca2+ 705 ions (Dupraz et al., 2009), combined with strong seasonal and highly alkaline runoff, as shown 706 in the groundwater chemistry (Table. 1), could have created conditions highly favourable to 707 spherulite and ooid precipitation. A more detailed interpretation of the formation of the 708 spherulites on Rottnest Island is beyond the scope of this paper and will be published 709 elsewhere. 710 The transition to gypsum precipitation below ~1.3 water depth may be attributable to 711 lake water chemistry as well as changes in microbial mat composition. The intimate association 712 between undulate to domical gypsum crusts and the perforate mat suggests a microbial 713 influence on the external morphology. Microbially-influenced gypsum deposits have been 714 reported in Israel (Canfield et al., 2004) and gypsum thrombolites are present in hypersaline 715 island lagoons in Venezuela (Petrash et al., 2012). Location of the transition between gypsum 716 granules and crusts at the thermocline also suggests periodic undersaturation and gypsum 717 dissolution at this depth interval (Cohen et al., 1977). 718 Apparent variation in microbial mat diversity between shallow and deep lakes likely 719 reflects differences in lake depth, size, chemistry and margin morphology. Small and/or 720 shallow lakes are more prone to seasonal dilution by meteoric input or concentration by 721 evaporation. Moreover, gentle slope margins in the shallow lakes result in greater seasonal 722 shoreline movement. Consequently, the lake water chemistry and shoreline locations are less 723 prone to change in the deep lakes versus the shallow lakes, consistent with the greater number

A41

Appendices

724 of microbial mat facies and microbialite macro-morphologies in the deep lakes, notably the 725 conical thrombolites that are the subject of ongoing research.

726 CONCLUSIONS

727 This study and comparison with published work from elsewhere shows that hypersaline 728 carbonate lakes in coastal settings display several distinctive features. 729  Very low sedimentation rates, related to their location in semi-arid climates with low- 730 lying catchment areas, coupled with hypersaline conditions precluding most flora and 731 fauna, result in lake systems that are largely in stasis in terms of biogenic carbonate 732 production and detrital input. 733  Consequently, surficial bioclastic sediment of marine origin in littoral areas comprise 734 relict grains formed during late Holocene sea level fluctuations. 735  The lakes are strongly microbially influenced: most deposits since the Holocene sea- 736 level highstands comprise biological and chemical precipitates, notably texturally and 737 morphologically diverse microbialites in littoral areas, as well as oncoids, spherulites 738 and ooids, and gypsum in open water. Therefore, it is possible to establish the 739 sedimentological evolution of the lakes since the Holocene highstands using the 740 surficial sediments. 741  Stromatolites and thrombolites are both present on Rottnest Island. This reinforces the 742 view that thrombolitic fabrics can be primary features formed by in situ mineral 743 precipitation. Hypersalinity, however, also hinders a grazing fauna that would disrupt 744 stromatolitic laminae and contribute to the formation of the thrombolites that 745 characterize most lacustrine microbialites. This suggests that modern stromatolites 746 formed by in situ precipitation are a characteristic feature of hypersaline carbonate 747 lakes, though they are not restricted to coastal settings. 748  Littoral facies differ between ephemeral and perennial lakes. Conical microbialites are 749 restricted to perennially submerged littoral areas of the deep lakes, whereas ephemeral 750 lakes contain tabular crusts and bioclastic sediment in deeper areas. This reflects both 751 water depth and ephemerality controlling accommodation space and influencing the 752 development of microbial mats, and the presence of a much larger catchment area 753 supplying sediment to the ephemeral lakes. 754 The Rottnest Island lake facies share features with other modern carbonate lake systems, 755 however the environmental gradients in hypersaline coastal lakes are steeper. Negligible 756 detrital input since the lagoonal phase of the lakes in the late Holocene, coupled with their

A42

Appendices

757 morphological heterogeneity, hypersalinity and alkalinity, has facilitated the development of 758 an exceptionally diverse suite of surficial facies. Together, these facies record a transition from 759 marine to lacustrine conditions, and provide a modern laboratory for studying microbially 760 influenced carbonates (e.g., microbialites, spherulites, ooids, oncoids) and evaporites. Ongoing 761 research will utilise the information presented here to develop mineralisation models for the 762 microbialites, and help clarify environmental changes recorded in the Rottnest Island lake and 763 lagoonal sediments.

764 ACKNOWLEDGEMENTS

765 This research was funded by the Rottnest Island Authority (RIA), School of Earth 766 Sciences (SES), University of Western Australia (UWA) and the Australian Nuclear Science 767 and Technology Organisation. Karl Bischoff was supported by an Australian Government 768 Research Training Program Scholarship at UWA, and an Australian Institute for Nuclear 769 Science and Engineering Postgraduate Research Award. We are grateful to Rosine Riera 770 (UWA) for her help in the field and Luke Wheat and Cassanya Thomas (RIA) for logistical 771 support. Anne Brearley (Indian Ocean Marine Research Centre, UWA) helped with mollusc 772 identification, and David Haig (SES, UWA) with foraminiferal identification. We thank Frank 773 Nemeth (UWA) for thin section preparation and Chris Brouwer (UWA) for performingthe 774 grain-size analyses. Electon microscropy was done at the Centre for Microscopy, 775 Characterisation and Analysis at UWA. Stanley Awramik provided a very useful review of an 776 earlier version of this manuscript. This work was performed in partial fulfilment of 777 requirements for a Ph.D. degree by KB at UWA.

778 REFERENCES

779 Arp, G., Thiel, V., Reimer, A., Michaelis, W., and Reitner, J., 1999, Biofilm exopolymers 780 control microbialite formation at thermal springs discharging into the alkaline Pyramid 781 Lake, Nevada, USA: Sedimentary Geology, v. 126, no. 1-4, p. 159-176. 782 Awramik, S. M., and Buchheim, H. P., 2012, The quest for microbialite analogs to the South 783 Atlantic Pre-Salt carbonate hydrocarbon reservoirs of Africa and South America: HGS 784 International Dinner Meeting talk. URL: (http://www.hgs.org/node/4875). 785 Bathurst, R. G., 1972, Carbonate sediments and their diagenesis, Elsevier. 786 Bohacs, K. M., 2000, Keys to Exploration: Lake-Basin Type, Source Potential, and 787 Hydrocarbon Character within an Integrated Sequence-Stratigraphic/Geochemical 788 Framework.

