post-glacial sea-level changes around the australian

74
University of Wollongong Research Online Faculty of Science, Medicine and Health - Papers Faculty of Science, Medicine and Health 2013 Post-glacial sea-level changes around the Australian margin: a review Stephen E. Lewis James Cook University Craig R. Sloss Queensland University of Technology, [email protected] Colin V. Murray-Wallace University of Wollongong, [email protected] Colin D. Woodroffe University of Wollongong, [email protected] Sco G. Smithers James Cook University Research Online is the open access institutional repository for the University of Wollongong. For further information contact the UOW Library: [email protected] Publication Details Lewis, S. E., Sloss, C. R., Murray-Wallace, C. V., Woodroffe, C. D. & Smithers, S. G. (2013). Post-glacial sea-level changes around the Australian margin: a review. Quaternary Science Reviews, 74 115-138.

Upload: others

Post on 21-May-2022

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Post-glacial sea-level changes around the Australian

University of WollongongResearch Online

Faculty of Science, Medicine and Health - Papers Faculty of Science, Medicine and Health

2013

Post-glacial sea-level changes around the Australianmargin: a reviewStephen E. LewisJames Cook University

Craig R. SlossQueensland University of Technology, [email protected]

Colin V. Murray-WallaceUniversity of Wollongong, [email protected]

Colin D. WoodroffeUniversity of Wollongong, [email protected]

Scott G. SmithersJames Cook University

Research Online is the open access institutional repository for the University of Wollongong. For further information contact the UOW Library:[email protected]

Publication DetailsLewis, S. E., Sloss, C. R., Murray-Wallace, C. V., Woodroffe, C. D. & Smithers, S. G. (2013). Post-glacial sea-level changes around theAustralian margin: a review. Quaternary Science Reviews, 74 115-138.

Page 2: Post-glacial sea-level changes around the Australian

Post-glacial sea-level changes around the Australian margin: a review

AbstractIt has been known since Rhodes Fairbridge's first attempt to establish a global pattern of Holocene sea-levelchange by combining evidence from Western Australia and from sites in the northern hemisphere that thedetails of sea-level history since the Last Glacial Maximum vary considerably across the globe. The Australianregion is relatively stable tectonically and is situated in the 'far-field' of former ice sheets. It therefore preservesimportant records of post-glacial sea levels that are less complicated by neotectonics or glacio-isostaticadjustments. Accordingly, the relative sea-level record of this region is dominantly one of glacio-eustatic (iceequivalent) sea-level changes. The broader Australasian region has provided critical information on the natureof post-glacial sea level, including the termination of the Last Glacial Maximum when sea level wasapproximately 125 m lower than present around 21,000-19,000 years BP, and insights into meltwater pulse 1Abetween 14,600 and 14,300 cal. yr BP. Although most parts of the Australian continent reveals a high degreeof tectonic stability, research conducted since the 1970s has shown that the timing and elevation of aHolocene highstand varies systematically around its margin. This is attributed primarily to variations in thetiming of the response of the ocean basins and shallow continental shelves to the increased ocean volumesfollowing ice-melt, including a process known as ocean siphoning (i.e. glacio-hydro-isostatic adjustmentprocesses). Several seminal studies in the early 1980s produced important data sets from the Australasianregion that have provided a solid foundation for more recent palaeo-sea-level research. This review revisitsthese key studies emphasising their continuing influence on Quaternary research and incorporates relativelyrecent investigations to interpret the nature of post-glacial sea-level change around Australia. These include asynthesis of research from the Northern Territory, Queensland, New South Wales, South Australia andWestern Australia. A focus of these more recent studies has been the re-examination of: (1) the accuracy andreliability of different proxy sea-level indicators; (2) the rate and nature of post-glacial sea-level rise; (3) theevidence for timing, elevation, and duration of mid-Holocene highstands; and, (4) the notion of mid- to lateHolocene sea-level oscillations, and their basis. Based on this synthesis of previous research, it is clear thatestimates of past sea-surface elevation are a function of eustatic factors as well as morphodynamics ofindividual sites, the wide variety of proxy sea-level indicators used, their wide geographical range, and theirindicative meaning. Some progress has been made in understanding the variability of the accuracy of proxyindicators in relation to their contemporary sea level, the inter-comparison of the variety of dating techniquesused and the nuances of calibration of radiocarbon ages to sidereal years. These issues need to be thoroughlyunderstood before proxy sea-level indicators can be incorporated into credible reconstructions of relative sea-level change at individual locations. Many of the issues, which challenged sea-level researchers in the latterpart of the twentieth century, remain contentious today. Divergent opinions remain about: (1) exactly whensea level attained present levels following the most recent post-glacial marine transgression (PMT); (2) theelevation that sea-level reached during the Holocene sea-level highstand; (3) whether sea-level fell smoothlyfrom a metre or more above its present level following the PMT; (4) whether sea level remained at thesehighstand levels for a considerable period before falling to its present position; or (5) whether it underwent aseries of moderate oscillations during the Holocene highstand.

Keywordsaustralian, around, margin, changes, level, review, glacial, post, sea, GeoQuest

DisciplinesMedicine and Health Sciences | Social and Behavioral Sciences

This journal article is available at Research Online: http://ro.uow.edu.au/smhpapers/921

Page 3: Post-glacial sea-level changes around the Australian

Publication DetailsLewis, S. E., Sloss, C. R., Murray-Wallace, C. V., Woodroffe, C. D. & Smithers, S. G. (2013). Post-glacial sea-level changes around the Australian margin: a review. Quaternary Science Reviews, 74 115-138.

This journal article is available at Research Online: http://ro.uow.edu.au/smhpapers/921

Page 4: Post-glacial sea-level changes around the Australian

1

1

2

3

4

5

6

Post-glacial sea-level changes around the Australian margin: A review 7

8 Stephen E. Lewis1*, Craig R. Sloss2, Colin V. Murray-Wallace3, Colin D. Woodroffe3, Scott 9

G. Smithers4 10 11

1 Catchment to Reef Research Group, TropWATER, James Cook University, Townsville 12 QLD 4811, Australia 13 2 Earth, Environmental and Biological Sciences, Queensland University of Technology, GPO 14 Box 2434, Brisbane, QLD 4001, Australia 15 3 School of Earth & Environmental Sciences, University of Wollongong, Wollongong, NSW 16 2522, Australia 17 4 School of Earth and Environmental Sciences, James Cook University, Townsville QLD 18 4811, Australia 19 * Corresponding author: phone: +61 7 4781 6629; fax: +61 7 4781 5589 email: 20 [email protected] 21 22

23

24

25

26

27

28

29

30

31

32

33

34

35

36

37

Page 5: Post-glacial sea-level changes around the Australian

2

Abstract 38

It has been known since Rhodes Fairbridge’s first attempt to establish a global pattern of 39

Holocene sea-level change by combining evidence from Western Australia and from sites in 40

the northern hemisphere that the details of sea-level history since the last glacial maximum 41

vary considerably across the globe. The Australian region is relatively stable tectonically and 42

is situated in the ‘far-field’ of former ice sheets. It therefore preserves important records of 43

post-glacial sea levels that are less complicated by neotectonics or glacio-isostatic 44

adjustments. Accordingly, the relative sea-level record of this region is dominantly one of 45

glacio-eustatic (ice equivalent) sea-level changes. The broader Australasian region has 46

provided critical information on the nature of post-glacial sea level, including the termination 47

of the last glacial maximum when sea level was approximately 125 m lower than present 48

around 21,000 to 19,000 years BP, and insights into meltwater pulse 1A between 14,600 and 49

14,300 cal. yr BP. Although most parts of the Australian continent reveals a high degree of 50

tectonic stability, research conducted since the 1970’s has shown that the timing and 51

elevation of a Holocene highstand varies systematically around its margin. This is attributed 52

primarily to variations in the timing of the response of the ocean basins and shallow 53

continental shelves to the increased ocean volumes following ice-melt, including a process 54

known as ocean siphoning (i.e. glacio-hydro-isostatic adjustment processes). 55

56

Several seminal studies in the early 1980’s produced important datasets from the Australasian 57

region that have provided a solid foundation for more recent palaeo-sea-level research. This 58

review revisits these key studies emphasising their continuing influence on Quaternary 59

research and incorporates relatively recent investigations to interpret the nature of post-60

glacial sea-level change around Australia. These include a synthesis of research from the 61

Northern Territory, Queensland, New South Wales, South Australia and Western Australia. A 62

focus of these more recent studies has been the re-examination of: 1) the accuracy and 63

reliability of different proxy sea-level indicators; 2) the rate and nature of post-glacial sea-64

level rise; 3) the evidence for timing, elevation, and duration of mid-Holocene highstands; 65

and, 4) the notion of mid to late-Holocene sea-level oscillations, and their basis. 66

67

Based on this synthesis of previous research it is clear that estimates of past sea-surface 68

elevation are a function of eustatic factors as well as morphodynamics of individual sites, the 69

wide variety of proxy sea-level indicators used, their wide geographical range, and their 70

indicative meaning. Some progress has been made in understanding the variability of the 71

Page 6: Post-glacial sea-level changes around the Australian

3

accuracy of proxy indicators in relation to their contemporary sea level, the inter-comparison 72

of the variety of dating techniques used and the nuances of calibration of radiocarbon ages to 73

sidereal years. These issues need to be thoroughly understood before proxy sea-level 74

indicators can be incorporated into credible reconstructions of relative sea-level change at 75

individual locations. Many of the issues which challenged sea-level researchers in the latter 76

part of the twentieth century remain contentious today. Divergent opinions remain about: 1) 77

exactly when sea level attained present levels following the most recent post-glacial marine 78

transgression (PMT); 2) the elevation that sea-level reached during the Holocene sea-level 79

highstand; 3) whether sea-level fell smoothly from a metre or more above its present level 80

following the PMT; 4) whether sea level remained at these highstand levels for a considerable 81

period before falling to its present position; or 5) whether it underwent a series of moderate 82

oscillations during the Holocene highstand. 83

84

85

86

Keywords: sea level; Australia; sea-level indicators; sea-level change; Holocene; post-glacial 87

88

89

90

91

92

93

94

95

96

97

98

99

100

101

102

103

104

105

Page 7: Post-glacial sea-level changes around the Australian

4

1. Introduction 106

The relative tectonic stability of the Australian continent is due to its intra-plate setting. It is 107

also far from the major ice sheets that covered large areas of the northern hemisphere 108

continents during the Last Glacial Maximum (LGM) and has thus been unaffected by 109

significant ice accumulation or the effects of glacio-isostatic rebound. Accordingly, the 110

Australian margin is a suitable place to examine evidence of former shorelines dominated by 111

glacio-eustatic sea-level changes with variable influences of hydro-isostasy (Bryant et al., 112

1992; Nakada and Lambeck, 1989; Murray-Wallace and Belperio, 1991). The relative 113

tectonic stability and limited isostatic influence around the Australian coastal margin renders 114

the region an important setting for reconstructing post-glacial glacio-eustatic (ice-equivalent) 115

sea-level changes during the most recent marine transgression which followed the LGM 116

(Ferland et al., 1995; Murray-Wallace et al., 1996, 2005). However, much of the Australian 117

coast is bounded by a relatively shallow continental shelf of variable width that has 118

experienced limited hydro- and sedimentary isostasy during the last glacial cycle which has 119

had an influence in terms of the Holocene highstand (Larcombe et al., 1995; Baker and 120

Haworth, 2000a; Belperio et al., 2002; Collins et al., 2006; Sloss et al., 2007; Lewis et al., 121

2008; Woodroffe, 2009). In 1983, David Hopley and colleagues published a synthesis of 122

Holocene sea-level data from around the Australian mainland highlighting the key sites and 123

studies up to that time (Hopley, 1983a). This review reflects on the continued influence of the 124

key sites identified in Hopley’s synthesis, and critically reviews the palaeo-sea level evidence 125

presented in subsequent studies. 126

127

While the Australian continental margin reveals evidence for localised neotectonic uplift and 128

subsidence (up to 10 m but typically < 2 m) based on the displacement of last interglacial 129

(marine isotope sub-stage 5e ca. 125 ka) coastal successions that clearly relate to different 130

geotectonic domains (Murray-Wallace, 2002), these rates of crustal movement have had 131

negligible influence on resolving the general pattern of Holocene sea-level changes. For 132

example, one of the more tectonically active regions in southern Australia has experienced 133

uplift of approximately 0.7 m during the Holocene (Cann et al., 1999). This contrasts with 134

adjacent plate-margin sites such as the Huon Peninsula in Papua New Guinea which has 135

undergone rapid and apparently constant tectonic uplift of 20 – 30 m over the same period, 136

where one of the most detailed records of relative sea-level changes for the past ~ 400 ka has 137

been established (e.g. Chappell and Shackleton, 1986; Chappell et al., 1996). 138

139

Page 8: Post-glacial sea-level changes around the Australian

5

The analysis in this review uses a similar approach to previous studies; it emphasises the 140

continued significance of the previously published data (particularly age estimates and the 141

reliability of proxy data to contemporary sea levels) and presents additional marine (corals) 142

and estuarine (sedimentary and mangrove deposits) proxy indicators, together with other 143

proxies now being examined using new techniques (e.g. fixed intertidal biological indicators 144

or encrusting organisms). This work also examines the principal research trends over the past 145

25 years that recognised regional and local tectonic influences around the Australian margin, 146

and regional variations in eustatic influences, leading to the reconstruction of sea-level 147

histories on a more localised scale. Improved methods of dating enable a greater precision in 148

the assignment of ages to fossil material, but the accuracy with which any inference about sea 149

level can be made means that each sample must be interpreted in the context of the 150

geomorphological setting in which it occurred. Indeed the review highlights the potential 151

problems associated with the interpretation of sea-level indicators, most notably the 152

assumptions that coastal boundary morphodynamic conditions have been constant through 153

time. 154

155

The key sites with ongoing research on post-glacial sea-level changes examined in this 156

review include (Fig. 1): The Huon Peninsula, Papua New Guinea (Chappell and Polach, 157

1991; Ota and Chappell, 1999) and the Joseph Bonaparte Gulf (Yokyama et al. 2000, 2001a) 158

and its comparison with the Sunda Shelf (Hanebuth et al. 2000, 2009); Spencer Gulf and the 159

Gulf of St Vincent (Belperio et al., 2002); eastern Australia (Baker et al., 2001a, b; Sloss et 160

al., 2007; Lewis et al., 2008); and Western Australia (Searle and Woods, 1986; Baker et al., 161

2005; Collins et al., 2006). To accurately examine the sea-level histories following the LGM, 162

this review synthesises available data for sea-level change around the Australian margin over 163

the past 20,000 years. We focus on two separate intervals: 1) during the post-glacial 164

transgression (20,000 to 7000 years BP); and 2) the mid-late Holocene (8000 years BP to 165

present). We examine the key mechanisms that appear to explain sea-level change since the 166

LGM at regional and local scales, recognising the legacy provided by the pioneering 167

researchers and identifying remaining gaps in our understanding. 168

169

2. Background 170

The significance of both the relatively tectonically stable continent of Australia and the 171

uplifted coast of Papua New Guinea for palaeo-sea-level investigations was realised by 172

Rhodes W. Fairbridge. His studies in Western Australia in collaboration with Curt Teichert 173

Page 9: Post-glacial sea-level changes around the Australian

6

(Teichert, 1950) identified a series of well-preserved sea-level indicators on the Western 174

Australian coast at Point Peron, and offshore on Rottnest Island and the Abrolhos Islands. 175

Fairbridge also realised the potential of the uplifted suite of reef terraces on the north coast of 176

the Huon Peninsula (Fairbridge, 1960). In 1961, Fairbridge presented a thorough global 177

synthesis and critique of sea level and climatic studies and painstakingly considered (and 178

discounted) several alternative explanations for sea-level variability since the LGM including 179

continental flexure, the ‘oscillating margin’ hypothesis, geodetic forcing and different forms 180

of eustasy (listed below). He produced an integrated theory based on changes in ice volume 181

(glacio-eustasy), tectonic influences (tectono-eustasy), sediment deposition in ocean basins 182

(sedimento-eustasy) and other influences such as inputs from volcanic activity and thermo-183

expansion/contraction. Fairbridge (1961) also recognised the influence of local tectonic 184

processes and crustal adjustments to water loading and unloading (hydro-isostasy) on sea-185

level records, although he did not attempt to quantify them in his work. 186

187

Fairbridge’s sea-level curve of global eustasy (Fairbridge, 1961), developed from a series of 188

studies undertaken in the 1940’s and 1950’s with Teichert, propelled the evidence from 189

emergent shorelines in Western Australia onto the world stage as key type localities for 190

several highstands. At this time the prevailing view was that sea level had risen gradually up 191

to present at a decelerating rate (e.g. Jelgersma and Pannekoek, 1960; Shepard, 1961). In 192

light of that, Fairbridge included post-glacial sea-level evidence from the Bahamas and 193

Florida positioned below present sea level. In Fairbridge’s eustatic curve, a series of 194

highstands based on evidence from Australian sites were interspersed with lowstands based 195

on evidence from sites in the Bahamas and Florida. It soon became apparent that these 196

Atlantic and Pacific data sets could not be compiled into a single eustatic sea-level record. 197

This was due to global and regional variations in the eustatic and isostatic response to 198

changes in ice and ocean volumes, resulting in different relative sea-level histories (Bloom, 199

1977). 200

201

In Australia, Bruce Thom and John Chappell (1975) produced the first widely accepted sea-202

level envelope of post-glacial sea-level rise for Australia based on a compilation of 203

radiocarbon ages from eastern Australia. Their study also suggested the possibility of mid-204

late Holocene emergence on some parts of the Australian coast, and recognised the 205

significance of a stratigraphic record of prograded coastal plains that had accumulated over 206

the past 6000 years. This was in contrast to the decelerating pattern of sea-level rise widely 207

Page 10: Post-glacial sea-level changes around the Australian

7

recognised in the North Atlantic, where present sea level was not attained until sometime 208

between 4000 to 2000 yr BP (Thom and Chappell, 1975; Pirazzoli, 1991). Whether or not 209

Holocene sea levels had been higher than present, particularly along the eastern Australian 210

coast remained a contentious issue. A vigorous exchange of views had been expressed in the 211

early volumes of Marine Geology, with evidence proposed in support of sea level having 212

been above its present level observed in Queensland and Victoria (Gill and Hopley, 1972) 213

and refuted along other sections of the coast (Thom et al., 1969, 1972). 214

215

In 1973, the Great Barrier Reef Expedition by the Royal Society of London and the 216

Universities of Queensland, led by David Stoddart, undertook particularly extensive mapping 217

and surveying of numerous reefs in the northern Great Barrier Reef. The results, published in 218

1978 in two volumes of the Philosophical Transactions of the Royal Society, provided 219

unequivocal evidence for higher Holocene sea levels (McLean et al., 1978). It was as part of 220

this expedition that the significance of coral microatolls as accurate sea-level indicators was 221

realised (Scoffin and Stoddart, 1978). Shortly thereafter an extensive sequence of microatolls 222

from the continental margin of north-east Queensland was described by John Chappell and 223

colleagues (Chappell, 1983; Chappell et al., 1982, 1983). The two decades following saw 224

concentrated coring and dating of a series of reefs which contributed to our understanding of 225

the way in which the Great Barrier Reef had developed during the Holocene in response to 226

sea-level variations. These are summarised by Hopley (1982) and updated in Hopley et al. 227

(2007). 228

229

Stratigraphical and chronological studies of sand barrier and estuarine evolution along the 230

south-east coasts of Australia by Bruce Thom and Peter Roy refined the understanding of sea-231

level history in this region (Roy and Thom, 1981; Thom, 1984a, b) and consolidated the 232

earlier work of Thom and Chappell (1975). Thom and Roy (1983, 1985) subsequently plotted 233

a sea-level envelope between the terrestrial (tree stumps, wood and charcoal) and marine 234

(shells) palaeo-sea-level indicators recognising that these were ‘directional’ (relational) 235

indicators, rather than providing an accurate (fixed) estimate of sea-level position. As a result, 236

these sea-level data display a wide scatter of sample depth in relation to their contemporary 237

sea level. The samples collected for radiocarbon dating were collected from locations 238

extending from the Gold Coast in Queensland to Badger Head in Tasmania, representing an 239

extensive geographical range (1,400 km in latitude) along Australia’s eastern seaboard. The 240

additional data incorporated by Thom and Roy (1983, 1985), notably the mangrove stump 241

Page 11: Post-glacial sea-level changes around the Australian

8

data, shifted the sea-level envelope to attain present levels much earlier than the previous 242

work by Thom and Chappell (1975) thus showing that the rates of sea-level rise were faster 243

than previously thought (we note that Thom and Roy also used ‘environmentally corrected’ 244 14C ages for the marine reservoir effect in radiocarbon dating for their reconstruction). These 245

studies laid the groundwork for the recognition that local sea levels are partly dependent on 246

the geomorphological setting, and that geomorphological evolution of the coast may both 247

preserve evidence of relative sea-level change and influence how other sea-level indicators 248

preserve evidence of past sea-level changes. Nevertheless, the data allowed the construction 249

of a broad envelope to constrain the position of early to mid-Holocene relative sea level for 250

the east coast of Australia, indicating that sea level reached its present height by 6000 251

radiocarbon yr BP. 252

253

The focused research on sea level during the late 1970’s and early 1980’s enabled David 254

Hopley and colleagues to compile a comprehensive data set of Holocene sea-level indicators 255

across Australia in 1983 (see Hopley, 1983a, b, 1987; Hopley and Thom, 1983) with a further 256

compilation by John Chappell in 1987 (Chappell, 1987). Critical issues identified in these 257

compilations included the determination of when following the post-glacial marine 258

transgression present sea level was reached, the elevation and duration of any mid-Holocene 259

highstand, whether sea-level oscillations of 1 to 2 m lasting up to 1000 years have occurred 260

since present sea level was initially reached, and understanding regional variations in 261

Holocene relative sea-level histories around the Australian coastline and across other parts of 262

the world (Hopley, 1983a, 1987; Chappell, 1987). 263

264

3. Sea-level indicators 265

A variety of proxy sea-level indicators has been used in previous studies and on-going 266

research and a critical assessment of their indicative meaning is presented here. Interpreting 267

past sea levels has involved the use of several different sea-level indicators to reconstruct sea-268

level change relative to a specific datum (commonly Australian Height Datum, which broadly 269

approximates estimated present mean sea level (PMSL)). In some parts of the world, there is 270

clear evidence that the sea was once higher, preserved as erosional notches cut into cliffs. 271

Erosional features formed a substantial component of the evidence for higher sea level 272

recognised by Fairbridge and others from Western Australia. However, constructional coasts, 273

such as coral reefs, built by a framework of coral and with a range of secondary organisms, 274

sedimentary and diagenetic processes, tend to preserve a wider range of potential indicators 275

Page 12: Post-glacial sea-level changes around the Australian

9

than erosional coasts. On other coasts, stratigraphical sequences of sedimentary facies 276

