Academia.eduAcademia.edu
A LITTLE FROM COLUMN A, A LITTLE FROM COLUMN B: MERGING HIGH-RESOLUTION RESOURCE MODELLING WITH ARCHAEOLOGICAL DATA TO CONSTRUCT AN INDIGENOUS OCCUPATION HISTORY FOR INLAND SOUTHWESTERN AUSTRALIA. Alana M. Rossi BA (Hons), MA. Submitted in fulfillment of the requirements for the Doctor of Philosophy School of Arts and Sciences Fremantle November, 2020 Declaration To the best of the candidate’s knowledge, this thesis contains no material previously published by another person, except where due acknowledgement has been made. This thesis is the candidate’s own work and contains no material which has been accepted for the award of any other degree or diploma in any institution. Signature: Name: Alana M. Rossi Date: 1 November 2020 ii ABSTRACT In Noongar country, southwestern Australia, most archaeological research has focussed on the coastal margin. Only one inland site has been excavated and dated – Mulka's Cave – but interpretations were hampered by the lack of comparative data (Rossi 2010, 2014a). This research aims to address the knowledge gap for the inland southwest by creating an occupation history for the area within 50 km of Mulka's Cave, by merging high-resolution resource modelling with archaeological data. The study area was divided into Landscape Divisions based on landform and vegetation characteristics; monthly availability of food and water were quantified for each, using published sources, agricultural tools (Department of Primary Industries and Regional Development 2016a) and a specially devised model. Optimal Foraging Theory was then used to define the optimal seasonal route for the study area occupants. Archaeological investigations focussed on three sites: Anderson Rocks, Gibb Rock and Mulka's Cave. Stone artefact attributes were interpreted using Kuhn's (1995) provisioning systems, whereby different levels of mobility favour contrasting technological strategies. The archaeological data and occupation model identified a system of predictable and unpredictable residential moves; predictability and duration were dictated by the availability of water on-site and elsewhere in the landscape. Occupation was longer and more predictable in the northern and central study area, where groups were tethered to more reliable water sources during summer and autumn. More frequent, unpredictable moves occurred in the southern study area, as winter rains allowed people to disperse throughout the landscape. The archaeological data demonstrate that the study area was visited as early as 9650 cal. BP, but regular occupation occurred from 7775 cal. BP. Considering the reliance on ephemeral water sources, it is unlikely that the inland southwest could have sustained regular occupation during the Pleistocene, when conditions were drier. This study highlights the value in combining theoretical occupation models with archaeological data to maximise interpretive value. iii ACKNOWLEDGEMENTS First and foremost, I would like to thank my primary supervisor Dr Shane Burke for always being approachable, encouraging, and for addressing my [very] lengthy emails and chapter drafts in a timely fashion. I am also extremely grateful for the financial assistance, which helped cover unanticipated field costs. I would also like to thank my secondary supervisor, R. Esmee Webb for her assistance, particularly during the fieldwork at Mulka's Cave during my MA research. Her support and advice were especially appreciated when meeting with Indigenous land councils and navigating the application processes associated with the Aboriginal Heritage Act (1972). The following individuals and institutions are thanked for their support or assistance in particular areas: o The Australian Government, for providing fees offset under the Research Training Program (RTP) scheme; o Graduate Women Western Australia, whose $5000 scholarship for radiocarbon dating was vital to this research; o Fisher Rare and Special Collections, University of Sydney, for providing a high-resolution scan of Curr's (1887) map; o South-West Aboriginal Land and Sea Council (SWALSC), particularly the Ballardong Traditional Owners who accompanied me to Twine Reserve: Clarence Barron, Donna Hill and Marques Ugle. The Ballardong's lawyer, Peter Nettleton, is also thanked for facilitating meetings and providing assistance and advice as required; o The Research Office, University of Notre Dame, Fremantle, for financial and administrative assistance; o Staff from the School of Arts and Sciences and Student Administration and Fees, University of Notre Dame, Fremantle for ongoing administrative assistance and other support; o Professor Chris Clarkson, Professor Peter Hiscock and Professor Simon Holdaway for examining this thesis, and providing valuable feedback. iv The penultimate thanks must go to my family and friends, for their support throughout this incredibly long journey. I am particularly grateful for their efforts in suppressing the "WHAT ARE YOU GOING TO DO AFTERWARDS?!?" question that everyone wants to ask/yell but represses for fear it might initiate some sort of breakdown. My immediate family were especially supportive during the last 6 weeks prior to draft