A43

Appendices

789 Borch, C., Bolton, B., and Warren, J., 1977, Environmental setting and microstructure of 790 subfossil lithified stromatolites associated with evaporites, Marion Lake, South 791 Australia: Sedimentology, v. 24, no. 5, p. 693-708. 792 Bots, P., Benning, L., Rickaby, R., and Shaw, S., 2011, The role of SO4 in the switch from 793 calcite to aragonite seas: Geology, v. 39, no. 4, p. 331-334. 794 Bouton, A., Vennin, E., Pace, A., Bourillot, R., Dupraz, C., Thomazo, C., Brayard, A., 795 Désaubliaux, G., and Visscher, P. T., 2016, External controls on the distribution, fabrics 796 and mineralization of modern microbial mats in a coastal hypersaline lagoon, Cayo 797 Coco (Cuba): Sedimentology, v. 63, no. 4, p. 972-1016. 798 Braissant, O., Decho, A. W., Przekop, K. M., Gallagher, K. L., Glunk, C., Dupraz, C., and 799 Visscher, P. T., 2009, Characteristics and turnover of exopolymeric substances in a 800 hypersaline microbial mat: FEMS microbiology ecology, v. 67, no. 2, p. 293-307. 801 Brooke, B., 2001, The distribution of carbonate eolianite: Earth-Science Reviews, v. 55, no. 1- 802 2, p. 135-164. 803 Bryan, E., Meredith, K. T., Baker, A., Andersen, M. S., and Post, V. E., 2017, Carbon dynamics 804 in a Late Quaternary-age coastal limestone aquifer system undergoing saltwater 805 intrusion: Science of The Total Environment, v. 607, p. 771-785. 806 Bryan, E., Meredith, K. T., Baker, A., Post, V. E., and Andersen, M. S., 2016, Island 807 groundwater resources, impacts of abstraction and a drying climate: Rottnest Island, 808 Western Australia: Journal of Hydrology, v. 542, p. 704-718. 809 Bufarale, G., and Collins, L. B., 2015, Stratigraphic architecture and evolution of a barrier 810 seagrass bank in the mid-late Holocene, Shark Bay, Australia: Marine Geology, v. 359, 811 p. 1-21. 812 Bunn, S., and Edward, D., 1984, Seasonal meromixis in three hypersaline lakes on Rottnest 813 Island, Western Australia: Marine and Freshwater Research, v. 35, no. 2, p. 261-265. 814 Burne, R., and Colwell, J., 1982, Temperate carbonate sediments of northern Spencer Gulf, 815 South Australia: a high salinity ‘foramol’province: Sedimentology, v. 29, no. 2, p. 223- 816 238. 817 Burne, R. V., and Moore, L. S., 1987, Microbialites: organosedimentary deposits of benthic 818 microbial communities: Palaios, p. 241-254. 819 Canfield, D. E., Sørensen, K., and Oren, A., 2004, Biogeochemistry of a gypsum‐encrusted 820 microbial ecosystem: Geobiology, v. 2, no. 3, p. 133-150. 821 Carminatti, M., Wolff, B., and Gamboa, L., New exploratory frontiers in Brazil, in Proceedings 822 19th World Petroleum Congress2008, World Petroleum Congress.

A44

Appendices

823 Cawood, P., and Nemchin, A., 2000, Provenance record of a rift basin: U/Pb ages of detrital 824 zircons from the Perth Basin, Western Australia: Sedimentary Geology, v. 134, no. 3- 825 4, p. 209-234. 826 Chagas, A. A., Webb, G. E., Burne, R. V., and Southam, G., 2016, Modern lacustrine 827 microbialites: Towards a synthesis of aqueous and carbonate geochemistry and 828 mineralogy: Earth-Science Reviews, v. 162, p. 338-363. 829 Chidsey, T. C., Berg, M. D. V., and Eby, D. E., 2015, Petrography and characterization of 830 microbial carbonates and associated facies from modern Great Salt Lake and Uinta 831 Basin's Eocene Green River Formation in Utah, USA: Geological Society, London, 832 Special Publications, v. 418, no. 1, p. 261-286. 833 Cohen, A. S., 2003, Palaeolimnology: the history and evolution of lake systems, Oxford 834 University Press, New York. 835 Cohen, A. S., and Thouin, C., 1987, Nearshore carbonate deposits in Lake Tanganyika: 836 Geology, v. 15, no. 5, p. 414-418. 837 Cohen, Y., Krumbein, W. E., Goldberg, M., and Shilo, M., 1977, Solar lake (Sinai). 1. Physical 838 and chemical limnology: Limnology and Oceanography, v. 22, no. 4, p. 597-608. 839 Davaud, E., and Girardclos, S., 2001, Recent freshwater ooids and oncoids from western Lake 840 Geneva (Switzerland): indications of a common organically mediated origin: Journal 841 of Sedimentary Research, v. 71, no. 3, p. 423-429. 842 Davies, P. J., Bubela, B., and Ferguson, J., 1978, The formation of ooids: Sedimentology, v. 843 25, no. 5, p. 703-730. 844 Davies, P. J., and Martin, K., 1976, Radial aragonite ooids, Lizard Island, great barrier reef, 845 Queensland, Australia: Geology, v. 4, no. 2, p. 120-122. 846 De Deckker, P., 1988, Biological and sedimentary facies of Australian salt lakes: 847 Palaeogeography, Palaeoclimatology, Palaeoecology, v. 62, no. 1-4, p. 237-270. 848 Decho, A. W., Visscher, P. T., and Reid, R. P., 2005, Production and cycling of natural 849 microbial exopolymers (EPS) within a marine stromatolite, Geobiology: Objectives, 850 Concepts, Perspectives, Elsevier, p. 71-86. 851 Dunham, R. J., 1962, Classification of carbonate rocks according to depositional textures. 852 Dupraz, C., Reid, R. P., Braissant, O., Decho, A. W., Norman, R. S., and Visscher, P. T., 2009, 853 Processes of carbonate precipitation in modern microbial mats: Earth-Science Reviews, 854 v. 96, no. 3, p. 141-162.