(commonly dating intertidal and subtidal shells) have provided a first-order indication of 277

transgressive or regressive tendency and function as directional indicators of previous 278

limiting sea levels. Further sea-level evidence has been derived from beachrock, the remains 279

of intertidal plants (mangroves), encrusting organisms (oysters, barnacles, tubeworms), 280

foraminiferal assemblages and supratidal deposits, as well as intertidal erosional indicators 281

(including notches, abrasion platforms and benches). In this section, we review the 282

development, reliability, accuracy and limitations of the key proxy indicators that are 283

commonly used to reconstruct palaeo-sea-level. A detailed review of sea-level indicators on 284

coral reefs has recently been provided by Smithers (2011 and references therein) and other 285

reviews including Hopley and Thom (1983), Davies and Montaggioni (1985), van de 286

Plassche (1986), Pirazzoli, (1991, 1996) and Hopley et al. (2007) also provide excellent 287

broader overviews. This section, therefore, provides only a brief summary of commonly used 288

proxy indicators, and how they have been used in reconstructions in the Australian region. 289

We examine the integrity and indicative meaning of each of these proxies and highlight their 290

contribution to reconstructing relative sea levels around the Australian mainland. While the 291

sea-level indicators described below have different indicative meanings relative to predicted 292

tidal datums (e.g. corals: below or at mean low water springs; tubeworms: mean high water 293

neaps), the difference between the elevation of the relict indicator and that of the highest or 294

lowest living counterpart provides evidence of relative sea-level change compared to PMSL. 295

296

It is important to recognise that the indicators above preserve sea-level evidence that may be 297

related to a level associated with specific coastal features or processes, but may never be 298

directly tied to a surveyed and statistically derived tidal datum such as mean sea level (MSL). 299

Biological indicators are mediated by the physiological response to environmental factors 300

such as exposure, which may relate broadly to a tidal stage. However, tidal planes themselves 301

are unlikely to be horizontal over even short distances, are variable in time and space due to 302

interactions with local geomorphology, and open waters are rarely calm and still. These at-a-303

site factors which introduce variability and noise are less critical for assessments of rapidly 304

rising post-glacial sea levels, but become pivotal in the interpretation of the details of sea-305

level histories after present sea level was initially reached. Indeed, a change in relative sea 306

level as recorded by sea-level indicators can be the result of eustatic (ice volume change), 307

isostatic (uplift or subsidence from changing ice/water loads on the continental crust), 308

Page 13: Post-glacial sea-level changes around the Australian

10

climatic (steric, wind forcing) or coastal morphodynamic (e.g. closing of a high energy 309

window) changes. 310

311

The reliability with which indicators can be used in sea-level reconstructions is also partly 312

dependent on their potential for accurate age-control (Sloss et al., in press). Dating can be 313

performed by various methods including radiocarbon (via Accelerator Mass Spectrometry: 314

AMS or conventional methods), U-series, Amino Acid Racemisation (AAR) and Optical 315

Stimulated Luminescence (OSL). Radiocarbon (14C) dating has unquestionably been the most 316

widely applied numeric dating method in studies of post-glacial sea-level changes; however, 317

ages need to be calibrated to account for secular atmospheric variations and the marine 318

reservoir effects for marine samples (both global and regional; Sloss et al., in press). The 14C 319

ages presented in this review have been calibrated to sidereal years with 2 sigma standard 320

deviations using the latest calibration programs and applying the currently recommended 321

regional reservoir effect (ΔR) for each site (Stuiver et al., 2005; Weninger et al., 2011). 322

Previous studies have shown that ΔR varies around different sections of the Australian 323

coastline and also with the type of indicator (e.g. seagrass fibres, bivalves, corals); it is also 324

likely to diverge from the open-water ΔR value in estuaries where there is significant dilution 325

by fluvially-derived water with a different radiocarbon activity, as well as having varied 326

considerably over time (Eisenhauer et al., 1993; Ulm, 2002; McGregor et al., 2008; Lewis et 327

al., 2012). Insufficient data are currently available to account for these additional influences 328

on ΔR, but compilation of age data and continued attention to disparities in the ΔR value 329

promise to reduce discrepancies that arise from such issues. Before reliable radiocarbon 330

calibration programs, ‘environmentally corrected’ 14C ages where 450 years were subtracted 331

from marine samples (commonly molluscs and corals) following the work of Gillespie and 332

Polach (1979) to take account of the ‘pre-inclusion’ age of radiocarbon were commonly 333

applied in sea-level reconstructions. The development of calibration datasets (e.g. Reimer et 334

al., 2009) allows 14C ages to be calibrated to sidereal years and this advance has shown that 335

present sea level around Australia was reached somewhat earlier than previously thought (e.g. 336

Sloss et al., 2007; Lewis et al., 2008). U-series has become the favoured method for the 337

dating of corals as the ages do not require corrections for the marine reservoir effects and the 338

analytical precision has greatly improved with the establishment of new analytical procedures 339

(Yu and Zhao, 2010; Clark et al., 2012). 340

341

3.1 Coral reefs 342

Page 14: Post-glacial sea-level changes around the Australian

11

Reef-building scleractinian corals flourish in shallow tropical waters, limited in their 343

poleward extension by sea-surface temperatures (coral reef growth is mostly confined above 344

the 18ºC isotherm), constrained in their upward growth by exposure at low water spring tide, 345

and rare below water depths of 50 m where symbiotic photosynthetic algae receive 346

insufficient light (Smithers, 2011). Individual fossil corals within Holocene coral reefs can be 347

used as ‘directional’ sea-level indicators where relative sea-level position (compared with 348

today) must have been higher at the time of coral growth (e.g. Veeh and Veevers, 1970). 349

Individual massive or ‘brain’ corals grow at rates of about 1 to 2 cm per year, whereas many 350

branching corals may extend 5-10 cm each year. Coral reefs, however, which are the 351

aggregated outcome of a framework of corals and other calcifying reef organisms, tempered 352

by erosion, accrete vertically at rates of just a few millimetres per year. Accordingly, they 353

grew more slowly than rapidly rising sea levels following the LGM, with the result that the 354

reef growth curve differs from the sea-level curve (Hopley 1986a; Eisenhauser et al., 1993). 355

Three modes of coral reef response to sea-level rise have been identified: keep-up where a 356

reef tracked sea-level rise; catch-up where a reef lagged behind the rising sea level but 357

reached the sea surface after sea level had stabilised close to its present position; and give-up 358

where a reef was drowned by rapid sea-level rise and remained as a submerged reef, 359

commonly with an algal pavement over its surface (summarised in Hopley, 1982; Neumann 360

and Macintyre, 1985; Davies and Montaggioni, 1985). 361

362

Recently it has become apparent that there are numerous drowned, give-up reefs in the 363

Australian region. ‘Drowned reefs’ along the shelf edge of the Great Barrier Reef have been 364

reported at depths between 50 and 70 m (Carter and Johnson, 1986; Harris and Davies, 1989; 365

Beaman et al., 2008; Abbey et al., 2011; Webster et al., 2011; Yokoyama et al., 2011); these 366

reefs are likely to provide important sea-level data, although they have yet to be dated. 367

Drowned reefs at depths of 20 to 30 m have also been identified in the southern Gulf of 368

Carpentaria, where they were unexpected because of relatively high turbidity and the fact that 369

the area was a terrestrial environment with an inland lake for a significant part of the latter 370

portion of the last glacial cycle (Chivas et al., 2001). Terrace reefal successions from the 371

Cootamundra Shoals, north-western Australia ranging from 20 to 90 m depth have also been 372

identified with radiocarbon ages of the 20 and 28 m terraces yielding 8460 ± 100 and 7630 ± 373

100 cal. yr BP, respectively (Flemming, 1986). 374

375

Page 15: Post-glacial sea-level changes around the Australian

12

Much has been learnt about the internal structure of reefs from detailed studies of the 376

tectonically-active margin of the Huon Peninsula, Papua New Guinea where coral reef 377

terraces have been uplifted and preserved since at least 400,000 years ago (Chappell, 1980; 378

Chappell et al., 1996; Pandolfi, 1996; Pandolfi et al., 2006). Where rates of tectonic uplift are 379

well-constrained, these terrace successions reliably record previous interglacial and 380

interstadial highstands (and possibly lowstands), but coring on the lowermost terrace has also 381

provided important data on sea-level rise after the LGM (Chappell and Polach, 1991; 382

Chappell et al., 1998; Ota and Chappell, 1999). 383

384

Submerged reefs may also be used as indicators where the rate of subsidence is known, and 385

where reef flats and reef pavement deposits are located above the contemporary limit to coral 386

growth they can indicate formerly higher sea levels (Hopley, 1986a; Matthews, 1990; Collins 387

et al., 2006). At the Houtman Abrolhos Islands, Western Australia raised coral pavements 388

were used to reconstruct Holocene sea levels with an accuracy of ±0.5 m (Collins et al., 389

2006), where they were interpreted to form during gradual sea-level fall (Eisenhauer et al., 390

1999). Corals grow prolifically in shallow water with a gradually rising sea, but may be 391

almost absent if a mature reef flat becomes emergent because of a relative sea-level fall 392

(Smithers et al., 2006). 393

394

3.2 Coral microatolls 395

Microatolls are disc-shaped coral colonies with dead tops and live sides. They form on reef 396

flats and can grow up to several metres in diameter. Colony height varies with depth of the 397

substrate, but for most microatolls used for sea-level research on the Great Barrier Reef it is < 398

1 m. As many as 23 coral genera have been observed to adopt a microatoll growth form on 399

the Great Barrier Reef, but the most common microatolls are of Porites and Faviid spp. 400

(Rosen, 1978; Scoffin and Stoddart, 1978; Stoddart and Scoffin, 1979). Massive Porites grow 401

upwards in open water conditions until they reach an elevation close to mean low water 402

springs (MLWS) or slightly higher, after which they cannot grow further vertically, but 403

continue to extend laterally to form microatolls (Scoffin and Stoddart, 1978). Although the 404

MLWS limit has been observed across much of the GBR for Porites, Faviid microatolls have 405

been found up to 0.5 m higher (Perry and Smithers, 2011) and Porites microatolls on the 406

Cocos (Keeling) Islands also grow to a higher level in open waters (between MLWS and 407

mean low water neap tides; Smithers and Woodroffe, 2000). Microatolls can be locally 408

moated behind rubble storm ridges (Hopley and Isdale, 1977), thus indicating a locally-409

Page 16: Post-glacial sea-level changes around the Australian

13

elevated water level (discussed below), which can complicate sea-level interpretations where 410

moated and open water microatolls are not easily distinguished. 411

412

Although flat-topped corals were described from Yule Point by Bird (1971), fossil 413

microatolls elevated above their living counterparts were first recognised as high resolution 414

sea-level indicators on the Great Barrier Reef by Scoffin and Stoddart (1978). Dated 415

microatolls were an important component of the evidence used by McLean et al. (1978) to 416

provide an estimate of a higher sea level (approximately +1.0 m) in the mid-Holocene. An 417

estimate of when sea level first attained its modern elevation was given at 6310 ± 90 14C 418

years for a microatoll on Fisher Island (6760 ± 230 cal. yr BP; McLean et al., 1978). For this 419

particular study, the elevations of the fossil microatolls were measured conservatively to the 420

current mean high neap tidal mark to account for possible ‘moating effects’. 421

422

The most comprehensive sea-level reconstruction using coral micoatolls was performed by 423

Chappell et al. (1983) where groups or ‘fields’ of raised fossil microatolls were measured 424

relative to their modern counterparts on fringing reefs along the continental islands and 425

mainland of the Great Barrier Reef (‘moated’ microatolls were excluded from the analysis). 426

These data remain as one of the highest resolution records produced in Australia and have 427

been used to ‘calibrate’ hydro-isostatic models and support the ‘smooth-fall’ hypothesis of 428

Holocene sea levels following the onset of the highstand (e.g. Chappell et al., 1982). Baker 429

and Haworth (2000b) statistically reanalysed Chappell et al.’s (1983) original data and 430

suggested that either the linear fall trend or a seventh-order polynomial (oscillating) trend 431

could fit the data equally well. 432

433

Yu and Zhao (2010) demonstrated the higher analytical precision of U-series methods by 434

analysing 11 coral microatolls from Magnetic Island obtaining 2σ errors commonly within 435

±50 years. However, neither the studies of Chappell et al. (1983) nor those of Yu and Zhao 436

(2010) report where on the microatoll surface the samples were collected for their analyses 437

(i.e. collected from the outer rim, centre or elsewhere); this limits the interpretation of the 438

timing of sea-level position. One of the most important considerations for sea-level studies is 439

the precision to which elevation is measured and understanding the influence of other 440

phenomena that may limit upward coral growth such as wave climate, ENSO and tidal 441

patterns; such data on both spatial and temporal scales are rare. One such record exists from 442

Christmas Island, Pacific Ocean, where detailed survey data of microatolls showed the 443

Page 17: Post-glacial sea-level changes around the Australian

14

influence of the geoid effect where living colonies from the southern section of the island 444

were ~ 1 m lower than those on the northern side (Woodroffe et al., 2012). 445

446

While coral reef flats are typically restricted to low water springs tidal levels, shingle 447

ramparts or algal rims on the reef flat can form moats or pools which effectively retain water 448

above these tidal levels. These moats allow coral microatolls to grow to much higher 449

elevations (the height largely dependent on tidal range), in some cases up to mean sea level 450

(up to 2 m higher in macrotidal situations), but more commonly no higher than the low water 451

neap tides (approximately 1 m higher than low water springs; Scoffin and Stoddart, 1978; 452

Hopley, 1982, 1986a, 1987). Moating can be identified on reef flats from elevated pools at 453

low water levels where coral colonies may display different surface morphology than their 454

open water counterparts such as ‘smoother’ tops or terraces (Scoffin and Stoddart, 1978; 455

Hopley, 1986a). Moating of fossil microatolls can be more difficult to identify. A tropical 456

cyclone in the Great Barrier Reef in 1918 breached a shingle rampart on Holbourne Island 457

killing the moated microatolls on the reef flat (Hopley and Isdale, 1977); without this prior 458

knowledge these microatolls may have been interpreted as representing a higher sea level. 459

Chappell et al. (1983) suspected some microatolls dated in their study were moated based on 460

these microatolls being younger than those found offshore. These microatolls were excluded 461

from their sea-level curve. Importantly, whereas coral microatolls can overestimate sea levels 462

due to moating, it is not possible for them to underestimate the relative sea-level position; 463

therefore other palaeo-sea-level indicators that suggest lower sea levels than the microatolls 464

need to be treated with caution. 465

466

In microtidal settings outside the cyclone belt, moating of microatolls is less likely. Extensive 467

surveys of several hundred Porites microatolls on the Cocos (Keeling) Islands revealed that 468

their upper surfaces are constrained between mean low water springs and mean low water 469

neaps range (Smithers and Woodroffe, 2000; Smithers, 2011). Open water microatoll ‘fields’ 470

grow within a ±10 cm elevation range. They can be influenced by wave climate and tidal 471

range but the upper surface of modern microatolls is clearly constrained by a tidal plane, and 472

the former upper surface of fossil microatolls represents one of the more reliable indicators of 473

the tidal level at which they were confined (Woodroffe and McLean, 1990; Smithers and 474

Woodroffe, 2001). 475

476

3.3 Encrusting organisms 477

Page 18: Post-glacial sea-level changes around the Australian

15

Encrusting organisms, such as oysters, tubeworms and barnacles (also referred to as fixed 478

biological indicators; Chappell, 1987), are confined to a restricted range within the intertidal 479

zone on rocky shorelines and have been used as sea-level indicators (Baker and Haworth, 480

1997; Lewis et al., 2008). For example, encrusting intertidal organisms may be buried or 481

submerged as a consequence of rapid sea-level rise, but are more likely to be preserved under 482

a falling sea level. Although their potential as sea-level indicators has long been recognised 483

(Smith, 1978; Bird, 1988) and their indicative relationships to tidal levels at a particular site 484

have been relatively well-established (e.g. Endean et al., 1956; Dakin, 1987), ages of elevated 485

relict deposits were not obtained until 1988 and into the 1990s (Playford, 1988; Flood and 486

Frankel, 1989; Beaman et al., 1994; Baker and Haworth, 1997). Subsequently, researchers 487

have used relict oyster bed, barnacle and tubeworm deposits in sea-level reconstructions 488

along the east and west coasts of Australia to delineate the magnitude and duration of the 489

mid-Holocene highstand (Beaman et al., 1994; Baker and Haworth, 1997, 2000a; Baker et al., 490

2001a, b, 2005; Lewis et al., 2008). 491

492

In north-eastern Australia, the oysters Saccostrea cuccullata and S. echinata (also referred to 493

as Crassostrea amasa) form large beds/visors of up to 1 m width in their upper ~20 cm 494

vertical growth range (approximately mean high water neaps). This upper limit can be traced 495

horizontally to within ±0.15 m across an embayment with a 2 km headland, although in areas 496

of wave run-up deposits can occur at higher elevations in crevices or with changing wave 497

exposure. The nature of these fossil oyster bed deposits, which can accumulate over periods 498

of up to 1000 years, suggest that sea levels were stable or fluctuated around a common 499

position for lengthy periods in the mid-Holocene; this evidence is difficult to reconcile with 500

the smoothly falling sea levels inferred from sequences of coral microatolls and other 501

prograding accretionary deposits (Beaman et al., 1994; Lewis et al., 2008). Fossil barnacles 502

of Octomeris brunnea and Tetraclitella sp. from north-eastern Australia (Beaman et al., 1994; 503

Higley, 2000) clearly have potential as indicators of past water levels in this region, although 504

detailed studies are required before their full significance can be determined (Hopley et al., 505

2007). 506

507

Serpulid tubeworms (commonly Galeolaria caespitosa) are constrained between the low and 508

mid-tide range (up to high water neap); their use as sea-level indicators was described by 509

Bird (1988). The transitional boundary between tubeworm and the barnacles Chamaesipho 510

tasmanica and Tesseropora rosea forms a distinct marker indicative of a former tidal level 511

Page 19: Post-glacial sea-level changes around the Australian

16

reported to have a reproducible elevation range within ±10 cm; where this transitional 512

boundary is not observed, an accuracy of ±25 cm has been reported (Baker and Haworth, 513

1997, 2000a; Baker et al., 2001a, b). Elevation of this tubeworm-barnacle boundary can vary 514

under different exposure regimes (Baker and Haworth, 1997). G. caespitosa typically grow in 515

sheltered environments and two separate growth morphologies have been recognised which 516

have been linked to their vertical growth range (Baker and Haworth, 2000a). However, 517

tubeworm and barnacle deposits can occupy a much broader depth envelope in regions of 518

larger tidal range or higher wave exposure (e.g. Laborel and Laborel-Deguen, 1996), 519

diminishing their potential use for high-resolution reconstructions of past sea-level changes 520

(Sloss et al., 2007). Indeed, relict tubeworms from Wasp Head and Clear Point, New South 521

Wales, a region with relatively higher wave/swell exposure, were reported at +2.7 to 3.0 m 522

above the modern assemblage (Baker et al., 2001b). Baker et al. (2001b) rejected these 523

deposits as reliable indicators of former sea level, arguing that the high elevation was an 524

artefact of the energetic wave regime at the sample location. Recognition of wave exposure 525

as a possible confounding factor is important, as is acknowledgement that a key assumption 526

when using these indicators to examine past sea-level changes is that wave conditions have 527

not significantly changed over the period of investigation (see Goodwin, 2003). 528

529

Variations in the upper growth limit of these organisms at a single location is another issue 530

that constrains the use of encrusting organisms, implying that they are more likely to preserve 531

evidence of local variations in past conditions at that site rather than indicating a regional sea-532

level trajectory. This is particularly so as encrusting organisms appear to be best preserved in 533

sheltered settings where local effects can be important, and less well preserved on open water 534

settings where they can be directly compared with their modern counterparts (e.g. Scoffin, 535

1977; Hopley et al., 2007; Perry and Smithers, 2011; Smithers, 2011). However, suitable 536

growth patterns for oyster deposits can sometimes be distinguished in the ancient record 537

based on their morphology such as those that form a thick (> 20 cm) bed/visor on a horizontal 538

plane with clear separation of visor and wave splash zone (Lewis, 2005). As most relict 539

oyster bed deposits are preserved in caves and crevices, they may record a relatively higher 540

sea level compared with more open-water indicators such as microatolls (Lewis et al., 2008; 541

Smithers, 2011). Nevertheless, these deposits can accumulate over long periods and their 542

continued presence at similar elevations across large parts of eastern, northern and Western 543

Australia provides evidence for a prolonged highstand over the mid-late Holocene. Moreover, 544

changes in the ecology, morphology and elevation of encrusting organisms (e.g. Baker and 545

Page 20: Post-glacial sea-level changes around the Australian

17

Haworth, 2000a; Baker et al., 2001b; Wright, 2011) are likely to provide important evidence 546

for changes in coastal boundary conditions such as water level, wave exposure and climate. 547

Important data have been gathered on both the geographical and elevation ranges of oyster 548

and barnacle species surveyed to tidal datum along the Great Barrier Reef coastline and 549

offshore islands extending from Cooktown to Gladstone (Wright, 2011). 550

551

3.4 Stratigraphical sequences 552

Distinctive boundaries between different sedimentary facies can be used to decipher past sea 553

levels where the depositional development of different facies are related to sedimentary 554

processes dominant at particular elevations with respect to sea level. Such facies relationships 555

have been used to establish relative changes in sea level in sedimentary settings including 556

raised beach/barrier systems preserved in coastal plains (Burne, 1982; Belperio et al., 1984, 557

2002; Searle and Woods, 1986), the sedimentary infill of incised coastal valleys (e.g. Jones et 558

al., 1979; Gagan et al., 1994; Sloss et al., 2005, 2006, 2007, 2011), and in sediment cores 559

from the continental shelf that preserve sequences of marine (onlap) and terrestrial (offlap) 560

facies related to inundation and emergence (e.g. Larcombe and Carter, 1998). Different 561

materials have been dated within these deposits, including mangrove wood, intertidal and 562

subtidal molluscs, charcoal, peat, the remains of seagrass fibres and organic rich muds. In 563

some cases, changes in sediment facies within sediment cores can be directly related to 564

intertidal deposition and processes linked to a specific tidal datum or depth. For example, 565

Belperio et al. (1983, 1984, 2002) produced detailed sea-level reconstructions for different 566

sections of Spencer Gulf and the Gulf of St Vincent, South Australia (partitioned according to 567

distance from the continental margin) using established relationships between transitions 568

from subtidal seagrass facies to intertidal sand flat facies, and from sand flat to mangrove 569

facies (material included seagrass, mangrove and molluscs). The work of Bruce Thom, Peter 570

Roy and colleagues on the geomorphology and sedimentology of back barrier deposits and 571

drowned river valley successions along the coasts of southern Queensland and New South 572

Wales also allowed different depositional environments to be identified and related to sea-573

level position (Martin, 1972; Macphail, 1974; Roy et al., 1980; Roy and Thom, 1981; Thom 574

et al., 1981, 1992; Thom, 1984b). These data on the stratigraphic sequences in New South 575

Wales provided important information on the post-glacial sea-level rise and when present sea 576

level was achieved. Moreover, the studies provided insights on the various responses of these 577

coastal deposits that could be interpreted in relation to local levels of the sea, including tidal 578

planes. 579

Page 21: Post-glacial sea-level changes around the Australian

18

580

Searle and Woods (1986) used the interface between sublittoral sands and beach sediments 581

(‘swash zone’ facies) to reconstruct sea-level changes along the southern Western Australian 582

coastline. In their reconstruction, they used radiocarbon ages from well-preserved ‘delicate’ 583

bivalves (Donax and Paphies sp.); these bivalves are easily weathered and broken and so 584

well-preserved materials were considered to indicate little or no reworking (Searle and 585

Woods, 1986). A transgressive sandsheet formed as rising seas breached remnants of Last 586

Interglacial barriers during the most recent post-glacial marine transgression (ca. 12,000 – 587

7000 cal. yr BP) is widely encountered in coastal embayments in eastern Australia (e.g. Sloss 588

et al., 2006, 2007), where it extends up to near present sea level. The transgressive deposit is 589

characterised by a shell-rich mix of marine and estuarine molluscan fauna and represents 590

reworked intertidal sand flats, tidal channel sands and flood tide delta sands (Sloss et al., 591

2006, 2007, 2011). Recognition of this and overlying facies provides another example of the 592

application of facies associations and stratigraphy as a sea-level indicator. However, such 593

deposits accumulate over a relatively long time period (ca. 2,000 years), can only provide a 594

time-averaged assessment of sea level at the time of deposition, and may contain reworked 595

molluscs which complicate interpretations of the age of the deposits (Sloss et al., 2005, 596

2007). As for the biological indicators, a key assumption in the interpretation is that coastal 597

boundary conditions such as wave exposure, tidal range and sediment supply have remained 598

relatively constant as the deposit has formed (see Chappell and Thom, 1986). 599

600

3.5 Mangrove deposits 601

Many of the numerous estuaries along the tropical and sub-tropical coasts of Australia 602

contain extensive mangrove forests. Transgressive mangrove muds also occur beneath the 603

deltaic plains or marginal estuarine deposits which record post-glacial inundation of former 604

valleys by rising seas. Such mangrove deposits are relatively good indicators of this rapid 605

sea-level rise (e.g. high preservation potential) and they are generally restricted to the upper 606

part of the intertidal zone. Several researchers have used organic clays, peats, roots, wood 607

fragments and in situ stumps that are associated with mangroves to indicate Holocene sea-608

level change where mangroves occur (e.g. Jennings, 1975; Belperio, 1979; Jones et al., 1979; 609

Thom and Roy, 1983, 1985; Grindrod and Rhodes, 1984; Woodroffe et al., 1987; Woodroffe, 610

1988; Sloss et al., 2007). The accuracy with which mangrove deposits can be used to define 611

past water level is likely to be less in areas of large tidal range, as mangroves grow from 612

mean sea level up to highest astronomical tide (HAT), with some isolated stands of mangrove 613

Page 22: Post-glacial sea-level changes around the Australian

19

persisting even above this level in a few locations. Mangrove wood and the fibrous remains 614

of the roots of Rhizophora are distinctive but these organic remains are highly compressible, 615

and the muds within which they occur can be compacted (Woodroffe, 1988; Grindrod et al., 616

1999). It is therefore important to choose basal samples that overlie bedrock or consolidated 617

pre-Holocene deposits if the evidence is to be used in reconstructing the final stages of post-618

glacial sea-level rise (Woodroffe et al., 1987, 1989). 619

620

Where intertidal mangrove stumps can be detected, they are particularly effective sea-level 621

indicators. Jones et al. (1979) describe the stumps of Avicennia exposed by storm erosion on 622

a beachface at Bulli in the Illawarra, and an age on these remains (7720 ± 260 cal. yr BP) 623

continues to support the argument for early attainment of present sea level in New South 624

Wales (Young et al., 1993; Sloss et al., 2007). However, it is noted that this is only one datum 625

and accordingly should be viewed with caution until the early age of the onset of the 626

Holocene highstand can be confidently replicated. Roy and Crawford (1981) obtained a 627

comparatively younger radiocarbon age (7030 ± 280 cal. yr BP) from a fossil mangrove 628

stump (Avicennia marina) from Kurnell Peninsula, New South Wales at PMSL. Mangrove 629

stumps, together with carbonate nodules containing the remains of mud lobsters (Thalassina) 630

crop out on erosional banks along many of the estuaries of northern and north-western 631

Australia, and are associated with ‘big swamp’ mangrove deposits that record a phase of 632

widespread mangrove forest development between 7900 and 5800 cal. yr BP as sea level 633

stabilised (Woodroffe et al., 1985). 634

635

Mangrove deposits are good indicators of rapid sea-level rise or when sea level attained a 636

particular elevation, but are less useful indicators of sea-level stillstands or slow transgression 637

because: they can grow in the upper part of the intertidal zone from around mean sea level to 638

the supratidal zone, and their roots penetrate into the muddy sediments for up to as much as 1 639

m (Bunt et al., 1985; Woodroffe, 1988; Smithers, 2011). Mangrove sediments are also prone 640

to post-depositional compaction (Chappell, 1987; Beaman et al., 1994; Larcombe et al., 641

1995); and the intrusion of younger rootlets presents contamination issues for radiocarbon 642

ages (Beaman et al., 1994; Hopley et al., 2007; Smithers, 2011). Despite these potential 643

problems, mangrove deposits have been used to indicate that mid-Holocene sea levels have 644

not been significantly higher than 1 m above PMSL in several locations including Western 645

Australia, Northern Territory and Queensland (e.g. Jennings, 1975; Belperio, 1979; 646

Woodroffe et al., 1987). These deposits appear to underestimate sea-level elevations that 647

Page 23: Post-glacial sea-level changes around the Australian

20

have been established with other sea-level indicators (e.g. microatolls and encrusting 648

organisms) that suggest sea levels were >1 m above PMSL, a result likely due to compaction 649

and erosion of mangrove deposits and/or modifications of the sea-level signal as the coastal 650

setting has evolved over the late Holocene (Chappell et al., 1983; Beaman et al., 1994; Baker 651

et al., 2001b). Chappell and Grindrod (1984) discuss this issue for Princess Charlotte Bay, 652

where the chenier plain and underlying mangrove sediments indicate a gradual progradation 653

of the coast with no detectable change in sea level, at odds with evidence of emergence on 654

nearby offshore islands. In this, as in other locations in northern Australia, it is likely that 655

morphodynamic changes in the landscape result in local alteration of tidal prism and 656

consequently evidence records water levels locally (Chappell and Thom, 1986). 657

658

3.6 Supratidal deposits 659

Supratidal deposits include sedimentary facies or organic remains that have been deposited 660

above the highest astronomical tide (HAT). Supratidal chenier ridges deposited over 661

mangrove muds provide an example, where the boundary between the two deposits has been 662

used to indicate a sea-level position in several studies. This relationship was applied on a 663

chenier sequence near Karumba, Gulf of Carpentaria, to establish a mid-Holocene highstand 664

of ~ 2.5 m above PMSL followed by a gradual sea-level fall (Rhodes et al., 1980; Rhodes, 665

1982; Chappell et al., 1982). At Broad Sound on the central Queensland coast, chenier 666

deposits have similarly been examined to infer sea-level changes across the eastern and 667

western sides of the sound (Cook and Polach, 1973; Cook and Mayo, 1977). 668

669

Chenier ridges are wave-built landforms comprising generally relatively coarse sediments 670

and commonly abundant accumulations of marine shells. Despite numerous investigations, 671

there remains insufficient evidence to demonstrate that chenier bases define a reproducible 672

datum level, or that they occur above the high tide plane. In Princess Charlotte Bay, it was 673

argued that the sedimentary deposits were compacted and thus particular ridges 674

underestimated sea level (Chappell and Grindrod, 1984; Chappell, 1987). Moreover, chenier 675

ridges are ‘complex response systems’ strongly influenced by changes in storm frequency, 676

sediment supply and shell production rates, which may confound the production and 677

interpretation of a reliable sea-level signal (Chappell and Thom, 1986). 678

679

Other supratidal indicators include terrestrial materials such as freshwater peats, organic-rich 680

muds, wood fragments and in situ tree stumps. They provide a ‘directional’ sea-level 681

Page 24: Post-glacial sea-level changes around the Australian

21

indicator, implying that sea level was lower than the elevation at which they occur at a 682

particular point in time, but they do not indicate how far below. In some cases, these 683

indicators can be used in conjunction with intertidal/estuarine sediments that bury or overlie 684

them. For example, in situ tree stumps that are covered by estuarine muds have been used to 685

define when sea level first reached a particular elevation (Sloss et al., 2007). Supratidal 686

deposits are prone to compaction, reworking and diagenesis, and thus are generally not relied 687

on to reconstruct accurate accounts of former sea-level position (Larcombe et al., 1995). 688

689

3.7 Beachrock 690

Beachrock comprises cemented beach sands and is usually found as clearly bedded and 691

dipping outcrops of consolidated or moderately consolidated beach sediment, reflecting the 692

gradient of the beach. The details of beachrock formation have been debated in the literature 693

for several decades, but it is now generally conceded that it can form through several 694

different processes (McLean, 2011, and references therein). It is largely agreed that 695

beachrock forms in the intertidal zone when unconsolidated sediments become lithified by 696

aragonitic and/or calcitic cements, preserving the internal fabric of the beach. Outcrops can 697

be over 3 m in thickness in areas with relatively high tidal ranges such as on the Great Barrier 698

Reef or exposed to large waves (Hopley, 1986b; McLean, 2011). While beachrock may 699

provide a relatively constrained sea-level indicator in microtidal environments, the exact 700

uppermost limit of formation is difficult to determine, particularly where the tide range is 701

large. Therefore there is both considerable variability and uncertainty as to what tidal level, if 702

any, the upper limit of beachrock indicates (Hopley, 1986b, 1987; Hopley et al., 2007). Based 703

on beachrock occurrence above assumed sea-level limits, Kelletat (2006) asserted that 704

beachrock forms in supra-tidal environments well above the highest water levels due to 705

splashing and spraying, however Knight (2007) dismissed this conclusion arguing that 706

beachrock forms in the beachface proper by well-established processes (described in 707

McLean, 2011). 708

709

Hopley (1971, 1975, 1980) interpreted raised beachrock preserved on continental islands in 710

north-eastern Australia, as evidence that sea level had been higher, reporting deposits 3 to 4 711

m above the current mean high water springs tide level (thought to represent the upper limit 712

of cementation). This original estimate of elevation was later revised down to 3 m above the 713

present HAT level (Hopley, 1983b, 1986b). Assuming that this elevation documents the peak 714

highstand sea level, it must be noted that it is still 1.0 to 1.5 m higher than microatoll (highest 715

Page 25: Post-glacial sea-level changes around the Australian

22

‘unmoated’ microatoll is +1.45 m) and oyster bed (+1.65 m: Beaman et al., 1994) data from 716

the Great Barrier Reef. 717

718

The use of beachrock as a sea-level indicator is also problematic because it is difficult to 719

determine a precise age on its formation. The constituent grains that formed the beach sand 720

before lithification are almost certain to be of widely differing ages and pre-dating the 721

cementation event. Accordingly, in principle it would be preferable to directly date the 722

cements instead. Well-cemented beachrock may have undergone several diagenetic phases 723

(Vousdoukas et al., 2007), and consequently the cements from the one deposit can generate 724

different ages to any biogenic carbonate (Desruelles et al., 2009). While the ages of 725

beachrock cements provide a minimum timing of deposition, the analysis of shell and coral 726

material in beachrock provides a maximum age (Hopley, 1983b, 1986b), and this could span 727

a wide range, for example shells on modern beaches can be up to 8000 years old (Donner and 728

Junger, 1981) or possibly even reworked from previous interglacial successions (Murray-729

Wallace and Belperio, 1994). Due to the issues related to uncertainties relative to sea level at 730

the time of formation, mixed age of biogenic carbonates and post-depositional diagenetic 731

processes, beachrock should only be used as a sea-level indicator with caution and is only 732

acceptable if the elevations agree with other recognised indicators (Smithers, 2011). 733

734

3.8 Foraminifera 735

Distinct changes in microfossils, particularly foraminiferal assemblages, have long been used 736

for sea-level reconstructions on salt marshes in temperate regions around the world, most 737

recently using complex statistical ‘transfer functions’ (Gehrels, 2000; Gehrels et al., 2001; 738

Woodroffe, 2009, and references therein). The use of foraminifera in this way requires 739

detailed studies on the distribution and zonation of foraminifera in modern environments 740

(Haslett, 2007), on the basis of which it may then be possible to interpret changes in 741

assemblages in sediment cores in terms of past sea-level adjustments (e.g. Haslett et al., 742

2010). Although microfossil studies have been conducted in southern Australia to reconstruct 743

the characteristics of the post-glacial marine transgression (Belperio et al., 1988; Cann et al., 744

1988, 1993, 2002, 2006; Wang and Chappell, 2001), it is only recently that the ‘transfer 745

function’ approach has been adopted in the Australasian region to decipher sea-level changes 746

over the past few millennia (Southall et al., 2006; Woodroffe, 2009; Callard et al., 2011; 747

Gehrels et al., 2012). 748

749

Page 26: Post-glacial sea-level changes around the Australian

23

In the Australian context, these techniques have been extended from temperate saltmarsh 750

systems to tropical mangrove environments, particularly those associated with the Great 751

Barrier Reef (Horton et al., 2003, 2007; Woodroffe et al., 2005). Adopting this approach, 752

Woodroffe (2009) produced a sea-level reconstruction for Cleveland Bay in north 753

Queensland based on benthic foraminifera. The results indicate a prolonged mid-late 754

Holocene highstand with sea level reaching elevations as high as 2.8 m above PMSL. These 755

elevations conflict with evidence for a less pronounced highstand using other sea-level 756

indicators for the region, such as coral microatolls and oyster beds. The taphonomy and 757

elevational accuracy with which these microfossils can be used as sea-level indicators has 758

been questioned as they can clearly be transported and redeposited (Smithers, 2011; Perry 759

and Smithers, 2011). Recent studies suggest that the foraminiferal transfer function may be 760

inappropriate in tropical environments due to complex mixing, bioturbation and degradation 761

processes that affect preservation potential and modify assemblage composition (Berkeley et 762

al., 2009a, b). A prograded coastal plain can provide a sedimentary record indicating little 763

overall change in sea level over the mid-late Holocene (e.g. Thom and Roy, 1985), but 764

intertidal microfossils (foraminifera, mangrove pollen) may preserve a more confused history 765

because of compaction, contamination or diagenetic changes (oxidation of near-surface 766

sediments and the development of potential acid-sulphate soils). The full potential of these 767

methods needs to be further examined in a broader range of environments around Australia, 768

but the approach appears to be most suitable for low-energy microtidal environments outside 769

of the tropics (Callard et al., 2011; Gehrels et al., 2012). 770

771

3.9 Intertidal erosional indicators 772

In many parts of the world, past sea-level highstands are best indicated by erosional evidence. 773

Intertidal erosional indicators are typically carved into softer sedimentary rocks (Pirazzoli, 774

1996; Smithers, 2011). Indeed, there has been a long history in Australia using erosional rock 775

platforms as evidence for higher sea level around the Australian mainland (reviewed in 776

Langford-Smith and Thom, 1969). Raised and submerged notches in Western Australia were 777

used in Fairbridge’s (1961) sea-level compilation and subsequent researchers have analysed 778

relict tubeworm deposits preserved within notches on Rottnest Island (Playford, 1988; Baker 779

et al., 2005), presuming that these deposits provide a minimum age for notch formation. 780

Although notches can provide visually persuasive evidence of former sea-level positions, 781

they can rarely be dated accurately. 782

783

Page 27: Post-glacial sea-level changes around the Australian

24

3.10 Summary 784

The broad coastal plains, sand barrier systems and estuaries of southern Australia, and the 785

coral reefs and extensive deltaic-estuarine plains of northern Australia have developed during 786

the mid and late Holocene when the sea has been relatively close to its present level. These 787

prograding stratigraphical sequences provide a first-order indication of the accumulation of 788

shoreline deposits, and they contrast with the coastal landforms that have developed in parts 789

of the world that have not experienced such relative sea-level stability, either because of 790

vertical tectonic movements or various isostatic influences on relative sea-level changes. 791

792

Particular sea-level indicators are more likely to be formed, and preserved, under different 793

sea-level scenarios; some are dependent on suitable accommodation space. Despite the 794

limitations of geographical distribution and the patchy spread of evidence, as well as the 795

uncertainties in indicative meanings that may be associated with proxy sea-level indicators, 796

they all clearly have a role in indicating aspects of past sea levels. The accuracy and precision 797

of each, and the reconciliation of the apparent discrepancies between them, are the key 798

challenges for those who seek a more regional insight into past sea-level change, as well as 799

those who are attempting to develop histories of local coastal landform development. The 800

following section provides examples of how different sea-level indicators have been used at 801

specific locations around Australia. It demonstrates how no one proxy sea-level indicator or 802

specific site contains enough data over a sufficient interval of the post-glacial to provide 803

evidence for a continuous sea-level history. Accordingly, in an attempt to provide a 804

comprehensive review of sea-level research around the Australian margin, we have split the 805

sections into two specific intervals of time (20,000 to 7000 cal. yr BP and 8000 cal. yr BP to 806

present) with the latter period further divided to reflect geographical distribution. 807

808

4 Post-glacial sea-level rise (Last Glacial Maximum to 7000 cal. yr BP) 809

The first estimate for the LGM sea-level lowstand in the Australian region was produced by 810

van Andel and Veevers (1967) who reported a -130 m sea level at ~18,000 years BP based on 811

shell material from the Sahul Shelf. Subsequently, a sea level of -175 to -150 m at ~17,000 812

years BP was reported for the southern Great Barrier Reef (offshore One Tree Island) based 813

on a submerged dredged Galaxea clavus coral colony (the modern range of this species can 814

occur up to 75 m below sea level) and a lithified sample interpreted as ‘beachrock’ (Veeh and 815

Veevers, 1970). Phipps (1970) found evidence in New South Wales for sea levels at -130 m 816

around 18,000 years BP. Lowstand sediments, containing alternating successions of fine-817

Page 28: Post-glacial sea-level changes around the Australian

25

grained mixed quartz-carbonate sand and densely packed mollusc-dominated sediments with 818

the shallow-water molluscs Pecten fumatus and Placamen placidium commonly represented, 819

were identified in vibrocores collected from the outer continental shelf off New South Wales 820

in depths between 123 and 152 m below PMSL, in broad agreement with depth suggested by 821

the earlier evidence. AAR dating of the shallow water shells indicates that these successions 822

represent deposition during the past three glacial maxima (Ferland et al., 1995; Murray-823

Wallace et al., 1996, 2005), confirming that sea levels during these maxima were at least 120 824

m below PMSL (Murray-Wallace et al., 2005). 825

826

A more detailed record from cores collected in the Joseph Bonaparte Gulf by Yokoyama et 827

al. (2000, 2001a) places sea level at -124 m (± 4 m) at ca. 20,000 years BP. While some 828

disagreement remains on the precise timing of the LGM (Peltier, 2002; Peltier and Fairbanks, 829

2006; De Deckker and Yokoyama, 2009), the magnitude of the sea-level lowstand at -120 to -830

130 m in the Australasian region has generally been agreed upon (e.g. Hopley and Thom, 831

1983; Chappell, 1987; Lambeck and Chappell, 2001; Murray-Wallace et al., 2005) and is 832

consistent with glacio-eustatic ice sheet models (e.g. Lambeck and Nakada, 1990; Peltier and 833

Fairbanks, 2006) and evidence collected from the Sunda Shelf (Hanebuth et al., 2000, 2009). 834

Forthcoming results from IODP expedition 325 to the Great Barrier Reef are likely to further 835

resolve the timing and depth of sea level at the LGM, and the nature of subsequent post-836

glacial sea-level rise (Webster et al., 2011; Yokoyama et al., 2011). 837

838

The key records of post-glacial sea-level rise from the Australian sites (Fig. 2) as well as 839

from Barbados and Tahiti show remarkable agreement when corrected for tectonic influences 840

(e.g. Lambeck and Chappell, 2001; Woodroffe, 2003). Although the broad pattern of eustatic 841

post-glacial sea-level rise has been independently supported by these other studies, there 842

continue to be further refinements to the detail, particularly associated with the relatively fast 843

rise during meltwater pulse 1A (and possibly 1B and 2), and a slowed rate of rise during the 844

Younger Dryas. Ooid deposits formed in shallow marine environments in the southern Great 845

Barrier Reef place sea level at approximately 100 m below PMSL at 16,800 years BP just 846

prior to meltwater pulse 1A (Yokoyama et al., 2006). Meltwater pulse 1A is particularly well-847

represented in the Sunda Shelf record (detected largely using mangrove facies). Sea level rose 848

about 16 m in only 300 years between 14,300 to 14,600 years BP (Hanebuth et al., 2000, 849

2009). Following this time, the coral core data from the uplifted reef platforms on Huon 850

Peninsula, Papua New Guinea, provide a relatively complete record from 13,000 to 7000 851

Page 29: Post-glacial sea-level changes around the Australian

26

years BP showing a continuous rise over this period (Chappell and Polach, 1991; Edwards et 852

al., 1993; Ota and Chappell, 1999). This period is also partially covered by other datasets 853

from locations including Christchurch, New Zealand (Gibb, 1986), Great Barrier Reef 854

(Larcombe et al., 1995; Larcombe and Carter, 1998), New South Wales (Sloss et al., 2007), 855

southern Australia (Belperio et al., 2002; Cann et al., 2006) and Western Australia (e.g. the 856

Abrolhos Islands, Eisenhauser et al., 1993) (Fig. 2). 857

858

In 1982 the Sirius Expedition, led by Nic Flemming, surveyed near Cootmaundra Shoal, 859

northwest of Bathurst Island and 240 km from Darwin, looking for archaeological evidence 860

of early human colonisation of this part of north-western Australia. Terraces at 90, 59, 48, 34 861

and 25 m depth were identified on the van Diemen Rise (van Andel and Veevers, 1967; 862

Chappell and Thom, 1977), from which reefal carbonate shoals arise, with Cootamundra 863

Shoal itself the shallowest at 16 m depth. The shoals presently support scattered live 864