A45

Appendices

855 Dupraz, C., Visscher, P. T., Baumgartner, L., and Reid, R., 2004, Microbe–mineral 856 interactions: early carbonate precipitation in a hypersaline lake (Eleuthera Island, 857 Bahamas): Sedimentology, v. 51, no. 4, p. 745-765. 858 Edward, D., and Watson, J., 1959, Fresh water and brackish water swamps of Rottnest Island: 859 Journal of the Royal Society of Western Australia, v. 42, p. 85-86. 860 Eisenhauer, A., Wasserburg, G., Chen, J., Bonani, G., Collins, L., Zhu, Z., and Wyrwoll, K., 861 1993, Holocene sea-level determination relative to the Australian continent: U/Th 862 (TIMS) and 14C (AMS) dating of coral cores from the Abrolhos Islands: Earth and 863 Planetary Science Letters, v. 114, no. 4, p. 529-547. 864 Embry, A. F., and Klovan, J. E., 1971, A late Devonian reef tract on northeastern Banks Island, 865 NWT: Bulletin of Canadian Petroleum Geology, v. 19, no. 4, p. 730-781. 866 Eugster, H. P., and Hardie, L. A., 1978, Saline lakes, Lakes, Springer, p. 237-293. 867 Eyles, N., Eyles, C., and Miall, A. D., 1988, Glacial facies models: Geology, v. 16, no. 4, p. 868 374-375. 869 Feldmann, M., and McKenzie, J. A., 1998, Stromatolite-thrombolite associations in a modern 870 environment, Lee Stocking Island, Bahamas: Palaios, v. 13, no. 2, p. 201-212. 871 Fernandez-Diaz, L., Putnis, A., Prieto, M., and Putnis, C. V., 1996, The role of magnesium in 872 the crystallization of calcite and aragonite in a porous medium: Journal of sedimentary 873 research, v. 66, no. 3, p. 482-491. 874 Flügel, E., 2010, New Perspectives in Microfacies, Microfacies of Carbonate Rocks, Springer, 875 p. 1-6. 876 Fyfe, W. S., and Bischoff, J. L., 1965, The calcite aragonite problem. 877 Gerdes, G., Dunajtschik‐Piewak, K., Riege, H., Taher, A., Krumbein, W., and Reineck, H. E., 878 1994, Structural diversity of biogenic carbonate particles in microbial mats: 879 Sedimentology, v. 41, no. 6, p. 1273-1294. 880 Ginsburg, R. N., and Lowenstam, H. A., 1958, The influence of marine bottom communities 881 on the depositional environment of sediments: The journal of geology, v. 66, no. 3, p. 882 310-318. 883 Gischler, E., Gibson, M. A., and Oschmann, W., 2008, Giant Holocene freshwater 884 microbialites, Laguna Bacalar, Quintana Roo, Mexico: Sedimentology, v. 55, no. 5, p. 885 1293-1309. 886 Given, R. K., and Wilkinson, B. H., 1985, Kinetic control of morphology, composition, and 887 mineralogy of abiotic sedimentary carbonates: Journal of Sedimentary Research, v. 55, 888 no. 1, p. 109-119.

A46

Appendices

889 Golubic, S., Seong-Joo, L., and Browne, K. M., 2000, Cyanobacteria: architects of sedimentary 890 structures, Microbial sediments, Springer, p. 57-67. 891 Gomez, F. J., Kah, L. C., Bartley, J. K., and Astini, R. A., 2014, Microbialites in a high-altitude 892 Andean lake: Multiple controls on carbonate precipitation and lamina accretion: 893 Palaios, v. 29, no. 6, p. 233-249. 894 Gouramanis, C., Dodson, J., Wilkins, D., De Deckker, P., and Chase, B., 2012, Holocene 895 palaeoclimate and sea level fluctuation recorded from the coastal Barker Swamp, 896 Rottnest Island, south-western Western Australia: Quaternary Science Reviews, v. 54, 897 p. 40-57. 898 Grey, K., Moore, L., Burne, R., Pierson, B., and Bauld, J., 1990, Lake Thetis, Western 899 Australia: and example of saline lake sedimentation dominated by benthic microbial 900 processes: Marine and Freshwater Research, v. 41, no. 2, p. 275-300. 901 Hearty, P. J., 2003, Stratigraphy and timing of eolianite deposition on Rottnest Island, Western 902 Australia: Quaternary Research, v. 60, no. 2, p. 211-222. 903 Hearty, P. J., and O’Leary, M. J., 2008, Carbonate eolianites, quartz sands, and Quaternary sea- 904 level cycles, Western Australia: a chronostratigraphic approach: Quaternary 905 Geochronology, v. 3, no. 1, p. 26-55. 906 Hobday, D., 1979, Geological evolution and geomorphology of the Zululand coastal plain: 907 Lake Sibaya, v. 36, p. 1. 908 Jahnert, R. J., and Collins, L. B., 2012, Characteristics, distribution and morphogenesis of 909 subtidal microbial systems in Shark Bay, Australia: Marine Geology, v. 303, p. 115- 910 136. 911 James, N. P., and Bone, Y., 2007, A late Pliocene–early Pleistocene, inner-shelf, subtropical, 912 seagrass-dominated carbonate: roe calcarenite, great Australian bight, western 913 Australia: Palaios, v. 22, no. 4, p. 343-359. 914 Jernakoff, P., Brearley, A., and Nielsen, J., 1996, Factors affecting grazer-epiphyte interactions 915 in temperate seagrass meadows: Oceanography and Marine Biology: an annual review. 916 John, J., Hay, M., and Paton, J., 2009, Cyanobacteria in benthic microbial communities in 917 coastal salt lakes in Western Australia: Algological Studies, v. 130, no. 1, p. 125-135. 918 Jones, B., 2011, Biogenicity of terrestrial oncoids formed in soil pockets, Cayman Brac, British 919 West Indies: Sedimentary Geology, v. 236, no. 1-2, p. 95-108. 920 Jones, L. J., Carballido-López, R., and Errington, J., 2001, Control of cell shape in bacteria: 921 helical, actin-like filaments in Bacillus subtilis: Cell, v. 104, no. 6, p. 913-922.