Turbinaria, Platygyra and Porites corals, but beneath a cemented algal crust, sticks of 865

branching Acropora were recovered. Radiocarbon ages of 8460 ± 100 years BP (ANU3259) 866

from 28 m and 7630 ± 100 years BP (ANU3257) from 20 m are broadly consistent with the 867

anticipated sea-level curve (Flemming, 1986), although a further 14 unpublished ages were 868

also obtained. Coral rubble was recovered to depths of at least 60 m, and inferred to 80 m, 869

underlain by incised siliceous sandstone. A prominent terrace level was noted at 24-28 m 870

below PMSL. 871

872

Detailed palaeoecological and sedimentary investigations reveal that rising post-glacial sea 873

levels breached the Arafura Sill ca. 12,200 cal. yr BP with full marine conditions in the Gulf 874

of Carpentaria established by 10,500 cal. yr BP (Chivas et al., 2001; Yokoyama et al., 2001b; 875

Reeves et al., 2008). U-series ages of the drowned coral reefs at depths of 20-30 m in the 876

southern Gulf of Carpentaria indicate that reef growth commenced between 10,500 and 9500 877

cal. yr BP associated with marine inundation into the Gulf and continued to flourish until ca. 878

7000 yr BP (Harris et al., 2008). Similar extensive fossil reefs in 30-50 m water depth have 879

been described from around Lord Howe Island, at present the southernmost limit to reef 880

growth in the Pacific Ocean, with radiocarbon and U-series ages from the same period 881

between 9100 and 7300 cal. yr BP (Woodroffe et al., 2010). 882

883

Some researchers have argued for a highly episodic post-glacial sea-level rise (up to nine 884

pauses) using data from the Great Barrier Reef and New Zealand (Carter and Johnson, 1986; 885

Page 30: Post-glacial sea-level changes around the Australian

27

Carter et al., 1986; Larcombe et al., 1995; Larcombe and Carter, 1998). These data (including 886

undated abrasion platforms/shorelines/drowned reefs and dated sea-level indicators) have 887

been interpreted to represent several stillstands or even a 6 m regression (centred around 8200 888

years BP) during this period. A large oscillation around this time was shown in the sea-level 889

curve proposed by Larcombe et al. (1995); this oscillation/regression was originally 890

suggested as a possibility by Thom and Chappell (1975). Harris (1999; see also Harris and 891

Davies, 1989) rejected this claim highlighting that there is no evidence for a 6 m sea-level fall 892

in ice sheet models or in the marine δ18O record, and Hopley et al. (2007) offered another 893

critique arguing that the elevation and error terms associated with the data were too large to 894

achieve the precision necessary to identify the proposed oscillation. There is an increasing 895

body of evidence from around the world that there may have been a rapid rise at around 8200 896

years BP. An inflection occurs in a sea-level reconstruction from Singapore (Bird et al., 897

2007) which appears to coincide with the back-stepping identified in several southeast Asian 898

deltas at that time (Hori and Saito, 2007). Such a rapid sea-level rise might explain the 899

apparent drowning of submerged reefs around Lord Howe Island, and even those in the Gulf 900

of Carpentaria (Harris et al., 2007, 2008; Woodroffe et al., 2010). However, although this 901

phase of reef drowning would be consistent with such an inflection in the sea-level curve 902

(meltwater pulse 2), there is no direct evidence for a sea-level stillstand or regression during 903

this period in the latest available data (Fig. 2). 904

905

5 Mid to late Holocene sea-level changes (~8000 cal. yr BP to present) 906

Since the earlier sea-level compilations by Hopley (1983a, 1987) and Chappell (1987), new 907

research has provided additional insights into Holocene sea-level variability around Australia. 908

While there seems to be a general consensus about when sea level attained its modern level 909

and the magnitude of the mid-Holocene highstand (albeit with some remaining 910

discrepancies), there appears to be regional variability and conflicting evidence regarding the 911

nature of late Holocene sea-level fall. In this section we divide the Australian mainland into 912

several regions and review the key sea-level evidence and interpretations from the mid-late 913

Holocene for each region. 914

915

5.1. Northern Australia (Gulf of Carpentaria/Northern Territory) 916

The principal mid to late Holocene records in northern Australia include studies focused on 917

the South Alligator River using mangrove sediments (Woodroffe et al., 1985, 1987, 1989) 918

and at Karumba using chenier deposits (Chappell et al., 1982; Rhodes et al., 1980; Rhodes, 919

Page 31: Post-glacial sea-level changes around the Australian

28

1982) (Fig. 1). Mangrove evidence from the South Alligator River indicates that the 920

transgression culminated in widespread mangrove forests, termed the ‘big swamp’ when sea 921

level reached its present level around 7400 ± 200 cal. yr BP. The mangroves within this 922

macrotidal setting extend across a vertical elevation of 3 m in the upper intertidal zone 923

(Woodroffe et al., 1987; Fig. 3), and the big swamp facies contains no evidence that the sea 924

was higher than present during the Holocene. Oxidation of the upper 1 – 2 m of the estuarine 925

plains may have destroyed mangrove material, although there does appear to be a transition 926

from mangrove pollen to pollen of grasses and sedges in the upper sections of some cores 927

(Woodroffe et al., 1985). 928

929

In contrast, the chenier data from Karumba, in the southern Gulf of Carpentaria, indicate that 930

sea level was much higher (approx. + 2.5 m) ca. 6400 cal. yr BP before falling smoothly to 931

present level after 1000 cal. years BP (original 14C ages in Rhodes, 1980; Chappell et al., 932

1982; Rhodes et al., 1982). This record of considerably higher sea level in the mid-Holocene 933

provided one of the key datasets for modelling of hydro-isostatic adjustment in northern 934

Queensland to explain the differences in shoreline elevation around the mainland (Chappell et 935

al., 1982, 1983). We note that the reliability of these storm deposited sediments as precise 936

palaeo-sea-level indicators is questionable (see section 3.6). Emergence along the southern 937

and eastern shores of the Gulf of Carpentaria is further supported by studies of the deltas of 938

the McArthur and Gilbert Rivers (Woodroffe and Chappell, 1993; Jones et al. 2003), and 939

younger radiocarbon ages (ca.1500 to 3300 cal. yr BP) on beachrock indicating a sea level of 940

+1 to 2 m near the McArthur River (Nott, 1996). 941

942

5.2 Queensland (east coast) 943

The Great Barrier Reef extends from Torres Strait, where only a few preliminary ages on 944

microatolls have been published (Woodroffe et al., 2000), south to Lady Elliot Island 945

(distance of 2,300 km), which comprises a sequence of ridges showing little variation in the 946

pattern of accumulation over the past 5000 years (Chivas et al., 1986). This is a particularly 947

wide area with large variability in the width of the continental shelf, and there are likely to be 948

substantial variations in relative sea-level change across this region which cannot be covered 949

in a single compilation (i.e. the importance of regionally/locally specific curves: Woodroffe, 950

2009; Lambeck et al., 2010). Geophysical modelling indicates a variable pattern of mid-951

Holocene sea-level change across the region (Chappell et al., 1982; Nakada and Lambeck, 952

1989). The time at which different reefs reached sea level varies from the outer reefs to the 953

Page 32: Post-glacial sea-level changes around the Australian

29

mainland coast (Hopley et al., 2007). In an early interpretation, Hopley (1984) invoked a 954

high-energy window, whereby landforms on the mainland coast contained evidence deposited 955

during high water levels associated with storm waves whose development was not 956

constrained by the outer reefs at that stage. In the late Holocene, when the outer reefs had 957

grown to sea level, that ‘window’ was closed, introducing a factor other than sea level to 958

which indicators have responded. There are relatively few useful data available from southern 959

Queensland, where fossil shell and coral suggest emergence of at least approximately 0.5 m 960

at 5670 ± 190 cal. yr BP (original 14C age in Flood, 1983). It is clear that there is a need for 961

further research to continue to refine the sea-level history of the Great Barrier Reef and 962

southern Queensland. 963

964

More recent studies have synthesised Holocene sea-level data from along the Queensland 965

coastline. Larcombe et al. (1995) compiled all available sea-level data (excluding beachrock) 966

from the central Great Barrier Reef (Cairns to Whitsunday Islands) over the past ca. 13,500 967

cal. yr BP. In summary, a synthesis of these data indicate that sea level approached its present 968

position (within 3 m) by 8170 ± 200 cal. yr BP (14C data from Hinchinbrook Island originally 969

reported in Grindrod and Rhodes, 1984). This is consistent with Woodroffe’s (2009) 970

conclusion that sea-level reached modern levels between 8000 and 6200 cal. yr BP. Lewis et 971

al. (2008) compiled the sea level data which they considered had the most reliably established 972

and precise indicative meaning (e.g. intertidal sea-level indicators such as coral microatoll 973

and encrusting organisms) from eastern Australian (Torres Strait to New South Wales, with a 974

concentration of data from Queensland) over the past 7000 cal. yr BP. Based on a relict 975

barnacle deposit from Magnetic Island sea level continued to rise to approximately +1 m by 976

7380 ± 240 cal. yr BP (Higley, 2000; Lewis et al., 2008) which is supported by ages from 977

fossil coral microatolls (+0.7 to 0.9 m) from Magnetic Island (7012 ± 22 cal. yr BP: Yu and 978

Zhao, 2010) and Orpheus Island (7530 ± 360 cal. yr BP: original 14C data in Zwartz, 1995). 979

980

The Queensland sea-level indicator evidence supports conflicting estimates for the elevation 981

of the mid-Holocene highstand. Sea-level estimates based on coral microatoll evidence places 982

relative sea level at +1.3 to 1.5 m between 6770 and 5750 cal. yr BP (original ages provided 983

in Chappell et al., 1983; Yu and Zhao, 2010), in good accord with that derived from oyster 984

bed data at +1.6 m between 6280 and 5720 cal. yr BP (Beaman et al., 1994; Higley, 2000; 985

Lewis et al., 2008). However, beachrock data suggest water levels of up to 2 to 3 m above 986

present sometime between 6450 and 3050 cal. yr BP (Hopley, 1980, 1983b, 1986b) and the 987

Page 33: Post-glacial sea-level changes around the Australian

30

highstand suggested by foraminiferal transfer function analyses is +2.8 to 3.3 m between 988

4270 and 3580 cal. yr BP (Woodroffe, 2009) (Fig. 4). While a higher mid-Holocene sea level 989

(> +2 m) cannot be completely ruled out, a +1.0 to 1.5 m highstand is currently the most 990

accepted elevation for this region (Hopley et al., 2007; Lewis et al., 2008; Perry and 991

Smithers, 2011). The most recent assessment indicates that microatolls and encrusting 992

organisms are more accurate sea-level indicators than beachrock or foraminifera (Perry and 993

Smithers, 2011). 994

995

There are conflicting interpretations of the mid to late Holocene sea-level fall in Queensland. 996

The earliest interpretation using microatoll data suggested that sea-level fell smoothly to 997

present levels (Chappell et al., 1983; Chappell, 1983). More recent data compilations have 998

indicated a sustained highstand followed by a pronounced fall after 2000 cal. yr BP (Lewis et 999

al., 2008; Woodroffe, 2009). Lewis et al. (2008) noted the possibility of up to two oscillations 1000

of about 1 m amplitude at approximately 4800 and 3000 cal. yr BP. Such marked oscillations 1001

have been discounted as an artefact of data interpolation (i.e. data gaps used to suggest 1002

oscillations and the validity of specific data points; Perry and Smithers, 2011). Perry and 1003

Smithers (2011) argue that the coral microatoll evidence should be given the most weight in 1004

sea-level reconstructions and contend that the smoothly-falling sea-level model remains the 1005

most plausible explanation for mid-late Holocene variability on this section of the 1006

Queensland coast. In contrast, the relict oyster bed deposits, which can accumulate for up to 1007

1000 years in relatively sheltered locations, indicate that sea levels remained stable for long 1008

periods and provide evidence for a stepped sea-level fall (Lewis, unpublished data). The 1009

disparity between sites is only partly attributable to the regional pattern of sea-level change, 1010

as it is clear from geophysical modelling that there has been a variable hydro-isostatic 1011

response over the region. Further anomalies can be ascribed to the fact that the indicators 1012

record different tidal or supratidal levels, and the fact that there have been local changes in 1013

these levels as the coast evolves. 1014

1015

5.3 South-eastern Australia 1016

The most recent compilation of Holocene sea-level data from south-eastern Australia presents 1017

many newly derived radiocarbon and AAR ages and a synthesis of previously published data 1018

from studies undertaken in the 1970’s and 1980’s (Sloss et al., 2004, 2007). A wider range of 1019

proxy sea-level indicators was considered, including molluscs from transgressive and 1020

estuarine sedimentary successions and in situ tree and mangrove stumps. The latter, in 1021

Page 34: Post-glacial sea-level changes around the Australian

31

particular, imply that modern sea level was attained between 7900 and 7700 cal. yr BP (Jones 1022

et al., 1979; Sloss et al., 2007). Although this age is more narrowly constrained than the range 1023

estimated from the Queensland data (between ca. 8200 – 7500 cal. yr BP), and is broadly 1024

consistent (if not earlier) with the inferred ages for other parts of Australia, its validity rests 1025

on a single radiocarbon measurement on mangrove wood previously published by Jones et al. 1026

(1979) and the time-averaged age of the transgressive facies that extend to present sea level 1027

(Sloss et al., 2007). 1028

1029

The compilation by Sloss et al. (2007) indicates that sea-level reached a highstand of +1.0 to 1030

1.5 m between 7700 and 7400 cal. yr BP, and remained at this elevation until 2000 cal. yr BP 1031

when sea level gradually fell to its present position (Fig. 5). The +1.0 to 1.5 m magnitude is 1032

marked by encrusting organisms/fixed biological indicators (tubeworms), molluscs and 1033

organic-rich muds and peats; this result is comparable with earlier reconstructions that 1034

suggest sea levels were no higher (or within ± 1 m) than present (e.g. Thom et al., 1969; 1035

Thom and Chappell, 1975; Thom and Roy, 1983, 1985). The higher mid-Holocene sea level 1036

suggested by Sloss et al. (2007) is also consistent with evidence from raised coral/boulder 1037

deposits and emerged carbonate mud deposits from Lord Howe Island that indicate mid-1038

Holocene sea levels as high as +1.0 to 1.5 m (Woodroffe et al., 1995). 1039

1040

The prolonged mid-Holocene sea-level highstand followed by a fall to present levels (with 1041

the possibility of oscillations) proposed by Baker et al. (2001a, b) conflicts with the smooth-1042

fall model that remains the favoured interpretation of Queensland data. The data comprising 1043

many ages on tubeworms reported by Baker and Haworth (2000a; Baker et al. 2001a, b) are 1044

central to their arguments. Baker and Haworth (2000b) suggest that there have been a series 1045

of sea-level oscillations based on statistical analysis of their data. They interpret up to three 1046

sea-level oscillations within the order of 1 m based on biostratigraphical relationships (i.e. 1047

assemblage changes etc). Sloss et al. (2007) and Perry and Smithers (2011) questioned the 1048

quality of the tubeworm and other encrusting organisms as accurate sea-level indicators, and 1049

suggested the sea-level oscillations more probably represent the adjustments of intertidal 1050

species to variations in coastal exposure and/or variable wave and climate conditions during 1051

the Holocene, or may simply be an artefact of missing data. This led Sloss et al. (2007) to 1052

conclude that the culmination of the Holocene marine transgression was followed by a sea-1053

level highstand that lasted until about 2000 years ago, followed by a relatively slow and 1054

smooth regression of sea-level from +1.5 m to present level. Donner and Junger (1981) found 1055

Page 35: Post-glacial sea-level changes around the Australian

32

evidence for a stable sea level with ‘no detectable fluctuations’ for the past 3200 years using 1056

radiocarbon ages of mollusc beds within low energy back barrier deposits from the 1057

Cullendulla Creek area (Bateman’s Bay), New South Wales. Seams of heavy minerals 1058

(ilmenite, rutile, zircon) within a prograding beachridge at Cudgen, New South Wales at or 1059

above the high water mark also provide evidence for a stable sea level over the period of 1060

sediment accumulation (Thom and Roy, 1985). The debate as to whether a smoothly falling 1061

sea level, a stable and prolonged highstand, or a highstand punctuated by oscillations 1062

occurred is currently unresolved and will continue until the precise indicative meaning and 1063

quality of existing data are better understood or new data are acquired (Baker et al., 2001a, b; 1064

Sloss et al., 2007; Perry and Smithers, 2011). In any case, the documented changes in the 1065

ecology, morphology, elevation and δ18O composition of tubeworm deposits (Baker and 1066

Haworth, 2000a; Baker et al., 2001a, b) are likely to provide important insights into changing 1067

water level, climate, wave exposure, ENSO and Pacific Decadal Oscillation. Indeed, there is 1068

potential that such indicators are recording short-term transient spikes from ENSO-related 1069

steric influences that are not recorded by ‘slow response systems’ such as mangrove deposits. 1070

1071

5.4 South Australia 1072

The most comprehensive sea-level compilation for South Australia was derived largely from 1073

a series of sediment cores and backhoe excavations along Spencer Gulf and the Gulf of St 1074

Vincent by Belperio et al. (2002). Transitions in sedimentary facies (seagrass, sandflat, 1075

mangrove, samphire and chenier ridge) coinciding with various tidal levels were recognised, 1076

measured relative to present sea-level position, and dated. As with the other sections of the 1077

Australian coast, the dataset has been compiled from over 25 years of research and contains 1078

an overview of the earlier studies, particularly those by Burne (1982), Belperio (1993), 1079

Belperio et al. (1983, 1984), Short et al., (1986), and Harvey et al. (1999). The data show that 1080

present sea level was reached between 8000 and 7500 cal. yr BP (original 14C data in 1081

Belperio et al., 2002 and also in Bowman and Harvey, 1986). 1082

1083

The elevation of the mid-Holocene highstand was variable along Spencer Gulf and the Gulf 1084

of St Vincent with higher sea levels (up to + 3 m) being recorded at the northern-most sites 1085

(Redcliff) farthest from the continental shelf-edge. Lower highstands (+ 1 m) were recorded 1086

at the more southern locations (Port Lincoln; Belperio et al., 2002; Fig. 6). While the earlier 1087

work in this region by Belperio and colleagues indicated that recent neotectonic activity 1088

within the Torrens Hinge Zone adjacent to the Flinders Ranges explained the highstand and 1089

Page 36: Post-glacial sea-level changes around the Australian

33

late Holocene sea-level fall (Belperio et al., 1983, 1984), this interpretation has since changed 1090

to account for the influence of hydro-isostasy (Chappell, 1987; Belperio et al., 2002). In fact, 1091

the dataset provides some of the strongest evidence for hydro-isostasy in the Australian 1092

region where it closely matches the modelled outputs (Nakada and Lambeck, 1989; Lambeck 1093

and Nakada, 1990). 1094

1095

The ‘earlier’ sea-level curves for Spencer Gulf suggested either a stepped or smooth sea-level 1096

fall over the mid-late Holocene (Burne, 1982; Belperio et al., 1983, 1984). Earlier evidence 1097

of a +3.8 m sea-level highstand for Spencer Gulf (Burne, 1982; Belperio et al., 1983) has 1098

since been revised down to +3 m (Belperio et al., 2002). Belperio et al. (2002) used only the 1099

most reliable data, conducted a more detailed statistical analysis, and highlighted the 1100

importance of separating the data to reflect geographical location. They showed that the mid-1101

late Holocene sea-level fall could be explained by either smooth-fall (linear trend) or an 1102

oscillating fall (polynomial fit); however, the polynomial analysis favoured ‘gaps’ in the data 1103

rather than genuine points and, based on the simple progradational coastal sedimentary 1104

deposits, they favoured the smooth-fall as the most likely option (Belperio et al., 2002). 1105

1106

5.5 Western Australia 1107

The sites that Fairbridge (1961) originally described in Western Australia remain important in 1108

investigations of Holocene sea-level changes; however, more recent research has included 1109

coral core records from the Houtman Abrolhos Islands (Eisenhauer et al., 1993) and reef-flat 1110

pavements (Collins et al., 2006). Other palaeo-sea level records from Western Australia 1111

include ‘swash zone’ facies from coastal plain deposits in the Perth Basin (Semeniuk, 1985, 1112

1996; Searle and Woods, 1986), and intertidal tubeworm deposits largely from Rottnest 1113

Island (Playford, 1988; Baker et al., 2005). The coral core data indicate that sea level was 1114

within 2 m of its modern level by 7100 ± 70 cal. yr BP (Eisenhauer et al., 1993) while swash 1115

zone deposits place sea level at modern levels by 8015 ± 230 cal. yr BP (original 14C age in 1116

Semeniuk, 1996). The time lag between the coral data and the other indicators is unsurprising 1117

as subtidal corals can lag sea level by long periods (e.g. Hopley, 1986a; Hopley et al., 2007). 1118

Unless the corals have formed microatoll or pavement morphology, they may be from a large 1119

depth range and should only be considered as directional sea-level indicators. 1120

1121

The reported magnitude of the mid-Holocene highstand is variable along the Western 1122

Australian coastline, with estimates of +2.0 to 3.5 m at Becher/Rockingham, Perth Basin 1123

Page 37: Post-glacial sea-level changes around the Australian

34

(swash zone facies: Semeniuk, 1985; Searle and Woods, 1986; Searle et al., 1988) and at 1124

Rottnest Island (tubeworms: Baker et al., 2001b, 2005), +2.05 m at the Houtman Abrolhos 1125

Islands (coral pavements: Collins et al., 2006), and +1.5 m in Shark Bay (shell material: 1126

Logan et al., 1970). However, the precise indicative meaning for most of these indicators is 1127

unresolved. Nevertheless, there is a degree of concordance between the tubeworm data from 1128

Rottnest Island (+2.2 m: Baker et al., 2005) and coral pavement data from the Houtman 1129

Abrolhos Islands (+2.0 m: Collins et al., 2006), which may provide an upper constraint on the 1130

elevation of the mid-Holocene highstand (Fig. 7). 1131

1132

The three most complete records of mid to late Holocene sea-level change in Western 1133

Australia are from swash zone facies (Searle and Woods, 1986), tubeworm deposits (Baker et 1134

al., 2005) and coral pavements (Collins et al., 2006) (Fig. 7). The swash zone facies and coral 1135

pavements are interpreted to represent a smooth sea-level fall from +2.0 to 2.5 m to present 1136

levels after 1000 cal. yr BP (Searle and Woods, 1986; Collins et al., 2006). In contrast, the 1137

tubeworm deposits have been interpreted to indicate an oscillating sea level (Baker et al., 1138

2005). However, these conflicting interpretations may simply reflect different statistical 1139

treatment of the data (see Belperio et al., 2002). Semeniuk and Searle (1986) reported 1140

completely different sea-level reconstructions (i.e. no highstand, smooth-fall, highstand then 1141

fall) for three sites separated only by 170 km within the Perth Basin. These discrepancies 1142

were thought be a function of local tectonism, although Baker et al. (2005) reported relatively 1143

consistent sea-level heights (approximately +1.5 m) from tubeworm deposits along the south-1144

western coastline. The accuracy of swash zone facies as sea-level indicators needs to be 1145

resolved as their uncertainties may be too large to produce high quality reconstructions. 1146