A47

Appendices

922 Kahle, C. F., 2007, Proposed origin of aragonite Bahaman and some Pleistocene marine ooids 923 involving bacteria, nannobacteria (?), and biofilms: Carbonates and Evaporites, v. 22, 924 no. 1, p. 10. 925 Kawaguchi, T., and Decho, A. W., 2002, A laboratory investigation of cyanobacterial 926 extracellular polymeric secretions (EPS) in influencing CaCO3 polymorphism: Journal 927 of Crystal Growth, v. 240, no. 1-2, p. 230-235. 928 Kennard, J. M., and James, N. P., 1986, Thrombolites and stromatolites: two distinct types of 929 microbial structures: Palaios, p. 492-503. 930 Kindler, P., and Hearty, P. J., 1996, Carbonate petrography as an indicator of climate and sea‐ 931 level changes: new data from Bahamian Quaternary units: Sedimentology, v. 43, no. 2, 932 p. 381-399. 933 Kirkham, A., and Tucker, M. E., 2018, Thrombolites, spherulites and fibrous crusts (Holkerian, 934 Purbeckian, Aptian): Context, fabrics and origins: Sedimentary Geology, v. 374, p. 69- 935 84. 936 Lambeck, K., Rouby, H., Purcell, A., Sun, Y., and Sambridge, M., 2014, Sea level and global 937 ice volumes from the Last Glacial Maximum to the Holocene: Proceedings of the 938 National Academy of Sciences, v. 111, no. 43, p. 15296-15303. 939 Lambiase, J. J., 1990, A model for tectonic control of lacustrine stratigraphic sequences in 940 continental rift basins: Lacustrine basin exploration: Case studies and modern analogs: 941 AAPG Memoir, v. 50, p. 265-276. 942 Last, F. M., Last, W. M., and Halden, N. M., 2010, Carbonate microbialites and hardgrounds 943 from Manito Lake, an alkaline, hypersaline lake in the northern Great Plains of Canada: 944 Sedimentary Geology, v. 225, no. 1, p. 34-49. 945 Last, W. M., and Deckker, P., 1990, Modern and Holocene carbonate sedimentology of two 946 saline volcanic maar lakes, southern Australia: Sedimentology, v. 37, no. 6, p. 967-981. 947 Last, W. M., and Vance, R. E., 1997, Bedding characteristics of Holocene sediments from salt 948 lakes of the northern Great Plains, Western Canada: Journal of Paleolimnology, v. 17, 949 no. 3, p. 297-318. 950 Laval, B., Cady, S. L., Pollack, J. C., and McKay, C. P., 2000, Modern freshwater microbialite 951 analogues for ancient dendritic reef structures: Nature, v. 407, no. 6804, p. 626. 952 Liu, Z., Liu, X., and Liao, C., 2008, Daytime deposition and nighttime dissolution of calcium 953 carbonate controlled by submerged plants in a karst spring-fed pool: insights from high 954 time-resolution monitoring of physico-chemistry of water: Environmental geology, v. 955 55, no. 6, p. 1159-1168.

A48

Appendices

956 Logan, B. W., 1961, Cryptozoon and associate stromatolites from the Recent, Shark bay, 957 Western Australia: The Journal of Geology, p. 517-533. 958 Magee, J., 1991, Late Quaternary lacustrine, groundwater, aeolian and pedogenic gypsum in 959 the Prungle Lakes, southeastern Australia: Palaeogeography, Palaeoclimatology, 960 Palaeoecology, v. 84, no. 1-4, p. 3-42. 961 Meybeck, M., 1995, Global distribution of lakes, Physics and chemistry of lakes, Springer, p. 962 1-35. 963 Monteiro, J. M., Vogwill, R., Bischoff, K., and Gleeson, D., in press, Comparative 964 metagenomics of microbial mats from hypersaline lakes at Rottnest Island (WA, 965 Australia), advancing our understanding of the effect of mat community and functional 966 genes on microbialite accretion: Limnology and Oceanography. 967 Moore, L. S., and Burne, R., 1994, The modern thrombolites of Lake Clifton, western 968 Australia, Phanerozoic stromatolites II, Springer, p. 3-29. 969 Morellón, M., VALERO‐GARCÉS, B., Anselmetti, F., Ariztegui, D., Schnellmann, M., 970 Moreno, A., Mata, P., Rico, M., and Corella, J. P., 2009, Late Quaternary deposition 971 and facies model for karstic Lake Estanya (North‐eastern Spain): Sedimentology, v. 56, 972 no. 5, p. 1505-1534. 973 Mossadegh, Z., Haig, D., Allan, T., Adabi, M., and Sadeghi, A., 2009, Salinity changes during 974 late Oligocene to early Miocene Asmari formation deposition, Zagros mountains, Iran: 975 Palaeogeography, Palaeoclimatology, Palaeoecology, v. 272, no. 1, p. 17-36. 976 Murphy, D. H., and Wilkinson, B. H., 1980, Carbonate deposition and facies distribution in a 977 central Michigan marl lake: Sedimentology, v. 27, no. 2, p. 123-135. 978 Nell, J., and Paterson, K., 1997, Salinity studies on the clams Katelysia rhytiphora (Lamy) and 979 Tapes dorsatus (Lamarck): Aquaculture Research, v. 28, no. 2, p. 115-119. 980 O'reilly, S., Mariotti, G., Winter, A., Newman, S., Matys, E., McDermott, F., Pruss, S., Bosak, 981 T., Summons, R., and Klepac‐Ceraj, V., 2017, Molecular biosignatures reveal common 982 benthic microbial sources of organic matter in ooids and grapestones from Pigeon Cay, 983 The Bahamas: Geobiology, v. 15, no. 1, p. 112-130. 984 Pace, A., Bourillot, R., Bouton, A., Vennin, E., Braissant, O., Dupraz, C., Duteil, T., 985 Bundeleva, I., Patrier, P., and Galaup, S., 2018, Formation of stromatolite lamina at the 986 interface of oxygenic–anoxygenic photosynthesis: Geobiology, v. 16, no. 4, p. 378-398. 987 Perri, E., Tucker, M., and Spadafora, A., 2012, Carbonate organo‐mineral micro‐and 988 ultrastructures in sub‐fossil stromatolites: Marion lake, South Australia: Geobiology, v. 989 10, no. 2, p. 105-117.

A49

Appendices

990 Petrash, D. A., Gingras, M. K., Lalonde, S. V., Orange, F., Pecoits, E., and Konhauser, K. O., 991 2012, Dynamic controls on accretion and lithification of modern gypsum-dominated 992 thrombolites, Los Roques, Venezuela: Sedimentary Geology, v. 245, p. 29-47. 993 Planavsky, N., and Ginsburg, R. N., 2009, Taphonomy of modern marine Bahamian 994 microbialites: Palaios, v. 24, no. 1, p. 5-17. 995 Platt, N., and Wright, V. P., 2009, Lacustrine carbonates: facies models, facies distributions 996 and hydrocarbon aspects: Lacustrine facies analysis, p. 57-74. 997 Playford, P. E., 1988, Guidebook to the geology of Rottnest Island, Geological Society of 998 Australia, WA Division and the Geological Survey of Western Australia. 999 Playford, P. E., Leech, R. E., and Kendrick, G. W., 1977, Geology and hydrology of Rottnest 1000 Island, Geological Survey of Western Australia. 1001 Post, F. J., 1977, The microbial ecology of the Great Salt Lake: Microbial Ecology, v. 3, no. 2, 1002 p. 143-165. 1003 Riding, R., 2000, Microbial carbonates: the geological record of calcified bacterial–algal mats 1004 and biofilms: Sedimentology, v. 47, no. s1, p. 179-214. 1005 Roduit, N., 2008, JMicroVision: Image analysis toolbox for measuring and quantifying 1006 components of high-definition images: Ver, v. 1, no. 7, p. 2002-2007. 1007 Saller, A., Rushton, S., Buambua, L., Inman, K., McNeil, R., and Dickson, J. T., 2016, Presalt 1008 stratigraphy and depositional systems in the Kwanza Basin, offshore Angola: AAPG 1009 Bulletin, v. 100, no. 7, p. 1135-1164. 1010 Sandberg, P. A., 1975, New interpretations of Great Salt Lake ooids and of ancient non‐skeletal 1011 carbonate mineralogy: Sedimentology, v. 22, no. 4, p. 497-537. 1012 Schlager, W., and Purkis, S., 2015, Reticulate reef patterns–antecedent karst versus self‐ 1013 organization: Sedimentology, v. 62, no. 2, p. 501-515. 1014 Shapiro, R. S., and Awramik, S. M., 2000, Microbialite morphostratigraphy as a tool for 1015 correlating Late Cambrian–Early Ordovician sequences: The Journal of Geology, v. 1016 108, no. 2, p. 171-180. 1017 Shiraishi, F., Hanzawa, Y., Okumura, T., Tomioka, N., Kodama, Y., Suga, H., Takahashi, Y., 1018 and Kano, A., 2017, Cyanobacterial exopolymer properties differentiate microbial 1019 carbonate fabrics: Scientific reports, v. 7, no. 1, p. 11805. 1020 Stirling, C., Esat, T., McCulloch, M., and Lambeck, K., 1995, High-precision U-series dating 1021 of corals from Western Australia and implications for the timing and duration of the 1022 Last Interglacial: Earth and Planetary Science Letters, v. 135, no. 1-4, p. 115-130.

A50

Appendices

1023 Strasser, A., 1986, Ooids in Purbeck limestones (lowermost Cretaceous) of the Swiss and 1024 French Jura: Sedimentology, v. 33, no. 5, p. 711-727. 1025 Summons, R., Bird, L., Gillespie, A., Pruss, S., Roberts, M., and Sessions, A., 2013, Lipid 1026 biomarkers in ooids from different locations and ages: evidence for a common bacterial 1027 flora: Geobiology, v. 11, no. 5, p. 420-436. 1028 Suosaari, E. P., Reid, R. P., Araujo, T. A. A., Playford, P. E., Holley, D. K., McNamara, K. J., 1029 and Eberli, G. P., 2016, Environmental Pressures Influencing Living Stromatolites In 1030 Hamelin Pool, Shark Bay, Western Australia: Palaios, v. 31, no. 10, p. 483-496. 1031 Theisen, C. H., and Sumner, D. Y., 2016, Thrombolite fabrics and origins: influences of diverse 1032 microbial and metazoan processes on Cambrian thrombolite variability in the Great 1033 Basin, California and Nevada: Sedimentology, v. 63, no. 7, p. 2217-2252. 1034 Turner, E. C., James, N. P., and Narbonne, G. M., 2000, Taphonomic control on microstructure 1035 in Early Neoproterozoic reefal stromatolites and thrombolites: Palaios, v. 15, no. 2, p. 1036 87-111. 1037 Van Kranendonk, M. J., 2006, Volcanic degassing, hydrothermal circulation and the 1038 flourishing of early life on Earth: A review of the evidence from c. 3490-3240 Ma rocks 1039 of the Pilbara Supergroup, Pilbara Craton, Western Australia: Earth-Science Reviews, 1040 v. 74, no. 3, p. 197-240. 1041 Verrecchia, E. P., Freytet, P., Verrecchia, K. E., and Dumont, J.-L., 1995, Spherulites in 1042 calcrete laminar crusts; biogenic CaCO 3 precipitation as a major contributor to crust 1043 formation: Journal of Sedimentary research, v. 65, no. 4a, p. 690-700. 1044 Wright, D. T., 1999, The role of sulphate-reducing bacteria and cyanobacteria in dolomite 1045 formation in distal ephemeral lakes of the Coorong region, South Australia: 1046 Sedimentary Geology, v. 126, no. 1, p. 147-157. 1047 Wright, D. T., and Wacey, D., 2005, Precipitation of dolomite using sulphate‐reducing bacteria 1048 from the Coorong Region, South Australia: significance and implications: 1049 Sedimentology, v. 52, no. 5, p. 987-1008. 1050 Wyrwoll, K.-H., Zhu, Z. R., Collins, L. B., and Hatcher, B. G., 2006, Origin of blue hole 1051 structures in coral reefs: Houtman Abrolhos, Western Australia: Journal of Coastal 1052 Research, p. 202-208. 1053

A51

Appendices

1054 FIGURES

1055 1056 Fig. 1. Geology (Department of Mines, Industry and Regulation, Government of Western 1057 Australia) and lake bathymetry (Rottnest Island Authority, Government of Western Australia) 1058 of the (A) regional setting and (B) study area also showing sample transect locations.

A52

Appendices

1059 1060 Fig. 2. Lacustrine depositional environments and lake types on Rottnest Island. Backshore 1061 areas are landward of blue lines; foreshore areas between blue to red lines; littoral areas 1062 between red to green lines; and basinal areas lakeward of green lines. (A) Lateral view of 1063 microbialite sheets on eastern foreshore of shallow ephemeral Pink Lake. (B) Aerial view of 1064 Pink Lake showing environmental distribution. (C) Lateral view of pustular microbial mat 1065 encrusting domical microbialites in foreshore area with heavy groundwater discharge (arrows) 1066 of deep perennial Lake Serpentine. (D) Aerial view of Lake Serpentine showing environmental 1067 division.

A53

Appendices

1068 1069 Fig. 3. Summary diagram of shallow ephemeral–perennial Pink Lake. From top: schematic cross section of lake; stratigraphic logs showing 1070 EPS-rich areas to the left (west) and microbialites to the right (east); and grainsize data to first change in lithology down section.

A54

Appendices

1071 1072 Fig. 4. Summary diagram of shallow perennial eastern Lake Baghdad. From top: schematic cross section of lake; stratigraphic logs showing 1073 EPS-rich areas to the left (west) containing spherulites and composite ooids; and grainsize data to first change in lithology down section.

A55

Appendices

1074 1075 Fig. 5. Summary diagram of shallow perennial western Lake Baghdad. From top: schematic cross section of lake; stratigraphic logs showing 1076 skeletal grainstone in foreshore areas and microbialites in littoral areas; and grainsize data to first change in lithology down section.

A56

Appendices

1077 1078 Fig. 6. Summary diagram of deep perennial southern Lake Serpentine. From top: schematic cross section of lake; stratigraphic logs showing 1079 bioclastic sand and oncoids in foreshore areas transitioning to microbialites in littoral areas and gypsum in basinal areas; and grainsize data to 1080 first change in lithology down section and; components to first change in lithology down section.