Wyrwoll et al. (1995) argued that these swash zone deposits are influenced by coastal 1147

morphodynamic conditions (in particular hydrodynamic controls) that can change over time 1148

and are also variable even across this particular stretch of coastline. 1149

1150

In northern Western Australia, mangrove muds that were deposited during the mid-Holocene 1151

‘big swamp’ phase also marked the time sea level stabilised around its present level 1152

(Woodroffe et al., 1985). Similar exposures of mangrove stumps were described from the 1153

tidal flats flanking the Ord River (Wright et al., 1972), but they remain undated. The emerged 1154

tidal flat deposits of the Ord River estuary is interpreted to have formed during a period 1155

where the adjacent embayment would have been much deeper (prior to the development of 1156

the Ord Delta and the progradation of coastal sediments) and as such the tidal 1157

Page 38: Post-glacial sea-level changes around the Australian

35

amplitude/wavelength would have been much greater at that time (Wright et al., 1972). The 1158

stratigraphy and sedimentology of the tidal flats flanking King Sound (tidal range > 12 m) 1159

have been described in detail by Semeniuk (1980a, b, 1981, 1982). His descriptions included 1160

reference to mangrove muds with abundant organic remains including in situ stumps, which 1161

he assigned to the Christine Point Clay. In early studies this clay was considered Pleistocene 1162

in age. However, radiocarbon ages extending back to 8415 ± 1740 cal. yr BP (original 14C 1163

age in Jennings, 1975) on similar muds in the adjacent Fitzroy River estuary suggest a 1164

Holocene age. Radiocarbon ages on wood fragments back to 7630 ± 200 cal. yr BP from 1165

Cambridge Gulf, adjacent to the Ord River estuary (originally reported in Thom et al., 1975) 1166

also suggest a mid-Holocene heritage. 1167

1168

5.6 Other locations 1169

Detailed time-series observations of Holocene sea-level changes in Victoria or Tasmania are 1170

rare. Gill (1983) reported highstands around +2.0 m in Victoria, and Haworth et al. (2002) 1171

described a relict tubeworm deposit at +1.5 m at Wilson’s Promontory dated at 5570 ± 250 1172

cal. yr BP (original 14C age in Haworth et al., 2002). The emerged shell bed evidence from 1173

the Gippsland Lakes presented by Gill (1983; Gill and Hopley, 1972) has since been 1174

explained by Thom (1984b) to reflect locally elevated water levels due to changing 1175

morphodynamics in the region (i.e. prior to the closure of the lagoon). Indeed, Thom (1984b) 1176

showed that the morphostratigraphical relationships of coastal sand barrier deposits in the 1177

Gippsland area are strongly influenced by changing energy conditions (wind and waves), 1178

sediment supply and sea level. For example, washover deposits become less common (or 1179

cease) in the stratigraphical record as the shorelines prograde seaward. 1180

1181

Bowden and Colhoun (1984) suggested that there was no clear evidence of a mid-Holocene 1182

highstand in Tasmania but Baker et al. (2001b) inferred a +0.5 m sea level from a relict 1183

tubeworm deposit at King Island dated at 2400 ± 330 cal. yr BP, and Colhoun (1983) 1184

cautiously suggested that shell material in a raised gravel beach deposit on Flinders island 1185

with an age of 3450 ± 400 cal. yr BP may indicate a highstand at +1.8 m. However, 1186

foraminiferal evidence (two dates: 4620 ± 330 and 6170 ± 280 cal. yr BP) from south-eastern 1187

Tasmania suggest that sea levels in the mid-Holocene were no higher than present (Gehrels et 1188

al., 2012). Murray-Wallace and Goede (1995) reported radiocarbon ages on the intertidal 1189

mollusc Katelysia sp. from back-barrier lagoon and embayment fill successions on Flinders 1190

Island. A Katelysia specimen from Cameron Inlet at + 0.5 m AHD yielded an age of 3600 ± 1191

Page 39: Post-glacial sea-level changes around the Australian

36

110 cal. yr BP (SUA-3001). Additional radiocarbon ages on Katelysia sp. ranging between 1192

5580 ± 80 (SUA-3010) and 2060 ± 100 (SUA-3015) were based on specimens collected at 1193

AHD suggesting that sea level was relatively stable over this period. Jack Davies reported 1194

seaward sloping raised beach ridge sequences from Tasmania as evidence for higher 1195

Holocene sea level , but the ages and exact elevations were not detailed (Davies, 1959, 1961). 1196

1197

5.7 Summary 1198

The evidence for higher Holocene sea level that Fairbridge described from Western Australia 1199

appeared from early radiocarbon ages to have occurred at different times within the past few 1200

millennia. His sea-level observations plotted with results from elsewhere in the world, 1201

produced a so-called global eustatic curve in the form of a succession of oscillations of sea-1202

level rising above and falling below the PMSL. Recognising the geographical differences 1203

between the pattern of decelerating relative sea level experienced in the Caribbean and their 1204

‘Australian sea-level curve’, Thom and Chappell (1975) decoupled areas with very different 1205

glacio-isostatic behaviour, and emphasised that the oscillations were not real, but an artefact 1206

of compiling data from a wide geographical range and differing relative sea-level histories. 1207

There has been a resurgence of the view that sea level has oscillated over a smaller amplitude 1208

during the mid-late Holocene (see Woodroffe and Horton, 2005; Baker and Haworth, 2000b; 1209

Lewis et al., 2008). Baker et al. (2005) specifically revisited Fairbridge’s sites from Western 1210

Australia and with newly acquired radiocarbon ages from tubeworms, demonstrated that an 1211

oscillating curve could be fitted to their available data. A smoothly falling interpretation is 1212

considered preferable for dated reef-flat evidence from the Abrolhos Islands and also from 1213

other parts of Western Australia (Searle and Woods, 1986; Collins et al., 2006). 1214

1215

The ‘early’ sea-level studies used uncorrected (or ‘environmentally corrected’; i.e. marine 1216

reservoir effect) 14C ages; these suggested that sea level attained its modern elevation 1217

between 5500 and 6500 years BP (Hopley, 1983a). It has become increasingly clear based on 1218

calibrated 14C, U-series and AAR ages that sea level reached present levels somewhat earlier, 1219

between 7500 and 8000 cal. yr BP. 1220

1221

Despite the reports of a +3 m mid-Holocene sea level by Fairbridge (1961 and references 1222

therein) in Western Australia, early sea-level research focused on whether Holocene sea 1223

levels were higher than present with research concentrated along the east coast of Australia 1224

(e.g. Hails, 1965; Thom et al., 1969, 1972; Gill and Hopley, 1972; Cook and Polach, 1973; 1225

Page 40: Post-glacial sea-level changes around the Australian

37

Belperio, 1979; Jones et al., 1979; Hopley, 1980). The sea-level evidence at this time was 1226

largely centred around relatively low-resolution indicators (beachrock, raised shell beds, 1227

freshwater peats, mangrove deposits) until the studies from the Great Barrier Reef Expedition 1228

and John Chappell’s (1983) work on coral microatolls confirmed a +1.0 to 1.5 m mid-1229

Holocene sea level on the inner shelf in north-east Queensland. Whereas most studies now 1230

recognise that sea levels around most parts of mainland Australia in the mid-Holocene 1231

reached between 1 and 2 m above present levels (e.g. Baker and Haworth, 2000a; Baker et 1232

al., 2005; Sloss et al., 2007; Lewis et al., 2008; Perry and Smithers, 2011), studies purporting 1233

a higher level between 2 and 3 m cannot be discounted entirely (e.g. Hopley 1971, 1975, 1234

1978, 1980; Searle and Woods, 1986; Woodroffe, 2009). While some of this variation in 1235

magnitude may be explained by the geographical variability in the hydro-isostatic response 1236

around the Australian margin, further research is required to resolve the uncertainties 1237

associated with the various proxy sea-level indicators used (i.e. beachrock, swash zone facies 1238

and foraminiferal transfer function) before they can be fully integrated into sea-level 1239

reconstructions. In particular, these indicators need to be avoided where their elevations do 1240

not concur with the more established coral microatolls and encrusting organisms at the same 1241

location. The Gulfs in Southern Australia record magnitudes up to + 3 m (Belperio et al., 1242

2002) and similar magnitudes have also been reported for Western Australia (e.g. Searle and 1243

Woods, 1986). Such variations in elevation and timing of changes in sea-level trends reveal 1244

that it is inappropriate to derive a single Holocene sea-level curve for the entire coastal 1245

margin of Australia and that there is regional variability in sea-level histories due to different 1246

isostatic influences, antecedent geomorphology of coastal margins, shelf characteristics, and 1247

oceanographic and climatic controls. Accordingly, the regional variability in sea-level 1248

histories as well as the subtle variations in the indicative meanings of different sea-level 1249

indicators in different settings and locations validates the construction of regional sea-level 1250

envelopes rather than a single sea-level curve. 1251

1252

The nature of the mid-late Holocene sea-level fall around mainland Australia is perhaps the 1253

most contentious issue remaining to be resolved. Proponents are split between a smoothly-1254

falling sea level, a prolonged and stable mid-late Holocene highstand followed by a later fall, 1255

and an oscillating mid-late Holocene sea level. In essence, the debate is centred on two 1256

different types of sea-level indicators that include those that are part of accretionary 1257

landforms, such as prograding coastal and estuarine sedimentary successions, or coral reefs, 1258

coral microatolls on reef flats, coral pavements and unconformities in coastal plain 1259

Page 41: Post-glacial sea-level changes around the Australian

38

successions; these indicators are more likely to record a relative sea-level fall. The other 1260

indicators include encrusting organisms, where numerous ages on scattered remnant deposits 1261

are suited to record a prolonged highstand with possible superimposed intermittent 1262

oscillations. 1263

1264

Clearly, continuous location-specific sea-level reconstructions based on a range of different 1265

types of sea-level indicators are highly desirable, and are more likely to credibly resolve 1266

relative sea-level history around the Australian margin. Indeed, indicators that record a 1267

regional pattern related to hydro-isostasy must be separated from those that record changes in 1268

local water levels produced by changes in coastal geomorphology over time. However, such 1269

diverse multi-proxy records remain elusive, and may remain so. Such a reconstruction is 1270

problematic because open water coral microatoll data for the past ca. 4000 years is sparse for 1271

north-eastern Australia (Lewis et al., 2008), probably reflecting the age structure of host reefs 1272

and the limited accommodation space for unmoated microatoll establishment once mature 1273

reef flats have been established (Smithers et al., 2006; Perry and Smithers, 2011). The 1274

encrusting organisms (particularly tubeworms) are well represented over this period (see 1275

Lewis et al., 2008) and indicate a prolonged sea-level highstand. While clear geographical 1276

transitions in biostratigraphical assemblages coupled with δ18O fluctuations in the tubeworm 1277

data as well as some shifts in vertical elevation undoubtedly coincide with climatic and 1278

environmental changes (e.g. sea temperature and salinity), it remains disputed whether these 1279

changes represent genuine sea-level oscillations (see Lambeck et al., 2010; Perry and 1280

Smithers, 2011). Furthermore, oscillations in the tubeworm data as postulated by Baker et al. 1281

(2001a, 2005) could be the result of changing wave exposure regimes over the Holocene 1282

(Goodwin, 2003, Sloss et al., 2007) or reflect specific climate/ocean interaction (e.g. ENSO 1283

and steric effect of the East Australian Current) which are known to influence the elevations 1284

of such deposits (Laborel and Laborel-Deguen, 1996). The larger scale oscillations (~1 m) 1285

proposed by Baker and Haworth (2000b; Baker et al., 2001a, b, 2005) using their original 1286

tubeworm data (and re-assessed in Sloss et al., 2007) need to be discriminated from statistical 1287

artefacts where ‘gaps’ in the data are weighted towards oscillations (e.g. Belperio et al., 2002; 1288

Perry and Smithers, 2011). 1289

1290

6 Modelling Holocene sea-level changes 1291

Recognition that various locations around the world have different relative Holocene sea-1292

level histories and that there is no single ‘global’ solution (e.g. the work of the International 1293

Page 42: Post-glacial sea-level changes around the Australian

39

Geological Correlation Programme; Bloom, 1977; Hopley, 1983a), prompted other 1294

explanations for varying sea-level patterns including tectonic and isostatic influences. A key 1295

breakthrough came when geophysical modellers revisited an early hypothesis that the land 1296

could subside and rebound with increased and decreased ice and water loads (Walcott, 1972; 1297

Clark et al., 1978; Clark and Lingle, 1979); these models provided a powerful tool to explain 1298

relative sea-level changes around the world. Walcott (1972) highlighted the different 1299

responses that would be expected based on the location of former ice sheets coupled with sea-1300

level indicator data from around the world. Further models to predict sea-level curves for 1301

more specific locations were developed by Chappell (1974), Clark et al. (1978), Clark and 1302

Lingle (1979) and Nakiboglu et al. (1983) where the contribution of hydro-isostasy for 1303

locations in the ‘far-field’ of former ice sheets (relevant for most parts of Australasia) were 1304

better quantified. However, the contribution of hydro-isostatic factors along the Australian 1305

mainland were not quantified until Chappell et al. (1982) and later Nakada and Lambeck 1306

(1989) and Lambeck and Nakada (1990) completed specific model predictions for this 1307

region. In the following discussion, we review these models (and later revisions) noting 1308

specific comparisons with the sea-level indicator data as well as examining other potential 1309

influences of mid-late Holocene sea-level changes. 1310

1311

6.1 Hydro-isostatic adjustments 1312

The concept of hydro-isostasy relates to the viscoelastic response or flexure of the continental 1313

crust and mantle to varying water loads. Thus the key parameters (that are most difficult to 1314

quantify) within the models include mantle viscosity and lithospheric thickness as well as the 1315

‘ice’ model such as the timing of meltwater influx to ocean basins and when it reduced or 1316

ceased (Lambeck and Nakada, 1990). In fact, the glacio-hydro-isostatic model of Lambeck 1317

and Nakada (1990) originally predicted that sea level reached its present position between 1318

9000 and 7000 cal. yr BP which is in accord with the present concensus as discussed 1319

previously in this review. The hydro-isostatic model provides insights into the melting of 1320

different ice sheets and the volumes required to produce sea-level changes of the magnitude 1321

observed in the post-glacial sea-level rise. The Lambeck and Nakada (1990) model was the 1322

first to examine changes at far-field sites that combined the regional (water loading on shelf) 1323

and global data (i.e. glacio-isostatic changes). 1324

1325

The predictions of a smooth sea-level fall for the far-field sites stemmed from both 1326

observational evidence as well as the input data based on a negligible contribution from ice 1327

Page 43: Post-glacial sea-level changes around the Australian

40

melting after 7000 to 6000 cal. yr BP. In particular, Chappell et al. (1982) ‘calibrated’ the 1328

hydro-isostatic model to observational data from Karumba (chenier ridges) and the Great 1329

Barrier Reef (coral microatolls) by matching the ‘smooth-fall’ response apparent in these 1330

indicators. They also introduced the concept of a ‘hinge’ response where the continental shelf 1331

margin down-warps to accommodate lower relative sea levels on the outer shelf and a 1332

corresponding upward flexure occurs in coastal areas, producing a relatively higher sea level 1333

on the inner shelf. The slow deformation of the crust-mantle system in the continental shelf 1334

and ocean margin areas following the post-glacial transgression results in a smooth sea-level 1335

fall along coastlines that are located adjacent to relatively wide shelves. Variations in the 1336

shelf width and depth result in variable mid-late Holocene sea-level histories; for example the 1337

highstands at Cairns, Queensland where the shelf width is around 80 km is much lower than 1338

the southern end of the Great Barrier Reef where the shelf extends to ~ 300 km offshore. 1339

Coral microatoll (up to +0.95 m: Chappell et al., 1983) and oyster bed (+1.2 m: Higley, 2000; 1340

Lewis et al., 2008) data from Princess Charlotte Bay (north of Cairns) show lower elevations 1341

compared with their southern counterparts (+1.4 and 1.6 m, respectively). The smooth sea-1342

level fall from highstands of varying magnitudes depending on shelf width is also supported 1343

by the dataset from Spencer Gulf and the Gulf of St Vincent, South Australia, where the most 1344

inland locations coincide with the highest emergence (Belperio et al., 2002). We note, 1345

however, that some researchers are less convinced that the magnitude of the mid-Holocene 1346

highstand always varies with the width of the eastern Australian continental shelf (see 1347

Haworth et al., 2002). 1348

1349

Interestingly, adjustments to the timing of the meltwater contribution (i.e. continuing after 1350

6000 to 7000 years BP) in the hydro-isostatic models will produce a prolonged mid-Holocene 1351

sea-level highstand (see Thom and Chappell, 1978; Lambeck, 2002; Peltier, 2002). Goodwin 1352

(1998) suggested that changes in Antarctic ice melt volumes during the Holocene could have 1353

also potentially affected late-Holocene sea-level fall. Indeed, the continued influx of 1354

meltwater was postulated by Sloss et al. (2007) and Woodroffe (2009) to account for the 1355

prolonged sea-level highstand observed in their datasets from eastern Australia. 1356

1357

While the hydro-isostatic modelling appears to yield outputs that agree with the regional 1358

datasets (e.g. Searle and Woods, 1986; Belperio et al., 2002; Collins et al., 2006), there are 1359

also discrepancies and anomalies in some regions. For example, mid-Holocene highstands on 1360

Rottnest Island and southern Australia (Haworth et al., 2002; Baker et al., 2005), and possibly 1361

Page 44: Post-glacial sea-level changes around the Australian

41

also Lord Howe Island (Woodroffe et al., 1995) do not match the predicted hydro-isostatic 1362

response (Nakada and Lambeck, 1989; Lambeck and Nakada, 1990; Lambeck, 2002). Local 1363

tectonic influences have been implied for south-western Australia including Rottnest Island 1364

(e.g. Semeniuk and Searle, 1986; Playford, 1988; Nakada and Lambeck, 1989; Lambeck and 1365

Nakada, 1990), although these have been questioned by Baker et al. (2005) who found 1366

similar magnitude highstands along the Western Australian coastline and at Rottnest Island. 1367

Moreover, regional tectonic uplift along southern Australia and Tasmania (see Bryant, 1992; 1368

Murray-Wallace and Goede, 1995; Murray-Wallace, 2002) may be superimposed on the 1369

hydro-isostatic influence (e.g. Lambeck and Nakada, 1990). Indeed, Baker et al. (2001) and 1370

Haworth et al. (2002) found higher sea levels of +0.5 m at King Island and +1.5 m at 1371

Wilson’s Promontory, both locations where hydro-isostatic modelling suggests there should 1372

be no highstand. 1373

1374

At this stage, there are inadequate observational spatial and temporal datasets (termed 1375

‘fragmentary’) to satisfactorily inform the hydro-isostatic and neotectonic models (Lambeck 1376

and Nakada, 1990; Lambeck, 2002). While many valuable location-specific datasets have 1377

been produced, many of which have been used to ‘calibrate’ the model, a continuous high-1378

resolution reconstruction remains elusive. Only when the apparent discrepancies between the 1379

‘prograding and accretionary’ indicators (smooth-fall) and the encrusting organisms 1380

(prolonged highstand) are reconciled and the precise indicative meaning of included 1381

indicators is resolved can modelling be performed to account for the addition of extra 1382

meltwater during the late Holocene. 1383

1384

6.2 Other influences 1385

It has been postulated from those in favour of possible ~1 m sea-level oscillations in the mid-1386

late Holocene that they were produced by the break-up of ice sheets in either the north 1387

Atlantic (‘Bond Cycles’) or Antarctica (e.g. Baker et al., 2005; Lewis et al., 2008). Edmund 1388

Gill was another supporter of eustatic mid-late Holocene sea-level changes, and while he 1389

acknowledged the possible influence of hydro-isostasy he was a strong advocate for 1390

oscillating sea levels concluding his 1983 paper with “It is no longer a question of whether 1391

oscillations occurred, but what the oscillations were” (Gill, 1983, p. 63). Other studies have 1392

dismissed larger (~ 1 m) oscillations but consider the possibility of smaller-scale fluctuations 1393

related to climate forcing such as ENSO, Pacific Decadal Oscillation and changes in wave 1394

climate (e.g. Chappell, 1987; Goodwin, 2003; Sloss et al., 2007). These smaller scale changes 1395

Page 45: Post-glacial sea-level changes around the Australian

42

in sea level would be superimposed on the eustatic and isotactic sea-level history and thus 1396

add complications in using proxy sea-level indicators that may be influenced by such ocean-1397

atmospheric influences. 1398

1399

‘Recent’ tectonic movements have also been invoked to explain variations in the magnitude 1400

of the mid-Holocene highstand along the Queensland coast (e.g. Hopley, 1975) and in South 1401

Australia (Belperio et al., 1983, 1984), although latter work has shown that the magnitude 1402

variability in the South Australian evidence is due to hydro-isostasy (Belperio et al., 2002). 1403

Research on chenier plain deposits at Broad Sound, Queensland highlighted the possibility of 1404

local tectonic movements, where deposits on the western side of the Sound suggest no mid-1405

Holocene highstand (Cook and Polach, 1973) but those on the eastern side suggest a 1406

highstand perhaps as large as 4-6 m above PMSL (Cook and Mayo, 1977). 1407

1408

Bryant (1992) reviewed the variable sea-level highstands of the last interglacial (based on the 1409

analysis of Murray-Wallace and Belperio, 1991) and mid-Holocene around Australia and 1410

found that there was possible down warping of northern Australia and up warping along the 1411

southern edge of the continent (including Tasmania). Most of the east coast of New South 1412

Wales and west coast of Western Australia were classed as relatively stable. Byrant (1992) 1413

argued this tectonic influence also explained the apparent variability in mid-Holocene sea-1414

level maxima across Australia and contrasted the findings of the Lambeck and Nakada (1990) 1415

models which suggested higher sea levels to the north of the continent. However, Nott (1996) 1416

subsequently discovered evidence of a raised (approximately +3 m) fossil in situ reef 1417

believed to be of last interglacial age (although diagenesis prevented reliable dating of the 1418

deposit), suggesting that subsidence of the northern Australia margin to the degree predicted 1419

by Bryant (1992) is unlikely. 1420

1421

Another possibility that would influence the reliability of sea-level indicators is historical 1422

changes in the tidal range. The most precise sea-level indicators grow to a specific sea-level 1423

datum such as coral microatolls (approximately mean low water springs) and encrusting 1424

organisms (approximately high water neaps). As such, a change in tidal range could account 1425

for discrepancies between these indicators. Cook and Mayo (1977) demonstrated that there 1426

had been a change in the tidal range at Broad Sound of close to 1 m in the past 6000 years, 1427

although the biggest absolute vertical changes are likely to occur where tidal ranges are large 1428

(e.g. Broad Sound has up to a 9 m tidal range). 1429

Page 46: Post-glacial sea-level changes around the Australian

43

1430

7 Conclusions 1431

Fairbridge’s pioneering research led, not to a global eustatic curve as he had anticipated, but 1432

to the recognition that the pattern of relative sea-level change in the Australian region 1433

differed from that observed in the Atlantic. A series of seminal sea-level studies were 1434

undertaken in the following 25 years. The stabilisation of sea level close to its present 1435

elevation in the mid-Holocene set the scene for the detailed reconstructions that were 1436

undertaken at different locations around the Australian mainland. The comprehensive reviews 1437

by Hopley (1983a) and Thom (1984a) summarise the state of knowledge about sea level and 1438

geomorphology around the Australian coast at that time. These early studies provide a solid 1439

foundation that remains the basis for current data compilations of sea-level studies in 1440

Australia over the ensuing 25 years (Hopley, 1982; Hopley et al., 2007). However, additional 1441

carefully selected and measured indicator data from specific locations are required to further 1442

refine the nature of mid-late Holocene sea-level changes at various locations. 1443

1444

Significant progress has been made to extend sea-level research in regard to the post-glacial 1445

sea-level rise, including research on the Huon Peninsula (e.g. Chappell and Polach, 1991), 1446

Joseph Bonaparte Gulf (Yokoyama et al., 2000, 2001a) and the Sunda Shelf (Hanebuth et al., 1447

2000, 2009). Preliminary regional compilations and syntheses of sea-level data have been 1448

produced for north-eastern Australia (Larcombe et al., 1995; Lewis et al., 2008), New South 1449

Wales (Sloss et al., 2007) and South Australia (Belperio et al., 2002). There is broad 1450

agreement about a sea-level envelope encompassing a range of evidence within a broad swath 1451

of data variability, but there remains active debate about the specific details of post-glacial, 1452

and particularly Holocene sea-level behaviour at any one site. Establishing broader regional 1453

patterns of relative sea-level change has long been a goal of those researchers who have 1454

compiled datasets, but it is clear that sea level has behaved differently at different localities 1455

around our coast. The calibration of 14C ages has allowed previous datasets to be directly 1456

aligned by accounting for global and regional marine reservoir effects and places sea level 1457

attaining modern elevations between 7500 and 8000 cal. yr BP around the coastal margin of 1458

Australia. Depending on the indicators of preference, sea level either fell smoothly from a +1 1459

to 2 m highstand or remained at these levels for a considerable period before falling or 1460

oscillating to its present position. 1461

1462

Page 47: Post-glacial sea-level changes around the Australian

44

Clearly, the selection of sea-level data for a synoptic assessment of sea-level change requires 1463

a thorough understanding of the limitations and integrity of each data point as accurate 1464

palaeo-sea-level indicators. However, the inherent uncertainty of the data can be somewhat 1465

difficult to estimate from the previous studies (i.e. the precision and accuracy of the elevation 1466

measurement to tidal datum, whether the sample was in situ). A way forward may be the 1467

development of systematic criteria in the selection of data points to ensure the quality of such 1468

compilations. It is recommended that sea-level studies in the future should ground truth 1469

indicators to sea-level datum using laser levelling technology, high precision GPS or satellite 1470

altimetry. Moreover, the key to quality sea-level reconstructions is a complete understanding 1471

of the indicative meaning, limitations and reproducibility of each indicator. The influence of 1472

tidal range and variability, and short-term climatic fluctuations (such as ENSO) on the 1473

various sea-level indicators also needs to be understood so that short-term and long-term 1474

deviations can be discriminated. 1475

1476

Great progress has been made since Fairbridge’s initial compilation that has improved the 1477

knowledge of post-glacial sea-level change around the Australian continent, but there is still 1478

much to be resolved. A clearer understanding of past sea-level changes and their causes is 1479

urgently needed to better inform our ability to forecast future changes. A concerted effort is 1480

required, through the compilation of existing data, renewed fieldwork, dating analysis and 1481

modelling to address the issues of whether there have been oscillations of the sea surface and 1482

if so, of what magnitude. The pattern and rate of fall from the Holocene highstand to modern 1483

levels, and of the contributions of the various factors to this change, both global ‘eustatic’ or 1484

‘steric’ components and local geophysical, tectonic and land instability issues also need to be 1485

addressed (see Belperio, 1993). 1486

1487

8 Acknowledgements 1488

This research was conducted with the support of funding from the Australian Government’s 1489

National Environmental Research Program grant to SEL and SGS for project 1.3 1490

‘Characterising the cumulative impacts of global, regional and local stressors on decadal, 1491

centennial and millennial scales to quantify the effects of European settlement on the 1492

biodiversity of GBR coral reef communities’. This review represents a contribution to the 1493

INQUA Commission of Coastal and Marine Processes, and the INQUA1001 Late Quaternary 1494

records of sea-level change. CDW and SGS were supported by ARC discovery grant 1495

DP0773019. SEL acknowledges the support of the Australian Institute of Nuclear Science 1496

Page 48: Post-glacial sea-level changes around the Australian

45

and Engineering for providing travel assistance to attend OZ-INTIMATE group meetings. 1497

We are very grateful to two anonymous reviewers who made excellent comments and greatly 1498

improved the manuscript. 1499

1500

9 References 1501

Abbey, E., Webster, J.M., Beaman, R.J., 2011. Geomorphology of submerged reefs on the 1502 shelf edge of the Great Barrier Reef: The influence of oscillating Pleistocene sea-1503 levels. Marine Geology 288, 61-78. 1504

1505 Baker, R.G.V., Haworth, R.J., 1997. Further evidence from relic shellcrust sequences for a 1506

late Holocene higher sea level for eastern Australia. Marine Geology 141, 1-9. 1507 1508 Baker, R.G.V., Haworth, R.J., 2000a. Smooth or oscillating late Holocene sea-level curve? 1509

Evidence from the palaeo-zoology of fixed biological indicators in east Australia and 1510 beyond. Marine Geology 163, 367-386. 1511

1512 Baker, R.G.V., Haworth, R.J., 2000b. Smooth or oscillating late Holocene sea-level curve? 1513

Evidence from cross-regional statistical regressions of fixed biological indicators. 1514 Marine Geology 163, 353-365. 1515 1516

Baker, R.G.V., Haworth, R.J., Flood, P.G., 2001a. Warmer or cooler late Holocene marine 1517 palaeoenvironments?: interpreting southeast Australian and Brazilian sea-level 1518 changes using fixed biological indicators and their δ18O composition. 1519 Palaeogeography, Palaeoclimatology, Palaeoecology 168, 249-272. 1520

1521 Baker, R.G.V., Haworth, R.J., Flood, P.G., 2001b. Inter-tidal fixed indicators of former 1522

Holocene sea levels in Australia: a summary of sites and a review of methods and 1523 models. Quaternary International 83-85, 257-273. 1524

1525 Baker, R.G.V., Haworth, R.J., Flood, P.G., 2005. An oscillating Holocene sea-level? 1526

Revisiting Rottnest Island, Western Australia, and the Fairbridge eustatic hypothesis. 1527 Journal of Coastal Research Special Issue 42, 3-14. 1528

1529 Beaman, R., Larcombe, P., Carter, R.M., 1994. New evidence for the Holocene sea-level high 1530

from inner shelf central Great Barrier Reef, Australia. Journal of Sedimentary 1531 Research A64, 881-885. 1532

1533 Beaman, R.J., Webster, J.M., Wust, R.A.J., 2008. New evidence for drowned shelf edge reefs 1534

in the Great Barrier Reef, Australia. Marine Geology 247, 17-34. 1535 1536 Belperio, A.P., 1979. Negative evidence for a mid-Holocene high sea level along the coastal 1537

plain of the Great Barrier Reef province. Marine Geology 32, M1-M9. 1538 1539 Belperio, A.P., 1993. Land subsidence and sea level rise in the Port Adelaide estuary: 1540

implications for monitoring the greenhouse effect. Australian Journal of Earth 1541 Sciences 40, 359-368. 1542

1543

Page 49: Post-glacial sea-level changes around the Australian

46

Belperio, A.P., Hails, J.R., Gostin, V.A., 1983. A review of Holocene sea levels in South 1544 Australia, In: Hopley, D. (Ed.), Australian sea levels in the last 15 000 years: A 1545 review. Monograph Series Occasional Paper 3, Geography Department, James Cook 1546 University, Townsville, pp. 37-47. 1547

1548 Belperio, A.P., Smith, B.W., Polach, H.A., Nittrouer, C.A., DeMaster D.J., Presscott, J.R., 1549

Hails, J.R., Gostin, V.A., 1984. Chronological studies of the Quaternary marine 1550 sediments and palaeoclimatic interpretations, Spencer Gulf, South Australia. Marine 1551 Geology 61, 265-296. 1552

1553 Belperio, A.P., Gostin, V.A., Cann, J.H. and Murray-Wallace, C.V., 1988. Sediment-1554

organism zonation and the evolution of Holocene tidal sequences in southern 1555 Australia. In: de Boer, P.L., van Gelder, A., Nio, S.D. (Eds.), Tide-influenced 1556 Sedimentary Environments and Facies. Reidel, Dordrecht, pp. 475-497 1557

1558 Belperio, A.P., Harvey, N., Bourman, R.P., 2002. Spatial and temporal variability in the 1559

Holocene palaeosea-level record around the South Australian coastline. Sedimentary 1560 Geology 150, 153-169. 1561

1562 Berkeley, A., Perry, C.T., Smithers, S.G., Horton, B.P., Cundy, A.B., 2009a. Foraminiferal 1563

biofacies across mangrove-mudflat environments at Cocoa Creek, north Queensland, 1564 Australia. Marine Geology 263, 64-86. 1565

1566 Berkeley, A., Perry, C.T., Smithers, S.G., 2009b. Taphonomic signatures and patterns of test 1567

degradation on tropical, intertidal benthic foraminifera. Marine Micropaleontology 73, 1568 148-163. 1569

1570 Bird, E.C.F., 1971. The fringing reefs near Yule Point, north Queensland. Australian 1571

Geographical Studies 9, 107-115. 1572 1573 Bird, E.C.F., 1988. The tubeworm Galeolaria caespitose as an indicator of sea-level rise. 1574

Victorian Naturalist 105, 98-104. 1575 1576 Bird, M.I. Fifield, L.K., Teh, T.S., Chang, C.H., Shirlaw, N., Lambeck, K., 2007 An 1577

inflection in the rate of early mid-Holocene eustatic sea-level rise: a new sea-level 1578 curve from Singapore. Estuarine Coastal and Shelf Science 71, 523-536. 1579

1580 Bloom, A.L., 1977. Atlas of Sea-Level Curves. IGCP 61. Cornell University, Ithaca, New 1581

York. 1582 1583 Bowden, A.R., Colhoun, E.A., 1984. Quaternary emergent shorelines of Tasmania. In: Thom, 1584

B.G. (Ed.), Coastal geomorphology in Australia. Academic Press, Sydney, pp. 313-1585 342. 1586

1587 Bowman, G., Harvey, N., 1986.Geomorphic evolution of a Holocene beach-ridge complex, 1588

LeFevre Peninsula, South Australia. Journal of Coastal Research 2, 345-362. 1589 1590 Bryant, E., 1992. Last interglacial and Holocene trends in sea-level maxima around Australia: 1591

Implications for modern rates. Marine Geology 108, 209-217. 1592 1593

Page 50: Post-glacial sea-level changes around the Australian

47

Bryant, E.A., Young, R.W., Price, D.M., Short, S.A., 1992. Evidence for Pleistocene and 1594 Holocene raised marine deposits, Sandon Point, New South Wales. Australian Journal 1595 of Earth Sciences 39, 489-493. 1596

1597 Bunt, J.S., Williams, W.T., Bunt, E.D., 1985. Mangrove species distribution in relation to tide 1598

at the seafront and up rivers. Australian Journal of Marine and Freshwater Research 1599 36, 481-492. 1600

1601 Burne, R.V., 1982. Relative fall of Holocene sea level and coastal progradation, northeastern 1602

Spencer Gulf, South Australia. BMR Journal of Australian Geology and Geophysics 1603 7, 35-45. 1604

1605 Callard, S.L., Gehrels, W.R., Morrison, B.V., Grenfell, H.R., 2011. Suitability of salt-marsh 1606

foraminifera as proxy indicators of sea level in Tasmania. Marine Micropaleontology 1607 79, 121-131. 1608

1609 Cann, J.H., Belperio, A.P., Gostin, V.A., Murray-Wallace, C.V., 1988. Sea-level history 1610

45,000 to 30,000 yr B.P., inferred from benthic foraminifera, Gulf St. Vincent, South 1611 Australia. Quaternary Research 29, 153-175. 1612

1613 Cann, J.H., Belperio, A.P., Gostin, V.A., Rice, R.L., 1993. Contemporary benthic 1614

foraminifera in Gulf St Vincent, South Australia, and a refined Late Pleistocene sea-1615 level history. Australian Journal of Earth Sciences 40, 197-211. 1616

1617 Cann, J.H., Murray-Wallace, C.V., Belperio, A.P., Brenchley, A.J., 1999. Evolution of 1618

Holocene coastal environments near Robe southeastern South Australia. Quaternary 1619 International 56, 81-97. 1620

1621 Cann, J.H., Harvey, N., Barnett, E.J., Belperio, A.P., Bourman, R.P., 2002. Foraminiferal 1622

biofacies eco-succession and Holocene sea levels, Port Pirie, South Australia. Marine 1623 Micropaleontology 44, 31-55. 1624

1625 Cann, J.H., Murray-Wallace, C.V., Riggs, N.J., Belperio, A.P., 2006. Successive 1626

foraminiferal faunas and inferred palaeoenvironments associated with the postglacial 1627 (Holocene) marine transgression, Gulf St Vincent, South Australia. The Holocene 16, 1628 224-234. 1629

1630 Carter, R.M., Johnson, D.P., 1986. Sea-level controls on the post-glacial development of the 1631

Great Barrier Reef, Queensland. Marine Geology 71, 137-164. 1632 1633 Carter, R.M., Carter, L., Johnson, D.P., 1986. Submergent shorelines in the SW Pacific: 1634

evidence for an episodic post-glacial transgression. Sedimentology 33, 629-649. 1635 1636 Chappell, J., 1974. Late Quaternary glacio- and hydro-isostasy , on a layered Earth. 1637

Quaternary Research 4, 429-440. 1638 1639 Chappell, J., 1980. Coral morphology, diversity and reef growth. Nature 286, 249-252. 1640 1641 Chappell, J., 1983. Evidence for smoothly falling sea level relative to north Queensland, 1642

Australia, during the past 6,000 yr. Nature 302, 406-408. 1643

Page 51: Post-glacial sea-level changes around the Australian

48

1644 Chappell, J., 1987. Late Quaternary sea-level changes in the Australian region. In: Tooley, 1645

M.J., Shennan, I. (Eds.), Sea-level changes. The Institute of British Geographers 1646 Special Publications Series 20, pp. 296-331 1647

1648 Chappell, J., Grindrod, J., 1984. Chenier plain formation in northern Australia. In: Thom, 1649

B.G. (Ed.), Coastal Geomorphology in Australia. Academic Press, Australia, pp. 197-1650 232. 1651

1652 Chappell, J., Polach, H., 1991. Post-glacial sea-level rise from a coral record at Huon 1653

Peninsula, Papua New Guinea. Nature 349, 147-149. 1654 1655 Chappell, J., Shackleton, N.J., 1986. Oxygen isotopes and sea level. Nature 324, 137-140. 1656 1657 Chappell, J., Thom, B.G., 1977. Sea levels and coasts, In: Allen, J., Golson, J., Jones, R., 1658

(Eds.), Sunda and Sahul: prehistoric studies in Southeast Asia, Melanesia and 1659 Australia. Academic Press, London, pp. 275-291. 1660

1661 Chappell, J., Thom, B.G., 1986. Coastal morphodynamics in North Australia: review and 1662

prospect. Australian Geographical Studies 24, 110-127. 1663 1664 Chappell, J., Rhodes, E.G., Thom, B.G., Wallensky, E., 1982. Hydro-Isostasy and sea-level 1665

isobase of 5500 B.P. in North Queensland, Australia. Marine Geology 49, 81-90. 1666 1667 Chappell, J., Chivas, A., Wallensky, E., Polach, H.A., Aharon, P., 1983. Holocene Palaeo-1668

Environmental changes, central to north Barrier Reef inner zone. BMR Journal of 1669 Australian Geology and Geophysics 8, 223-235. 1670

1671 Chappell, J., Ota, Y., Berryman, K., 1996. Late Quaternary coseismic uplift history of Huon 1672

Peninsula, Papua New Guinea. Quaternary Science Reviews 15, 7-22. 1673 1674 Chappell, J., Ota, Y., Campbell, C., 1998. Decoupling post-glacial tectonism and eustasy at 1675

Huon Peninsula, Papua New Guinea. In: Stewart, I.S., Vita-Finzi, C. (Eds.), Coastal 1676 Tectonics, Geological Society, London, Special Publications 146, pp. 31-40. 1677

1678 Chivas, A., Chappell, J., Polach, H., Pillans, B., Flood, P., 1986. Radiocarbon evidence for 1679

the timing and rate of development, beachrock formation and phosphatization at Lady 1680 Elliot Island, Queensland, Australia. Marine Geology 69, 273-287. 1681

1682 Chivas, A.R., Garcia, A., van der Kaars, S., Couapel, M.J.J., Holt, S., Reeves, J.M., 1683

Wheeler,D.J., Switzer, A.D., Murray-Wallace, C.V., Banerjee, D., Price, D.M., Wang, 1684 S.X., Pearson, G., Edgar, N.T., Beaufort, L., De Deckker, P., Lawson, E.M., Cecil, 1685 C.B., 2001. Sea-level and environmental changes since the last interglacial in the Gulf 1686 of Carpentaria, Australia: an overview. Quaternary International 83-85, 19-46. 1687

1688 Clark, J.A., Lingle, C.S., 1979. Predicted relative sea-level changes (18,000 years B.P. to 1689

present) caused by late-glacial retreat of the Antarctic ice sheet. Quaternary Research 1690 11, 279-298. 1691

1692

Page 52: Post-glacial sea-level changes around the Australian

49

Clark, J.A., Farrell, W.E., Peltier, W.R., 1978. Global changes in postglacial sea level: A 1693 numerical calculation. Quaternary Research 9, 265-287. 1694

1695 Clark, T.R., Zhao, J-x., Feng, Y-x., Done, T.J., Jupiter, S., Lough, J., Pandolfi, J.M., 2012. 1696

Spatial variability of initial 230Th/232Th in modern Porites from the inshore region of 1697 the Great Barrier Reef. Geochimica et Cosmochimica Acta 78, 99-118. 1698

1699 Colhoun, E.A., 1983. Tasmanian sea level in the last 15 000 years. In: Hopley, D (Ed.), 1700

Australian sea levels in the last 15 000 years: A review. Monograph Series Occasional 1701 Paper 3, Geography Department, James Cook University, Townsville, pp. 54-58. 1702

1703 Collins, L.B., Zhao, J-X., Freeman, H., 2006. A high-precision record of mid-late Holocene 1704

sea level events from emergent coral pavements in the Houtman Abrolhos Islands, 1705 southwest Australia. Quaternary International 145-146, 78-85. 1706

1707 Cook, P.J., Polach, H.A., 1973. A chenier sequence at Broad Sound, Queensland, and 1708

evidence against a Holocene high sea level. Marine Geology 14, 253-268. 1709 1710 Cook, P.J., Mayo, W., 1977. Sedimentology and Holocene history of a tropical estuary 1711

(Broad Sound, Queensland). Bureau of Mineral Resources Geology and Geophysics 1712 Bulletin 170, 1-206. 1713

1714 Dakin, W.J., 1987. Australian Seashores 7th ed. Angus and Robertson, Sydney, 411 pp. 1715 1716 Davies, J.L., 1959. Sea level change and shoreline development in south eastern Tasmania. 1717

Papers and Proceeding of the Royal Society of Tasmania 93, 89-96. 1718 1719 Davies, J.L., 1961. Tasmanian beach ridge systems in relation to sea level change. Papers and 1720

Proceeding of the Royal Society of Tasmania 95, 35-41. 1721 1722 Davies, P.J., Montaggioni, L.F., 1985. Reef growth and sea-level change: the environmental 1723

signature. Proceedings of the 5th International Coral Reef Congress 3, 477-515. 1724 1725 De Deckker, P., Yokoyama, Y., 2009. Micropalaeontological evidence for Late Quaternary 1726

sea-level changes in Bonaparte Gulf. Global and Planetary Change 66, 85-92. 1727 1728 Desruelles, S., Fouache, E., Ciner, A., Dalongeville, R., Pavlopoulos, K., Kosun, E., 1729

Coquinot, Y., Potdevin, J.L., 2009. Beachrocks and sea level changes since Middle 1730 Holocene: comparison between the insular group of Mykonos-Delos-Rhenia 1731 (Cyclades, Greece) and the southern coast of Turkey. Global and Planetary Change 1732 66, 19-33. 1733

1734 Donner, J., Junger, H., 1981. Radiocarbon dating of marine shells from southeastern 1735

Australia as a means of dating relative sea-level changes. Annales Academy Science 1736 Fenicae Series A, 131, 5-44. 1737

1738 Edwards, R.L., Beck. J.W., Burr, G.S., Donahue, D.J., Chappell, J.M.A., Bloom, A.L., 1739

Druffel, E.R.M., Taylor, F.W., 1993. A large drop in atmospheric 14C/12C and reduced 1740 melting in the Younger Dryas, documented with 230Th ages of corals. Science 260, 1741 962-968. 1742

Page 53: Post-glacial sea-level changes around the Australian

50

1743 Eisenhauer, A., Wasserburg, G.J., Chen, J.H., Bonani, G., Collins, L.B., Zhu, Z.R., Wyrwoll, 1744

K.H., 1993. Holocene sea-level determination relative to the Australian continent: 1745 U/Th (TIMS) and 14C (AMS) dating of coral cores from the Abrolhos Islands. Earth 1746 and Planetary Science Letters 114, 529-547. 1747

1748 Eisenhauer, A., Heiss, G.A., Sheppard, C., Dullo, W.C., 1999. Reef and island formation and 1749

Late Holocene sea level changes in the Chagos islands. In: Sheppard, C.R.C., 1750 Seaward, M.R.D. (Eds.), Ecology of the Chagos Archipelago. Westbury Publishing, 1751 Otley, 350pp. 1752

1753 Endean, R., Stephenson, W., Kenny, R., 1956. The ecology and distribution of intertidal 1754

organisms on certain islands off the Queensland coast. Marine and Freshwater 1755 Research 7, 317-342. 1756

1757 Fairbridge, R.W., 1960. The changing level of the sea. Scientific American 202, 70-79. 1758 1759 Fairbridge, R.W., 1961. Eustatic changes in sea level. Physics and Chemistry of the Earth 4, 1760