A57

Appendices

1081

1082 1083 Fig. 7. Skeletal grainstone and floatstone facies. (A) Cross-bedding in eroded platform in 1084 foreshore area of Government House Lake. (B) Photomicrograph showing clast-supported 1085 abraded grains in skeletal grainstone. (C) Example of skeletal floatstone in plan view showing 1086 articulated bivalves in life positions (arrows). (D) Photomicrograph of skeletal floatstone 1087 showing bivalve fragment (B), echinoid spicule (E) and rotalid foraminifera (R).

A58

Appendices

1088 1089 Fig. 8. Sediment facies. (A) Bioclastic sand–gravel and (B) light photomicrograph showing 1090 Katelysia (Sk) and lithic fragments (Lf) of Tamala Limestone. (C) Bivalve sand–gravel and 1091 (D) showing intact Katelysia scalarina valves (K), skeletal fragments (Sk) and coralline algae 1092 (Cl). (E) Lithic gravel and (F) light photomicrographs showing Katelysia scalarina valve (K) 1093 and lithic fragments (Lf) of. (G) Gastropod sand–gravel (S1D) and (H) light photomicrographs 1094 showing Batallaria estuarina (B), Hydrococcus sp. (arrow) gastropods, serpulid tubes (Sp) and 1095 miliolid foraminifera (M). (I) Peloidal silt–sand showing faecal pellets (P1), micritised grains 1096 (P2) and micropeloids (P3).

A59

Appendices

1097 1098 Fig. 9. Spherulite and coated grain facies. (A) Light photomicrographs of spherulite aggregate 1099 and (B) fused and coalescent spherulites with radial crystal orientation. (C) Light 1100 photomicrographs of composite ooid aggregates (D) exhibiting shared cortical laminae 1101 (arrows) around peloidal nuclei. (E) Foreshore area of heavy groundwater discharge in which 1102 (F) oncoids with indistinct and tapering laminae (arrows) containing skeletal debris in the 1103 nuclei were observed in thin section under a light microscope.

A60

Appendices

Gypsum

Gypsum

1104 1105 Fig. 10. Microbial mat and gypsum facies. (A) Non-lithifying flocculent mat lateral and (B) in 1106 cross-section. (C) Lithifying flocculent mat plan view and (D) cross-section photographs 1107 showing underlying microbialite. (E) Pustular mat plan and (F) cross-section photographs 1108 showing view showing pustules (Ps) and area containing purple coloration (arrow). (G) Blister 1109 mat plan view (H) cross-section photographs showing bulbous morphology and internal 1110 differentiation. (I) Nodular mat plan view and cross-section (J) photographs showing nodules 1111 (n) and internal differentiation. (K) Perforate mat in plan view with gypsum embedded in the 1112 surface. (L) Gypsum crusts that characterize the gypsum observed in association with the 1113 perforate mat, comprising aggregates of anhedral–lenticularm, shown here in plan view and 1114 cross-section.

A61

Appendices

1115 1116 Fig. 11. Thrombolite Type 1 facies. (A) Hand sample photograph showing sub-circular 1117 mesoclots (arrows) and abundant porosity. (B) Photomicrograph showing main microfabric 1118 components, including mesoclots and sparitic cement. (C) Photomicrograph of a mesoclot 1119 showing internal texture comprising aggregates of opaque peloids and isopachous rim (arrow).

A62

Appendices

1120 1121 Fig. 12. Thrombolite Type 2 and Stromatolite Type 1 facies. (A) Photograph of the hand 1122 sample showing thrombolite (T) and stromatolite (S) fabrics. AMS 14C sampling locations are 1123 shown with stars. (B) Photomicrograph of Thrombolite Type 2 showing aggregates of sub- 1124 circular features (arrows) within fibrous aragonite. (C) SEM-SE image of thrombolite from 1125 near the modern mat showing sub-circular features resembling coccoid cells (arrows). (D) 1126 SEM-SE image of area below (S) showing mineralized coccoid cells and EDS spectra 1127 exhibiting an elemental composition consistent with aragonite. (E) Photomicrograph of 1128 Stromatolite Type 1 (S) showing irregular sparitic laminae defined by darker areas (arrows). 1129 (F) Photomicrograph showing darker areas comprise microfossils/molds with dimensions 1130 consistent with the bacillus bacterial morphotypes (arrows).

A63

Appendices

1131 1132 Fig. 13. Stromatolite Type 2. (A) Photograph of hand sample showing location of Type 2 1133 Stromatolite above Type 1 Thrombolite (arrows). (B) Photomicrograph showing alternating 1134 sparitic and micritic laminae. (C) SEM-SE image of stromatolite from near the modern mat 1135 showing filament-like features (arrows). (D) SEM-SE image showing location of abundant 1136 sub-circular to elongate features within sparitic layer and beneath micrite. (E) SEM-BSE image 1137 showing a filament mold with well-defined segments (arrows). EDS spectra at circle location 1138 are consistent with aragonite.

A64

Appendices

1139 1140 Fig. 14. Facies trends in shallow lakes. (A) Foreshore samples plotted against observed 1141 foreshore groundwater discharge (GWD; 0=absent, 2=strong) and sediment input (bioclastic – 1142 peloidal). Foreshore areas with strong GWD and peloidal sediment form thick EPS layers 1143 offshore; bioclastic sediment with low–medium GWD associated with offshore microbialites. 1144 (B) Western areas of shallow lakes with strong GWD and thick sediment intervals form thick 1145 EPS in littoral. (C) Eastern areas of shallow lakes with lower permeability lithified pavement 1146 channeling groundwater into littoral areas along unconsolidated sediment underlying 1147 microbialitic biostromes (M).

A65

Appendices

1148 1149 Fig. 15. Facies and microbialite macromorphology trends in deep lakes. Above: distribution of 1150 microbialite and gypsum morphologies. Below: facies distributions. Lithified pavement 1151 typifies foreshore areas focusing groundwater into littoral areas beneath microbialites. Locally 1152 heavy groundwater discharge at the surface forms oncoids.

A66

Appendices

1153 1154 Fig. 16. Depositional models for the Rottnest Island lakes. (A) Marine phase ca. 2–5 kyr ago. (B) Modern shallow lakes. (C) Modern deep lakes.