99-185. 1761 1762 Ferland, M.A., Roy, P.S., Murray-Wallace C.V., 1995. Glacial lowstand deposits on the outer 1763

continental shelf of southeastern Australia. Quaternary Research 44, 294-299. 1764 1765 Flemming, N.C., 1986. A survey of the late Quaternary landscape of the Cootamundra 1766

Shoals, north Australia: A preliminary report, In: Flemming, N.C., Marchetti, F., 1767 Stefanon, A. (Eds.), Proceedings of the 7th International Diving Science Symposium 1768 of CMAS, 1983, Padova, Italy, pp. 149–180. 1769

1770 Flood, P.G., 1983. Holocene sea level data from the southern Great Barrier Reef and 1771

southeastern Queensland – A review, In: Hopley, D. (Ed.), Australian sea levels in the 1772 last 15 000 years: A review. Monograph Series Occasional Paper 3, Geography 1773 Department, James Cook University, Townsville, pp. 85-92. 1774

1775 Flood, P.G., Frankel, E., 1989. Late Holocene higher sea level indicators from eastern 1776

Australia. Marine Geology 90, 193-195. 1777 1778 Gagan, M.K., Johnson, D.P., Crowley, G.M., 1994. Sea level control of stacked late 1779

Quaternary coastal sequences, central Great Barrier Reef. Sedimentology 41, 329-351. 1780 1781 Gehrels, W.R., 2000. Using foraminiferal transfers functions to produce high-resolution sea-1782

level records from salt-marsh deposits, Maine, USA. The Holocene 10, 367-376. 1783 1784 Gehrels, W.R., Roe, H.M., Charman, D.J., 2001. Foraminifera, testate amoebae and diatoms 1785

as sea-level indicators in UK saltmarshes: a quantitative multiproxy approach. 1786 Journal of Quaternary Science 16, 201-220. 1787

1788 Gehrels, W.R., Callard, S.L., Moss, P.T., Marshall, W.A., Blaauw, M., Hunter, J., Milton, 1789

J.A., Garnett, M.H., 2012. Nineteneth and twentieth century sea-level changes in 1790 Tasmania and New Zealand. Earth and Planetary Science Letters 315-316, 94-102. 1791

1792

Page 54: Post-glacial sea-level changes around the Australian

51

Gibb, J.G., 1986. A New Zealand regional Holocene eustatic sea-level curve and its 1793 application to determination of vertical tectonic movements: A contribution to IGCP-1794 Project 200. Royal Society of New Zealand Bulletin 24, 377-395. 1795

1796 Gill, E.D., 1983. Australian sea levels in the last 15 000 years – Victoria, SE Australia, In: 1797

Hopley, D. (Ed.), Australian sea levels in the last 15 000 years: A review. Monograph 1798 Series Occasional Paper 3, Geography Department, James Cook University, 1799 Townsville, pp. 59-63. 1800

1801 Gill, E.D., Hopley, D., 1972. Holocene sea levels in eastern Australia – A discussion. Marine 1802

Geology 12, 223-242. 1803 1804 Gillespie, R., Polach, H.A., 1979. The suitability of marine shells for radiocarbon dating of 1805

Australian prehistory. In: Proceedings, Ninth International Conference on 1806 Radiocarbon Dating, University of California Press, 404-421. 1807

1808 Goodwin, I.D., 1998. Did changes in Antarctic ice volume influence late Holocene sea-level 1809

lowering? Quaternary Science Reviews 17, 319-332. 1810 1811 Goodwin, I.D., 2003. Unravelling climatic influences on Late Holocene sea-level variability. 1812

In: Mackay, A., Battarbee, R., Birks, J., Oldfield, F. (Eds.), Global change in the 1813 Holocene, Hodder Arnold, pp. 406-421. 1814

1815 Grindrod, J., Rhodes, E.G., 1984. Holocene sea-level history of a tropical estuary: Missionary 1816

Bay, north Queensland. In: Thom, B.G. (Ed.), Coastal Geomorphology in Australia. 1817 Academic Press, Australia, pp. 151-178. 1818

1819 Grindrod, J., Moss, P., van der Kaars, S., 1999. Late Quaternary cycles of mangrove 1820

development and decline on the north Australian continental shelf. Journal of 1821 Quaternary Science 14, 465-470. 1822

1823 Hails, J.R., 1965. A critical review of sea-level changes in eastern Australia since the last 1824

glacial. Australian Geographical Studies 3, 63-78. 1825 1826 Hanebuth, T., Stattegger, K., Grootes, P.M., 2000. Rapid flooding of the Sunda Shelf: A late-1827

glacial sea-level record. Science 288, 1033-1035. 1828 1829 Hanebuth, T.J.J, Stattegger, K., Bojanowski, A., 2009. Termination of the Last Glacial 1830

Maximum sea-level lowstand: The Sunda Shelf data revisited. Global and Planetary 1831 Change 66, 76-84. 1832

1833 Harris, P.T., 1999. Sequence architecture during the Holocene transgression: an example 1834

from the Great Barrier Reef shelf, Australia- Comment. Sedimentary Geology 125, 1835 235-239. 1836

1837 Harris, P.T., Davies, P.J., 1989. Submerged reefs and terraces on the shelf edge of the Great 1838

Barrier Reef, Australia. Coral Reefs 8, 87-89. 1839 1840

Page 55: Post-glacial sea-level changes around the Australian

52

Harris, P.T., Heap, A., Marshall, J.F., Hemer, M., Daniell, J., Hancock, A., Buchanan, C., 1841 Sbaffi, L., Brewer, D., Heales, D., 2007. Submerged coral reefs and benthic habitats 1842 of the southern Gulf of Carpentaria. Geoscience Australia Record, 2007/02: 134. 1843

1844 Harris, P.T., Heap, A.D., Marshall, J.F., McCulloch, M., 2008. A new coral reef province in 1845

the Gulf of Carpentaria, Australia: Colonisation, growth and submergence during the 1846 early Holocene. Marine Geology 251, 85-97. 1847

1848 Harvey, N., Barnett, E.J., Bourman, R.P., Belperio, A.P., 1999. Holocene sea-level change at 1849

Port Pirie, South Australia: a contribution to global sea level rise estimates from tide 1850 gauges. Journal of Coastal Research 15, 607-615. 1851

1852 Haslett, S.K., 2007. The distribution of foraminifera in surface sediments of the Clyde River 1853

Estuary and Batemans Bay (New South Wales, Australia). Revista Espanola de 1854 Micropaleontologia 38, 49-75. 1855

1856 Haslett, S.K., Davies-Burrows, R., Panayotou, K., Jones, B.G., Woodroffe, C.D., 2010. 1857

Holocene evolution of the Minnamurra River estuary, southeast Australia: 1858 foraminiferal evidence. Zeitschrift für Geomorphologie 54, Suppl. 3, 79-98. 1859

1860 Haworth, R.J., Baker, R.G.V., Flood, P.G., 2002. Predicted and observed Holocene sea-levels 1861

on the Australian coast: what do they indicate about hydro-isostatic models in far-1862 field sites? Journal of Quaternary Science 17, 581-591. 1863

1864 Higley, M., 2000. Fossil oyster beds of the mid-Holocene highstand of relative sea level: 1865

GBR shelf. Honours Thesis, School of Earth Sciences, James Cook University, 1866 Townsville, Australia. 1867

1868 Hopley, D., 1971. The origin and significance of north Queensland island spits. Zeitschrift für 1869

Geomorphologie 15, 371-389. 1870 1871 Hopley, D., 1975. Contrasting evidence for Holocene sea levels with special reference to the 1872

Bowen-Whitsunday area of Queensland. In: Douglas, I., Hobbs, J.E., Pigram, J.J. 1873 (Eds.), Geographical essays in honour of Gilbert J. Butland. Department of 1874 Geography, University of New England, Armidale, pp. 51-84. 1875

1876 Hopley, D., 1978. Sea level change on the Great Barrier Reef: an introduction. Philosophical 1877

Transactions of the Royal Society London A 291, 159-166. 1878 1879 Hopley, D., 1980. Mid-Holocene high sea levels along the coastal plain of the Great Barrier 1880

Reef province: A discussion. Marine Geology 35, M1-M9. 1881 1882 Hopley, D., 1982. The geomorphology of the Great Barrier Reef: Quaternary development of 1883

coral reefs. John Wiley and Sons, Toronto, 453 pp. 1884 1885 Hopley, D., 1983a. Australian sea levels in the last 15 000 years: A review. Monograph 1886

Series Occasional Paper 3, Geography Department, James Cook University, 1887 Townsville. 1888

1889

Page 56: Post-glacial sea-level changes around the Australian

53

Hopley, D., 1983b. Evidence of 15,000 years of sea level change in tropical Queensland. In: 1890 Hopley, D. (Ed.), Australian sea levels in the last 15 000 years: A review. Monograph 1891 Series Occasional Paper 3, Geography Department, James Cook University, 1892 Townsville, pp. 93-104. 1893

1894 Hopley, D., 1984. The Holocene ‘high energy window’ on the central Great Barrier Reef. In: 1895

Thom, B.G. (Ed.), Coastal Geomorphology in Australia. Academic Press, Australia, 1896 pp. 135-150. 1897

1898 Hopley, D., 1986a. Corals and reefs as indicators of paleo-sea levels with special reference to 1899

the Great Barrier Reef. In: van de Plassche, O. (Ed.), Sea-level research: a manual for 1900 the collection and evaluation of data. Geo Books, Norwich, pp. 195-228. 1901

1902 Hopley, D., 1986b. Beachrock as a sea-level indicator. In: van de Plassche, O. (Ed.), Sea-1903

level research: a manual for the collection and evaluation of data. Geo Books, 1904 Norwich, pp. 157-173. 1905

1906 Hopley, D., 1987. Holocene sea-level changes in Australasia and the southern Pacific. In: 1907

Devoy, R.J.N. (Ed.), Sea surface studies: A global view. Croom Helm, London, pp. 1908 375-408. 1909

1910 Hopley, D., Isdale, P., 1977. Coral micro-atolls, tropical cyclones and reef flat morphology: a 1911

north Queensland example. Search 8, 79-81. 1912 1913 Hopley, D., Thom, B.G., 1983. Australian sea levels in the last 15 000 years: A review. In: 1914

Hopley, D. (Ed.), Australian sea levels in the last 15 000 years: A review. Monograph 1915 Series Occasional Paper 3, Geography Department, James Cook University, 1916 Townsville, pp. 3-26. 1917

1918 Hopley, D., Smithers, S.G., Parnell, K., 2007. Geomorphology of the Great Barrier Reef: 1919

development, diversity and change. Cambridge University Press, 546pp. 1920 1921 Hori, K., Saito, Y., 2007. An early Holocene sea-level jump and delta initiation. Geophysical 1922

Research Letters 34, doi:10.1029/2007GL031029. 1923 1924 Horton, B.P., Larcombe, P., Woodroffe, S.A., Whittaker, J.E., Wright, M.R., Wynn, C., 2003. 1925

Contemporary foraminiferal distributions of a mangrove environment, Great Barrier 1926 Reef coastline, Australia: implications for sea-level reconstructions. Marine Geology 1927 198, 225-243. 1928

1929 Horton, B.P., Culver, S.J., Hardbattle, M.I.J., Larcombe, P., Milne, G.A., Morigi, C., 1930

Whittaker, J.E., Woodroffe, S.A., 2007. Reconstructing Holocene sea-level change for 1931 the central Great Barrier Reef (Australia) using sub-tidal foraminifera. Journal of 1932 Foraminiferal Research 37, 327-343. 1933

1934 James, N.P., Bone, Y., Kyser, T.K., Dix, G.R., Collins, L.B., 2004. The importance of 1935

changing oceanography in controlling late Quaternary carbonate sedimentation on a 1936 high energy, tropical oceanic ramp: north-western Australia. Sedimentology 51, 1179-1937 1205. 1938

1939

Page 57: Post-glacial sea-level changes around the Australian

54

Jelgersma, S., Pannekoek, A.J., 1960. Post-glacial rise of sea-level in the Netherlands (a 1940 preliminary report). Geologie en Mijnbouw 39, 201-207. 1941

1942 Jennings, J.N., 1975. Desert dunes and estuarine fill in the Fitzroy estuary (north-western 1943

Australia). Catena 2, 215-262. 1944 1945 Jones, B.G., Young, R.W., Eliot, I.G., 1979. Stratigraphy and chronology of receding barrier 1946

beach deposits on the northern Illawarra coast of New South Wales. Journal of the 1947 Geological Society of Australia 26, 255-264. 1948

1949 Jones, B.G., Woodroffe, C.D., Martin, G.R., 2003. Deltas in the Gulf of Carpentaria, 1950

Australia: forms, processes and products. In: Tropical Deltas of Southeast Asia – 1951 sedimentology, stratigraphy and petroleum geology, SEPM (Society for Sedimentary 1952 Geology) Special Publication 76, 21-43. 1953

1954 Kelletat, D., 2006. Beachrock as a sea-level indicator? Remarks from a geomorphological 1955

point of view. Journal of Coastal Research 22, 1558-1564. 1956 1957 Knight, J., 1997. Beachrock reconsidered. Discussion of Kelletat, D., 2006. Beachrock as a 1958

sea-level indicator? Remarks from a geomorphological point of view. Journal of 1959 Coastal Research 22, 1558-1564. Journal of Coastal Research 23, 1074-1078. 1960

1961 Laborel, J., Laborel-Deguen, F., 1996. Biological indicators of Holocene sea-level and 1962

climatic variations on rocky coasts of tropical and subtropical regions. Quaternary 1963 International 31, 53-60. 1964

1965 Lambeck, K., 2002. Sea level change from Mid Holocene to Recent time: An Australian 1966

example with global implications. In: Mitrovica, J.X., Vermeersen, B. (Eds.), Ice 1967 Sheets, Sea Level and the Dynamic Earth. Volume 29, AGU, 33-50. 1968

1969 Lambeck, K., Chappell, J., 2001. Sea level change through the last glacial cycle. Science 292, 1970

679-686. 1971 1972 Lambeck, K., Nakada, M., 1990. Late Pleistocene and Holocene sea-level change along the 1973

Australian coast. Palaeogeography, Palaeoclimatology, Palaeoecology 89, 143-176. 1974 1975 Lambeck, K., Woodroffe, C.D., Antonioli, F., Anzidei, M., Gehrels, W.R., Laborel, J., 1976

Wright, A.J., 2010. Paleoenvironmental records, geophysical modelling, and 1977 reconstruction of sea-level trends and variability on centennial and longer timescales. 1978 In: Church, J.A., Woodworth, P.L., Aarup, T., Wilson, W.S. (Eds.), Understanding 1979 sea-level rise and variability. Blackwell Publishing Ltd. pp. 61-121. 1980

1981 Langford-Smith, T., Thom, B.G., 1969. New South Wales coastal morphology. Journal of the 1982

Geological Society of Australia 16, 572-580. 1983 1984 Larcombe, P., Carter, R.M., 1998. Sequence architecture during the Holocene transgression: 1985

an example from the Great Barrier Reef shelf, Australia. Sedimentary Geology 117, 1986 97-121. 1987

1988 Larcombe, P., Carter, R.M., Dye, J., Gagan, M.K., Johnson, D.P., 1995. New evidence for 1989

Page 58: Post-glacial sea-level changes around the Australian

55

episodic post-glacial sea-level rise, central Great Barrier Reef, Australia. Marine 1990 Geology 127, 1-44. 1991

1992 Lewis, S.E., 2005. Environmental trends in the GBR lagoon and Burdekin River catchment 1993

during the mid-Holocene and since European settlement using Porites coral records, 1994 Magnetic Island, Queensland. Ph.D. Thesis, School of Earth Sciences, James Cook 1995 University, Townsville, Australia. 1996

1997 Lewis, S.E., Wüst, R.A.J., Webster, J.M., Shields, G.A., 2008. Mid-late Holocene sea-level 1998

variability in eastern Australia. Terra Nova 20, 74-81. 1999 2000 Lewis, S.E., Wüst, R.A.J., Webster, J.M., Shields, G.A., Renema, W., Lough, J.M., Jacobsen, 2001

G., 2012. Development of an inshore fringing coral reef using textural, compositional 2002 and stratigraphic data from Magnetic Island, Great Barrier Reef, Australia. Marine 2003 Geology 299-302, 18-32. 2004

2005 Logan, B.W., Read, J.F., Davies, G.R., 1970. History of carbonate sedimentation, Quaternary 2006

Epoch, Shark Bay, Western Australia. In: Logan, B.W., Davies, G.R., Read, J.F., 2007 Cebulski, D.E. (Eds.), Carbonate sedimentation and environments, Shark Bay, 2008 Western Australia. American Association of Petroleum Geologists Memoir 13, pp. 2009 38-84. 2010

2011 Macphail, M., 1974. Pollen analysis of a buried organic deposit on the backshore at Fingal 2012

Bay, Port Stephens, New South Wales. Proceedings of the Linnean Society of New 2013 South Wales 98, 222-233. 2014

2015 Matthews, R.K., 1990. Quaternary sea-level change. In: Revelle, R.R. (Ed.), Sea-level 2016

change. Committee on Global Change, National Research Council, Studies in 2017 Geophysics, National Academy Press, Washington, DC., pp. 88-103. 2018

2019 Martin, A.R.H., 1972. The depostitional environment of organic deposits on the foreshore at 2020

north DeeWhy, NSW. Proceedings of the Linnean Society of New South Wales 96, 2021 278-281. 2022

2023 McGregor, H.V., Gagan, M.K., McCulloch, M.T., Hodge, E., Mortimer, G., 2008. Mid-2024

Holocene variability in the marine 14C reservoir age for northern coastal Papua New 2025 Guinea. Quaternary Geochronology 3, 213-225. 2026

2027 McLean, R., 2011. Beachrock. In: Hopley, D. (Ed.), Encyclopaedia of coral reefs. pp 107-2028

111. 2029 2030 McLean, R.F., Stoddart, D.R., Hopley, D., Polach, H., 1978. Sea level change in the 2031

Holocene on the northern Great Barrier Reef. Philosophical Transactions of the Royal 2032 Society London A 291, 167-186. 2033

2034 Murray-Wallace, C.V., 2002. Pleistocene coastal stratigraphy, sea level highstands and 2035

neotectonism of the southern Australian passive margin – a review. Journal of 2036 Quaternary Science 17, 469-489. 2037

2038

Page 59: Post-glacial sea-level changes around the Australian

56

Murray-Wallace, C.V., Belperio, A.P., 1991. The last interglacial shoreline in Australia – A 2039 review. Quaternary Science Reviews 10, 441-461. 2040

2041 Murray-Wallace, C.V., Belperio, A.P., 1994. Identification of remanié fossils using amino 2042

acid racemisation. Alcheringa 18, 219-227. 2043 2044 Murray-Wallace, C.V., Goede, A., 1995. Aminostratigraphy and electron spin resonance 2045

dating of Quaternary coastal neotectonism in Tasmania and the Bass Strait Islands. 2046 Australian Journal of Earth Sciences 42, 51-67. 2047

2048 Murray-Wallace, C.V., Ferland, M.A., Roy, P.S., Sollar, A., 1996. Unravelling patterns of 2049

reworking in lowstand shelf deposits using amino acid racemisation and radiocarbon 2050 dating. Quaternary Science Reviews 15, 685-697. 2051

2052 Murray-Wallace, C.V., Ferland, M.A., Roy, P.S., 2005. Further amino acid racemisation 2053

evidence of glacial age, multiple lowstand deposition on the New South Wales outer 2054 continental shelf, southeastern Australia. Marine Geology 214, 235-250. 2055

2056 Nakada, M., Lambeck, K., 1989. Late Pleistocene and Holocene sea-level change in the 2057

Australia region and mantle rheology. Geophysical Journal 96, 497-517. 2058 2059 Nakiboglu, S.M., Lambeck, K., Aharon, P., 1983. Postglacial sealevels in the Pacific: 2060

Implications with respect to deglaciation regime and local tectonics. Tectonophysics 2061 91, 335-358. 2062

2063 Neumann, A.C., Macintyre, I., 1985. Reef response to sea level rise: keep-up, catch-up or 2064

give-up. Proceedings of the 5th International Coral Reef Congress 3, 105-110. 2065 2066 Nott, J., 1996. Late Pleistocene and Holocene sea-level highstands in northern Australia. 2067

Journal of Coastal Research 12, 907-910 2068 2069 Ota, Y., Chappell, J., 1999. Holocene sea-level rise and coral reef growth on a tectonically 2070

rising coast, Huon Peninsula, Papua New Guinea. Quaternary International 55, 51-59. 2071 2072 Ota, Y., Chappell, J., Kelley, R., Yonekura, N., Matsumoto, E., Nishimura, T., Head, J., 2073

1993. Holocene coral reef terraces and coseismic uplift of Huon Peninsula, Papua 2074 New Guinea. Quaternary Research 40, 177-188. 2075

2076 Pandolfi, J.M., 1996. Limited membership in Pleistocene reef coral assemblages from Huon 2077

Peninsula, Papua New Guinea: constancy during global change. Paleobiology 22, 2078 152-176. 2079

2080 Pandolfi, J.M., Tudhope, A.W., Burr, G., Chappell, J., Edinger, J., Frey, M., Steneck, R., 2081

Sharma, C., Yeates, A., Jennions, M., Lescinsky, H., Newton, A., 2006. Mass 2082 morality following disturbance in Holocene coral reefs in Papua New Guinea. 2083 Geology 34, 949-952. 2084

2085 Peltier, W.R., 2002. On eustatic sea level history: Last Glacial Maximim to Holocene. 2086

Quaternary Science Reviews 21, 377-396. 2087 2088

Page 60: Post-glacial sea-level changes around the Australian

57

Peltier, W.R., Fairbanks, R.G., 2006. Global glacial ice volume and Last Glacial Maximum 2089 duration from an extended Barbados sea level record. Quaternary Science Reviews 2090 25, 3322-3337. 2091

2092 Perry, C.T., Smithers, S.G., 2011. Cycles of coral reef ‘turn-on’, rapid growth and ‘turn-off’ 2093

over the past 8500 years: a context for understanding modern ecological states and 2094 trajectories (refer to supplementary material). Global Change Biology 17, 76-86. 2095

2096 Phipps, C.V.G., 1970. Dating of eustatic events from cores taken in the Gulf of Carpentaria 2097

and samples from the New South Wales continental shelf. Australian Journal of 2098 Science 32, 328-330. 2099

2100 Pirazzoli, P.A., 1991. World atlas of Holocene sea-level changes. Elsevier, Amsterdam, The 2101

Netherlands, 300 pp. 2102 2103 Pirazzoli, P.A., 1996. Sea-level changes: The last 20,000 years. John Wiley & Sons, West 2104

Sussex, England, 211 pp. 2105 2106 Playford, P.E., 1988. Guidebook to the geology of Rottnest Island. Geological Society of 2107

Australia & Geological Survey of Western Australia, Perth, Australia. 2108 2109 Reeves, J.M., Chivas, A.R., Garcia, A., Holt, S., Couapel, M.J.J., Jones, B., Cendon, D.I., 2110

Fink, D., 2008. The sedimentary record of palaeoenvironments and sea-level change 2111 in the Gulf of Carpentaria, Australia, through the last glacial cycle. Quaternary 2112 International 183, 3-22. 2113

2114 Reimer, P.J., Baillie, M.G.L., Bard, E., Bayliss, A., et al., 2009. INTCAL09 and Marine09 2115

radiocarbon age calibration curves 0-50,000 years cal BP. Radiocarbon 51, 1111-2116 1150. 2117

2118 Rhodes, E.G., 1982. Depositional model for a chenier plain, Gulf of Carpentaria, Australia. 2119

Sedimentology 29, 201-221. 2120 2121 Rhodes, E.G., Polach, H.A., Thom, B.G., Wilson, S.R., 1980. Age structure of Holocene 2122

coastal sediments: Gulf of Carpentaria, Australia. Radiocarbon 22, 718-727. 2123 2124 Rosen, B.R., 1978. The nature and significance of microatolls: Appendix: Determination of a 2125

collection of coral microatoll specimens from the northern Great Barrier Reef. 2126 Philosophical Transactions of the Royal Society London B 284, 115-122. 2127

2128 Roy, P.S., Crawford, E.A., 1981. Holocene geological evolution of the southern Botany Bay-2129

Kurnell region, central New South Wales coast. Records of the Geological Survey of 2130 New South Wales, Department of Mineral Resources 20, 159-250. 2131

2132 Roy, P.S., Thom, B.G., 1981. Late Quaternary marine deposition in New South Wales and 2133

southern Queensland - an evolutionary model. Journal of the Geological Society of 2134 Australia 28, 417-489. 2135

2136 Roy, P.S., Thom, B.G., Wright, L.D., 1980. Holocene sequences on an embayed high-energy 2137

coast: an evolutionary model. Sedimentary Geology 26, 1-19. 2138

Page 61: Post-glacial sea-level changes around the Australian

58

2139 Scoffin, T.P., 1977. Sea-level features of reefs in the northern province of the Great Barrier 2140

Reef. Proceedings of the Third International Coral Reef Symposium, USA, pp. 319-2141 324. 2142

2143 Scoffin, T.P., Stoddart, D.R., 1978. The nature and significance of microatolls. Philosophical 2144

Transactions of the Royal Society London B 284, 99-122. 2145 2146 Searle, D.J., Woods, P.J., 1986. Detailed documentation of a Holocene sea-level record in the 2147

Perth region, southern Western Australia. Quaternary Research 26, 299-308. 2148 2149 Searle, D.J., Semeniuk, V., Woods, P.J., 1988. The geomorphology, stratigraphy and 2150

Holocene history of the Rockingham-Becher plain. Journal of the Royal Society of 2151 Western Australia 70, 89-109. 2152

2153 Semeniuk, V., 1980a. Quaternary stratigraphy of the tidal flats, King Sound, Western 2154

Australia. Journal of the Royal Society of Western Australia 63, 65-78. 2155 2156 Semeniuk, V., 1980b. Geomorphology and Holocene history of the tidal flats, King Sound, 2157

north-western Australia. Journal of the Royal Society of Western Australia 65, 47-68. 2158 2159 Semeniuk, V., 1981. Sedimentology and the stratigraphic sequence of a tropical tidal flat, 2160

north-western Australia. Sedimentary Geology 29, 195-221. 2161 2162 Semeniuk, V., 1982. Geomorphology and Holocene history of the tidal flats, King Sound, 2163

north-western Australia. Journal of the Royal Society of Western Australia 65, 47-68. 2164 2165 Semeniuk, V., 1985. The age structure of a Holocene barrier dune system and its implications 2166

for sea level history reconstructions in southwestern Australia. Marine Geology 67, 2167 197-212. 2168

2169 Semeniuk, V., 1996. An early Holocene record of rising sea level along a bathymetrically 2170

complex coast in southwestern Australia. Marine Geology 131, 177-193. 2171 2172 Semeniuk, V., Searle, D.J., 1986. Variability of Holocene sea level history along the 2173

southwestern coast of Australia – Evidence for the effect of significant local 2174 tectonism. Marine Geology 72, 47-58. 2175

2176 Shepard, F.P., 1961. Sea level rise during the past 20,000 years. Zeitschrift für 2177

Geomorphologie Supplement 3, 30-35. 2178 2179 Short, A.D., Fotheringham, D.G., Buckley, R.C., 1986. Coastal morphodynamics and 2180

Holocene evolution of the Eyre Peninsula coast, South Australia. University of 2181 Sydney, Coastal Studies Unit Technical Report 86/2. 2182

2183 Sloss, C.R., Murray-Wallace, C.V., Jones, B.G., Wallin T., 2004. Aspartic acid racemisation 2184

dating of mid Holocene to recent estuarine sedimentation in New South Wales, 2185 Australia: a pilot study. Marine Geology 212, 45-59. 2186

2187

Page 62: Post-glacial sea-level changes around the Australian

59

Sloss, C.R., Jones, B.G., Murray-Wallace, C.V., McClennen, C.E., 2005. Holocene sea level 2188 fluctuations and the sedimentary evolution of a barrier estuary: Lake Illawarra, New 2189 South Wales, Australia. Journal of Coastal Research 21, 943-959. 2190

2191 Sloss, C.R., Jones, B.G., McClennen, C.E., de Carli, J., Price, D.M., 2006. The 2192

geomorphological evolution of a wave-dominated barrier estuary: Burrill Lake, New 2193 South Wales, Australia. Sedimentary Geology 187, 229-249. 2194

2195 Sloss, C.R., Murray-Wallace, C.V., Jones, B.G., 2007. Holocene sea-level change on the 2196

southeast coast of Australia: a review. The Holocene 17, 999-1014. 2197 2198 Sloss, C.R., Jones, B.G., Brooke, B.P., Heijnis, H., Murray-Wallace, C.V., 2011. Contrasting 2199

sedimentation rates in Lake Illawarra and St Georges Basin, two large barrier 2200 estuaries on the southeast coast of Australia. Journal of Paleolimnology 46, 561-577. 2201

2202 Sloss, C. R., Westerway, K., Hau, Q., Murray-Wallace, C.V., in press. Chapter: 14.36 An 2203

introduction to dating techniques: A guide for Geomorphologists. In: Treatise on 2204 Geomorphology, Vol. 14: Methods in Geomorphology, Elsevier. Accepted May, 2011. 2205

2206 Smith, A. 1978. Case study: Magnetic Island and its fringing reefs. In: Hopley, D. (Ed.), 2207

Geographical Studies of the Townsville area. Department of Geography monograph 2208 series no. 2, James Cook University, Townsville, pp. 59-65. 2209

2210 Smithers, S., 2011. Sea-level indicators. In: Hopley, D. (Ed.), Encyclopaedia of coral reefs. 2211

pp. 978-991. 2212 2213 Smithers, S.G., Woodroffe, C.D., 2000. Microatolls as sea-level indicators on a mid-ocean 2214

atoll. Marine Geology 168, 61-78. 2215 2216 Smithers, S.G., Woodroffe, C.D., 2001. Coral microatolls and 20th century sea level in the 2217

eastern Indian Ocean. Earth and Planetary Science Letters 191, 173-184. 2218 2219 Smithers, S.G., Hopley, D., Parnell, K.E., 2006. Fringing and Nearshore Coral Reefs of the 2220

Great Barrier Reef: Episodic Holocene Development and Future Prospects. Journal of 2221 Coastal Research 22, 175-187. 2222

2223 Southall, K.E., Gehrels, W.R., Hayward, B.W., 2006. Foraminifera in a New Zealand salt 2224

marsh and their suitability as sea-level indicators. Marine Micropaleontology 60, 167-2225 179. 2226

2227 Stoddart, D.R., Scoffin, T.P., 1979. Micro-atolls: review of form, origin and terminology. 2228

Atoll Research Bulletin 224, 1-17. 2229 2230 Stuiver, M., P.J. Reimer, R.W. Reimer, R.W., 2005. CALIB 5.0. www.calib.org 2231 2232 Teichert, C., 1950. Late Quaternary sea-level changes at Rottnest Island, Western Australia. 2233

Proceedings of the Royal Society of Victoria 59, 63-79. 2234 2235 Thom, B.G., 1984a. Coastal Geomorphology in Australia. Academic Press, Sydney, 349 pp. 2236 2237

Page 63: Post-glacial sea-level changes around the Australian

60

Thom, B.G., 1984b. Sand barriers of eastern Australia: Gippsland – A case study. In: Thom, 2238 B.G. (Ed.), Coastal geomorphology in Australia. Academic Press, Sydney, pp. 233-2239 261. 2240

2241 Thom, B.G., Chappell, J., 1975. Holocene sea levels relative to Australia. Search 6, 90-93. 2242 2243 Thom, B.G., Chappell, J., 1978. Holocene sea level change: an interpretation: Philosophical 2244

Transactions of the Royal Society London A 291, 187-194. 2245 2246 Thom, B.G., Roy, P., 1983. Sea level change in New South Wales over the past 15000 years, 2247

In: Hopley, D. (Ed.), Australian sea levels in the last 15 000 years: A review. 2248 Monograph Series Occasional Paper 3, Geography Department, James Cook 2249 University, Townsville, pp. 64-84. 2250

2251 Thom, B.G., Roy, P., 1985. Relative sea levels and coastal sedimentation in southeast 2252

Australia in the Holocene. Journal of Sedimentary Petrology 55, 257-264. 2253 2254 Thom, B.G., Hails, J.R., Martin, R.H., 1969. Radiocarbon evidence against higher postglacial 2255

sea levels in eastern Australia. Marine Geology 7, 161-168. 2256 2257 Thom, B.G., Hails, J.R., Martin, R.H., Phipps, C.V.G., 1972. Postglacial sea levels in eastern 2258

Australia – A reply. Marine Geology 12, 223-242. 2259 2260 Thom, B.G., Wright, L.D., Coleman, J.M., 1975. Mangrove ecology and deltaic-estuarine 2261

geomorphology Cambridge Gulf – Ord River, Western Australia. Journal of Ecology 2262 63, 203-232. 2263

2264 Thom, B.G., Bowman, G.M., Roy, P.S., 1981. Late Quaternary evolution of coastal sand 2265

barriers, Port Stephens-Myall Lakes area, Central News South Wales, Australia. 2266 Quaternary Research 15, 345-364. 2267

2268 Thom, B.G., Shepherd, M., Ly, C.K., Roy, P.S., Bowman, G.M., Hesp, P.A., 1992. Coastal 2269

geomorphology and Quaternary geology of the Port Stephens-Myall Lakes area. 2270 Department of Biogeography and Geomorphology, Australian National University 2271 Monograph 6, Australian National University, Canberra, 407 pp. 2272

2273 Ulm, S., 2002. Marine and estuarine reservoir effects in central Queensland, Australia: 2274

Determination of ΔR values. Geoarchaeology: An International Journal 17, 319-348. 2275 2276 van Andel, T.H., Veevers, J.J., 1967. Morphology and sediments of the Timor Sea. Bureau of 2277

Mineral Resources Geology and Geophysics Bulletin 83, 173 pp. 2278 2279 van de Plassche, O. (Ed.), 1986. Sea-level Research: A Manual for the Collection and 2280

Evaluation of Data. Geo Books, Norwich, UK, 618 pp. 2281 2282 Veeh, H.H., Veevers, J.J., 1970. Sea level at -175 off the Great Barrier Reef 13,600 to 17,000 2283

year ago. Nature 226, 536-537. 2284 2285 Vousdoukas, M.I., Velegrakis, A.F., Plotmaritis, T.A., 2007. Beachrock occurrence, 2286

characteristics, formation mechanisms and impacts. Earth-Science Reviews 85, 23-46. 2287

Page 64: Post-glacial sea-level changes around the Australian

61

2288 Walcott, R.I., 1972. Past sea levels, eustasy and deformation of the earth. Quaternary 2289

Research 2, 1-14. 2290 2291 Wang, P., Chappell, J., 2001. Foraminifera as Holocene environmental indicators in the 2292

South Alligator River, Northern Australia. Quaternary International 83-85, 47-62. 2293 2294 Webster, J. M., Yokoyama, Y., Cotterill, C., and Expedition 325 Scientists 2011. Proceedings 2295

of the Integrated Ocean Drilling Program 325, doi:10.2204/iodp.proc.325.2011. 2296 2297 Weninger, B., Jöris, O., Danzeglocke, U., 2011. CalPal-2007. Cologne Radiocarbon 2298

Calibration & Palaeoclimate Research Package. http://www.calpal.de/, accessed 2011-2299 04-29. 2300

2301 Woodroffe, C.D., 1988. Mangroves and sedimentation in reef environments: indicators of 2302

past sea-level changes, and present sea-level trends? Proceedings of the 6th 2303 International Coral Reef Symposium, Australia 3, pp. 535-539. 2304

2305 Woodroffe, C.D., 2003. Coasts: Form process and evolution. Cambridge University Press, 2306

Cambridge, 623 pp. 2307 2308 Woodroffe, C.D., Chappell, J., 1993. Holocene emergence and evolution of the McArthur 2309

River Delta, southwestern Gulf of Carpentaria, Australia. Sedimentary Geology 83, 2310 303-317. 2311

2312 Woodroffe, C., McLean, R., 1990. Microatolls and recent sea level change on coral atolls. 2313

Nature 344, 531-534. 2314 2315 Woodroffe, C.D., Thom, B.G., Chappell, J., 1985. Development of widespread mangrove 2316

swamps in mid-Holocene times in northern Australia. Nature 317, 711-713. 2317 2318 Woodroffe, C.D., Thom, B.G., Chappell, J., Wallensky, E., Grindrod, J., Head, J., 1987. 2319

Relative sea level in the South Alligator River region, north Australia, during the 2320 Holocene. Search 18, 198-200. 2321

2322 Woodroffe, C.D., Chappell, J., Thom, B.G., Wallensky, E., 1989. Depositional model of a 2323

macrotidal estuary and floodplain, South Alligator River, N.T. Australia. 2324 Sedimentology 36, 737-756. 2325

2326 Woodroffe, C.D., Murray-Wallace, C.V., Bryant, E.A., Brooke, B., Heijnis, H., Price, D.M., 2327

1995. Late Quaternary sea-level highstands in the Tasman Sea: evidence from Lord 2328 Howe Island. Marine Geology 125, 61-72. 2329

2330 Woodroffe, C.D., Kennedy, D.M., Hopley, D., Rasmussen, C.E., Smithers, S.G., 2000. 2331

Holocene reef growth in Torres Strait. Marine Geology 170, 331-346. 2332 2333 Woodroffe, C.D., Brooke, B.P., Linklater, M., Kennedy, D.M., Jones, B.G., Buchanan, C., 2334

Mleczko, R., Hua, Q., Zhao, J., 2010. Response of coral reefs to climate change: 2335 Expansion and demise of the southernmost Pacific coral reef. Geophysical Research 2336 Letters 37, L15602, doi:10.1029/2010GL044067. 2337

Page 65: Post-glacial sea-level changes around the Australian

62

2338 Woodroffe, C.D., McGregor, H.V., Lambeck, K., Smithers, S.G., Fink, D., 2012. Mid-Pacific 2339

microatolls record sea-level stability over the past 5000 years. Geology. 2340 doi:10.1130/G33344.1 2341

2342 Woodroffe, S.A., 2009. Testing models of mid to late Holocene sea-level change, north 2343

Queensland, Australia. Quaternary Science Reviews 28, 2474-2488. 2344 2345 Woodroffe, S.A., Horton, B.P., 2005. Holocene sea-level changes in the Indo-Pacific. Journal 2346

of Asian Earth Sciences 25, 29-43. 2347 2348 Woodroffe, S.A., Horton, B.P., Larcombe, P., Whittaker, J.E., 2005. Intertidal mangrove 2349

foraminifera from the central Great Barrier Reef shelf, Australia, implications for sea-2350 level reconstruction. Journal of Foraminiferal Research 35, 259-270. 2351

2352 Wright, L.D., Coleman, J.M., Thom, B.G., 1972. Emerged tidal flats in the Ord River estuary, 2353

Western Australia. Search 3, 339-341. 2354 2355 Wright, S.A., 2011. Fixed intertidal biological indicators and Holocene sea-level on the Great 2356

Barrier Reef coast. Ph.D. Thesis, School of Behavioural, Cognitive and Social 2357 Sciences, University of New England, Armidale, Australia. 2358

2359 Wyrwoll, K-H., Zhu, Z., Kendrick, G., Collins, L., Eisenhauer, A., 1995. Holocene sea-level 2360

events in Western Australia: Revisiting old questions. Journal of Coastal Research 2361 Special Issue 17, 321-326. 2362

2363 Yokoyama, Y., Lambeck, K., De Deckker, P., Johnson, P., Fifield, L.K., 2000. Timing of the 2364

last glacial maximum from observed sea-level minima. Nature 406, 713-716. 2365 2366 Yokoyama, Y., De Deckker, P., Lambeck, K., Johnson, P., Fifield, L.K., 2001a.Sea-level at 2367

the last glacial maximum: evidence from northwestern Australia to constrain ice 2368 volumes for oxygen isotope stage 2. Palaeogeography, Palaeoclimatology, 2369 Palaeoecology 165, 281-297. 2370

2371 Yokoyama, Y., Purcell, A., Lambeck, K., Johnston, P., 2001b. Shore-line reconstruction 2372

around Australia during the Last Glacial Maximum and Late Glacial Stage. 2373 Quaternary International 83-85, 9-18. 2374

2375 Yokoyama, Y., Purcell, A., Marshall, J.F., Lambeck, K., 2006. Sea-level during the early 2376

deglaciation period in the Great Barrier Reef, Australia. Global and Planetary Change 2377 53, 147-153. 2378

2379 Yokoyama, Y., Webster, J. M., Cotterill, C., Braga, J. C., Jovane, L., Mills, H., Morgan, S., 2380

Suzuki, A., and Expedition 325 Scientists, 2011. IODP Expedition 325: Great Barrier 2381 Reefs reveals past sea-level, climate and environmental changes during the end of the last 2382 Ice age. Scientific Drilling 12, 32-45, doi:10.2204/iodp.sd.12.04.2011. 2383

2384 Young, R.W., Bryant, E.A., Price, D.M., Wirth, L.M., Pease, M., 1993. Theoretical 2385

constraints and chronological evidence of Holocene coastal development in central 2386 and southern New South Wales, Australia. Geomorphology 7, 317-329. 2387

Page 66: Post-glacial sea-level changes around the Australian

63

2388 Yu, K.F., Zhao, J.X., 2010. U-series dates of Great Barrier Reef corals suggest at least +0.7 2389

m sea level ~7000 years ago. The Holocene 20, 161-168. 2390 2391 Zwartz, D.P., 1995. The recent history of the Antarctic Ice Sheet: constraints from sea-level 2392

change. Ph.D. Thesis, Research School of Earth Sciences, Australian National 2393 University, Canberra, Australia. 2394

2395 2396

2397

2398

2399

2400

2401

2402

2403

2404

2405

2406

2407

2408

2409

2410

2411

2412

2413

2414

2415

2416

2417

2418

2419

2420

2421

2422

2423

Page 67: Post-glacial sea-level changes around the Australian

64

Figure captions 2424

2425 Figure 1. Map of key sea-level sites around the Australasian region. 2426 2427 Figure 2. Summary of data showing the post-glacial sea-level rise for the Australasian region. 2428 The envelope is drawn to capture intertidal indicators and the zone between the terrestrial and 2429 marine directional indicators. Sites include New Zealand (NZ: e.g. Gibb, 1986), north-west 2430 shelf (NW shelf: e.g. Yokoyama et al., 2000, 2001a; James et al., 2004), Huon Peninsula 2431 (Huon: e.g. Chappell and Polach, 1991; Edwards et al., 1993; Ota et al., 1993; Chappell et al., 2432 1996), Queensland (QLD: e.g. Larcombe et al., 1995), Sunda Shelf (Sunda: Hanebuth et al., 2433 2000, 2009), Western Australia (WA: e.g. Einsenhauer et al., 1993; Semeniuk, 1985, 1996), 2434 Northern Territory (NT: Woodroffe et al., 1987), South Australia (SA: Belperio et al., 2002) 2435 and New South Wales (NSW: Sloss et al., 2007). The following vertical errors have been 2436 assigned to the data: ±3 m for the intertidal (Inter.) indicators, +10, -1 for the marine 2437 indicators and +1, -10 for the terrestrial (Terr.) indicators. Note that meltwater pulse 1A (1A) 2438 is well-represented in the Sunda Shelf dataset. 2439 2440 Figure 3. Data from mangrove material from the South Alligator River, Northern Territory 2441 (calibrated 14C ages from Woodroffe et al., 1987). 2442 2443 Figure 4. Summary of sea-level data for the Queensland region (a). Indicators include 2444 barnacles (Beaman et al., 1994; Higley, 2000), beachrock (Hopley, 1980), foraminiferal 2445 transfer function (Woodroffe, 2009), mangroves (Larcombe et al., 1995), coral microatolls 2446 (Chappell et al., 1983) and oyster beds (Beaman et al., 1994; Higley, 2000; Lewis et al., 2447 2008). Note the clear offset between the microatolls, barnacles and oysters compared with the 2448 beachrock and foraminifera data. The data fit within a tighter envelope when only the most 2449 reliable indicators are considered where the elevations can be directly measured to the 2450 modern counterparts (b; barnacles, microatolls and oyster beds). 2451 2452 Figure 5. Summary of key sea-level data from New South Wales (compiled in Sloss et al., 2453 2007). 2454 2455 Figure 6. A selection of data from Spencer Gulf, South Australia to highlight the difference in 2456 sea-level magnitudes between Port Lincoln (PL), Redcliff (Red) and Port Pirie (PP) that 2457 varies with distance from the continental shelf (calibrated 14C ages from Belperio et al., 2458 2002). 2459 2460 Figure 7. Summary of the key sea-level data from Western Australia including barnacles, 2461 tubeworms (Baker et al., 2005), swash zone deposits (Searle and Woods, 1986; Searle et al., 2462 1988) and coral pavements (Collins et al., 2006). 2463 2464 2465 2466 2467 2468 2469 2470 2471 2472 2473

Page 68: Post-glacial sea-level changes around the Australian

65

2474 Figure 1. 2475 2476

Page 69: Post-glacial sea-level changes around the Australian

66

2477 Figure 2. 2478 2479

Page 70: Post-glacial sea-level changes around the Australian

67

2480 Figure 3. 2481 2482

Page 71: Post-glacial sea-level changes around the Australian

68

2483 Figure 4. 2484 2485

Page 72: Post-glacial sea-level changes around the Australian

69

2486 Figure 5. 2487 2488

Page 73: Post-glacial sea-level changes around the Australian

70

2489 Figure 6. 2490 2491

Page 74: Post-glacial sea-level changes around the Australian

71

2492 Figure 7. 2493