A67

Appendices

1155 TABLES

1156 Table 1. Physical properties of study lakes, seawater and groundwater on Rottnest Island. Water level data courtesy of the Rottnest Island 1157 Authority, Government of WA. Pink Lake chemistry data from Monteiro (unpublished). Lake Serpentine chemistry data from Gillen (2016). 1158 Seawater and groundwater data collected during this study. AHD=Australian Height Datum. Water + 2- 2+ 2+ Size Date Depth Alk. Na Cl- SO4 Ca Mg Locality Type Permanence level pH Mg/Ca (ha) collected (m) meq/L mmol/L mmol/L mmol/L mmol/L mmol/L (AHD, m)

Ephemeral – Jul 2015 NR 0.77 8.0 2.4 1307.7 1353.3 67.0 28.0 153.3 5.5 Pink Lake Shallow 5.3 Perennial Jan 2016 NR 0.35 7.6 7.0 3892.0 4409.1 240.4 13.0 1078.4 83.1

Lake Jul 2015 -0.47 5.43 8.0 2.9 1772.9 1881.5 81.8 22.8 221.9 9.7 Deep Perennial 38.3 Serpentine Jan 2015 -0.96 4.94 8.1 3.6 3227.6 3213.6 160.4 29.7 371.0 12.5

Government Jul 2015 -0.48 7.82 8.0 2.7 2368.0 2738.5 25.4 23.6 287.6 12.2 Deep Perennial 55.4 House Lake Feb 2015 -1.09 7.21 7.8 2.7 3074.0 3405.5 65.3 30.4 367.8 12.1

Seawater - - - Oct 2018 -0.4 - 8.1 9.3 431.7 493.0 9.7 10.2 55.8 5.5

Groundwater - - - Oct 2018 - - 7.7 7.0 52.3 54.6 0.3 3.8 8.9 2.4

1159

A68

Appendices

1160 Table 2. Facies summary for the Rottnest Island lakes based on diagnostic components for limestone and sediment facies, morphology and fabric 1161 for microbial mats and microbialites, and morphology for gypsum. Mineral percentages by X-ray diffraction. HMC=High magnesium calcite (>5 1162 mol % MgCO3); LMC=low magnesium calcite (<5mol % MgCO3); TOC=total organic carbon; EPS=extracellular polymeric substances. Group Facies Principal components and description Sedimentary features Mineralogy Environments Figures

Skeletal Skeletal Tamala Limestone; large-scale cross bedding; Rhizoliths locally LMC 46%; Rottnest Island Fig. 7A, limestone grainstone abraded mollusk-dominant shell fragments, red HMC 28%; bedrock B bedrock algae, foraminifera, bryozoans, echinoids and aragonite quartz sand 24%; quartz 3%

Skeletal Herschell Limestone; tabular molluscan Eroded sub-circular–ovoid Aragonite Foreshore– Fig. 7C, rudstone– limestone beds up to 2.5 thick (Playford et al., domes up to 1.3 m wide 63%; Quartz littoral D floatstone 1977); friable to strongly cemented with a locally 14%; LMC micritic matrix; bivalves Katelysia scaralina 12%; HMC and Brachidontes sp. locally articulated and in 11% life positions

Group Facies Principal components Secondary components Mineralogy / Environments Figures TOC

Bioclastic Bioclastic Unabraded shell fragments (22–45%) Disarticulated bivalves (0– Aragonite Backshore– Fig. 8A, sediment sand–gravel dominated by bivalves; lithic grains <30 mm 7%) dominated by 34%; HMC foreshore B diameter (6–25%); peloids (17–25%) mostly Katelysia scalarina and 32%; LMC micritised skeletal grains; gastropods (0–9%) Irus irus; rotalids (0.9– 23%; Quartz Batallaria spp., and Cantharidus sp., with 2%) and miliolids (0– 11%. TOC=3– Hydrococcus sp. locally 0.7%) ; soil and roots 26%

A69

Appendices

locally; coralline algae up to 6%

Bivalve sand– Articulate and disarticulated bivalves (29– Serpulid tubeworms Aragonite 70– Backshore– Fig. 8C, gravel 45%) and shell fragments (3–21%) dominated locally common in 73%; HMC foreshore D by Katelysia scalarina, Brachidontes sp. and foreshore areas 12%; LMC Irus irus; lithic grains (10–36%) <20 mm 12%; Quartz diameter <4%. TOC=7–48%

Lithic gravel Lithic grains (33–70%) 2–50 mm diameter of Gastropods (0–12%) HMC 35– Backshore– Fig. 8E, skeletal grainstone or floatstone; bivalve Cantharidus sp., 39%; foreshore F fragments (22–45%) Katelysia sp.; Nassarius sp., Aragonite 24– disarticulated bivalves (0–20%) Katelysia Hydrococcus sp. and 31%; LMC scalarina, Brachidontes sp. and Irus irus with Diala sp. 10–21%; Fragnum erugatum locally Quartz <13%. TOC=4–13%

Gastropod Unabraded shell fragments (18–33%); peloids Miliolids rare (0.2%) to TOC=9–40% Foreshore– Fig. 8G, sand–gravel (6–26%) mostly micritised skeletal grains; common (8.7%); rotalids littoral of H gastropods (8–19%) Cantharidus sp., rare (0–1.1%); polychaete shallow lakes Batallaria spp., Cominella sp., Hydrococcus worm (Galeolaria sp. and Coxiella sp.; bivalves (8–11%) caespitosa) tube fragments articulated Katelysia scalarinaand Irus irus (diameter ~1 mm, length 3–6 mm) locally

A70

Appendices

Peloidal silt– Peloids (23–65%) mostly 50–250 μm in Shell fragments (1–11%); Aragonite 90– Littoral of Fig. 8I sand diameter probable faecal pellets and micro- miliolids (0–10%) locally 94%; HMC shallow lakes peloidal aggregates; ragged semi-opaque common; gastropods (0– 6%. TOC=16– material probable microbial mat fragments 7%) Batallaria spp. and 32% (22–71%) Nassarius sp. Irus irus locally Coated grains Spherulitic– Spherulites (~25%) and/or peloid-nucleated Peloids (15–43%) Aragonite 81– Littoral of Fig. oolitic sand– composite ooids (up to 73%) sharing cortical micritised skeletal grains 93%; HMC 3– shallow lakes 9A–D oolite laminae and faecal pellets 9%; Quartz <2 %

Oncolitic Irregularly laminated and elongate oncoids up Shell fragments (0–16%) Aragonite Deep lake Fig. gravel to 8 mm diameter (22–55%) and carbonate Katelysia scaralina; 78%; HMC foreshore areas 11E, F coated disarticulated Katelysia sp. valves (20– peloids (0–12%) 11%; LMC with heavy 65%) micritised skeletal grains 8%; Quartz groundwater 3%. TOC=2– discharge 13%

Group Facies External morphology and internal fabric Sedimentary features and associations Environments Figures

Microbial Flocculent Non-cohesive; yellow-brown layer 1–5 mm Stromatolite crusts (Type 2) and weakly Littoral of Fig. mats mat thick; purple layer locally <1 mm–2 mm thick; lithified gypsum crusts in eastern Lake shallow lakes 10A–D draped over gelatinous mat Baghdad; thick EPS intervals in western Lake Baghdad and Pink Lake

Pustular mat Undifferentiated cohesive; dark pigmented Encrusting sediment and rimming domal Foreshore with Fig. surface layer with a crenulated pustular depressions in littoral areas; associated with groundwater 10E, F surface; pustules 1–8 mm thick; purple and oncolitic gravel discharge to green patches within pustules

A71

Appendices

~20 cm water depth

Blister mat Planar to undulate; laminated cohesive; pink Type 1 stromatolites and Type 2 Littoral areas of Fig. layer 1 – 3 mm thick; green layer 0.2 – 1 mm thrombolites deep lakes 10G, H thick; purple layer 0.5–2 mm thick

Nodular mat Undulate–domal–conical; layered cohesive; Over gelatinous EPS and/or spherulitic Littoral; locally Fig. 12I, orange layer 1 mm thick with nodules 1–2 mm thrombolites in cones in basinal to ~2 J wide and 1–4 mm high aggregating on highs to m depth 10 mm wide and 12 mm high; green layer 0.3– 1 mm thick; purple layer 0–2 mm thickening in depressions

Perforate mat Undulate–domal and locally columnar; Gypsum crusts Basinal Fig. unlaminated and cohesive; green/grey/yellow 10K mat with stippled or 'perforated' surface

Group Facies Internal fabric External morphology Mineralogy Environments Figures

Thrombolites Type 1 Mesoclots comprising peloidal aggregates; Sheets Aragonite Foreshore– Fig. 11 mesoclots amoeboid to polylobate 1–2 mm 79%; HMC littoral of wide in packstone–grainstone; pore space 15%; Quartz shallow lakes locally infilled by fibrous aragonite 6%

Type 2 Diffuse irregularly shaped mesoclots Domes and sheets Aragonite Mostly littoral Fig. comprising mineralised coccoids in a sparitic 99%; HMC areas of deep 12A–D matrix 1% lakes; locally in shallow lakes

A72

Appendices

Stromatolites Type 1 Tapering and discontinuous alternating Domes and sheets Aragonite Fig. lensoidal laminae of sparite and micrite; 88%; HMC 12E, F skeletal grains locally; molds/body fossils of 12% bacillus bacterial morphotypes

Type 2 Alternating 0.1–0.2 mm laminae of sparite and As crusts ~1 cm thick in Aragonite Littoral of deep Fig. 13 micrite forming domes up to 10 mm high and surface undulations <1 cm 99%; HMC lakes; locally in 12 mm wide; abundant filament molds high and ~3 cm wide, 1% Lake Baghdad locally forming a East labyrinthine pattern

Group Facies Physical characteristics

Gypsum Gypsum Millimetric anhedral–lenticular gypsum granules partially aggregated into Gypsum Deep lakes granules crusts 100% between 0.7–2 m

Gypsum Sets of centimeter-thick alternating yellow-brown to green anhedral– Gypsum Basinal areas of Fig. crusts lenticular gypsum granules aggregated into crusts 100% deep lakes 10L 1163

A73

Appendices

1164 Table 3. Characteristics of coastal and non-coastal carbonate lakes mentioned in the text. NR=not recorded. Max. In situ Salinity, locality Genetic Area Water depth Mixed Lake ecology Littoral facies Basinal facies carbonate References (climate) type (ha) chemistry (m) minerals

Hyposaline

Lake Estanya Karstic 19 20 Mono- Highly Charo- and Bioturbated silts, Reworked Calcite, Morellón et (temperate) mictic alkaline macrophytes, ostracods, gastropods, littoral minor al. (2009) ostracods, calcified Chara remains sediments; dolomite gastropods and coatings black massive to laminated muds 2 Lake Clifton (semi- Coastal 33*10 3 No HCO3 and Charophytes, Thrombolites, calcareous NR Aragonite Moore et al. arid) Ca rich ostracods, mud, ostracods, (1994); gastropods, gastropods, calcified Konishi et al. fish, diatoms Chara remains, peloids (2001)

Lake Tangyanika Tectonic 33*105 1470 NR Mildly Charophytes, Calcareous silt–sand, NR High-Mg Cohen et al. (tropical) (rift) alkaline, gastropods, ooids, calcified Chara calcite (1987) Ca and fish remains, peloids, Mg rich gastropod coquinas, thrombolites

Hypersaline Lake Thetis (semi- Coastal 11 4 No Highly Diatoms, Thrombolites, mostly Calcareous Aragonite, Grey et al. arid) alkaline Coxiella marine derived bioclastic muds, shell minor (1990) sand–gravel fragments calcite

A74

Appendices

Rottnest Is. (semi- arid)

Shallow Coastal 1–75 ~2 No Highly Artemia, Thrombolites, Peloidal– Aragonite This study; alkaline, diatoms, stromatolites, spherulites, bioclastic silt– Playford et al. Ca and nematodes, ooids, gypsum granules, sand, gypsum (1977); Bunn

SO4 rich Coxiella in marine derived bioclastic (1984) ephemeral sand–gravel lakes

Deep Coastal 38–55 8 Mono- Highly Artemia, Oncoids, thrombolites, Gypsum Aragonite mictic alkaline, diatoms, stromatolites, marine Ca and nematodes derived bioclastic sand–

SO4 rich gravel

Lake Marion (semi- Coastal 30*10 2 No Highly Brine shrimp, Stromatolites, gypsum Calcareous Aragonite, Von der arid) alkaline, nematodes, granules, oncoids, mostly muds, gypsum minor Borch et al. Ca and Coxiella marine derived bioclastic high-Mg (1977); Perri

SO4 rich sediment calcite et al. (2011)

Great Salt Lake Contin- 44*104 10 Mero- Highly Brine shrimp, Mirabilite crusts, ooids, Reworked Aragonite Post (1977); (temperate) ental mictic alkaline, diatoms, oncoids, stromatolites, littoral Bouton et al. saline abundant ostracods thrombolites sediments; (2016)

NaSO4 black massive and wavy to laminated muds 1165

A75

Appendices

SUPPLEMENTARY INFORMATION

Fig. S1. Summary diagram of small ephemeral Lake Sirius. From top: schematic cross section of lake; stratigraphic logs; grainsize data to first change in lithology down section and; components to first change in lithology down section.

A76

Appendices

Fig. S2. Summary diagram of deep perennial Lake Serpentine. From top: schematic cross section of lake; stratigraphic logs showing bioclastic sand and oncolitic gravel in foreshore areas transitioning to microbialites in littoral and gypsum in basinal areas; grainsize data to first change.

A77

Appendices

Fig. S3. Summary diagram of deep perennial Government House Lake. From top: schematic cross section of lake; stratigraphic logs showing skeletal floatstone in foreshore transitioning to microbialites in littoral and gypsum in basinal areas; grainsize data to first change in lithology down section and; components to first change in lithology down section.

A78