submission, when I felt like it was all becoming too much. Simple things like a group chat with Simpsons references, terrible jokes, photos or updates from home always came at a time that I really needed them and reminded me that there was life outside the PhD bubble. Finally, I want to thank my future wife Jess Godfrey for her moral support and encouragement when I felt like giving up. She also read everything I wrote and provided feedback from a different perspective to my supervisors, which was greatly appreciated. I am particularly grateful for her patience when I was frustrated that, as a geologist, she couldn't identify raw material samples without smashing them into a million pieces; she also successfully hid her disappointment when I wouldn't let her do so. v SUMMARY TABLE OF CONTENTS Abstract .......................................................................................................................... iii Acknowledgements ......................................................................................................... iv Summary Table of Contents ............................................................................................ vi Table of Contents .......................................................................................................... vii List of Tables ................................................................................................................. xiii List of Figures ................................................................................................................ xix SECTION ONE: BACKGROUND TO THE RESEARCH Chapter 1 Introduction: Addressing the Scarcity of Archaeological Data from Inland Southwestern Australia .................................................................... 1 Theoretical Framework: Mobility, Subsistence and Technology ............... 16 Natural Environment and Landscape of the Study Area ........................... 36 Cultural Context ........................................................................................ 59 Chapter 2 Chapter 3 Chapter 4 SECTION TWO: RESOURCE AND OCCUPATION MODEL Chapter 5 Chapter 6 Chapter 7 Chapter 8 Constructing the Resource Model ............................................................ 76 Plant and Animal Foods .......................................................................... 101 Water Availability .................................................................................... 125 Occupation Model for the Study Area ...................................................... 151 SECTION THREE: ARCHAEOLOGICAL INVESTIGATIONS Chapter 9 Chapter 10 Chapter 11 Chapter 12 Chapter 13 Archaeological Methods ......................................................................... 173 Anderson Rocks ..................................................................................... 202 Gibb Rock ................................................................................................ 225 Mulka’s Cave ........................................................................................... 234 Identifying Technological Provisioning Systems ..................................... 275 SECTION FOUR: MERGING THEORY WITH ARCHAEOLOGY Chapter 14 Mobility, Site Use and Occupation Intensity ........................................... 297 Chapter 15 Conclusion: Occupation History of the Study Area ................................. 313 References ................................................................................................................... 343 APPENDICES Appendix A Appendix B Appendix C Appendix D Appendix E Appendix F Appendix G Flora Species Present in the Study Area ............................................... 376 Distribution of Plant Foods in the Study Area ......................................... 399 Distribution of Animal Foods in the Study Area ...................................... 418 Measuring Artefact Attributes ................................................................. 425 Anderson Rocks Stone Artefact Assemblage ........................................ 431 Gibb Rock Stone Artefact Assemblage ................................................... 445 Mulka’s Cave Stone Artefact Assemblage ............................................. 450 vi TABLE OF CONTENTS Abstract .......................................................................................................................... iii Acknowledgements ......................................................................................................... iv Summary Table of Contents ............................................................................................ vi Table of Contents .......................................................................................................... vii List of Tables ................................................................................................................. xiii List of Figures ................................................................................................................ xix SECTION ONE: BACKGROUND TO THE RESEARCH Chapter 1 Introduction: Addressing the Scarcity of Archaeological Data from Inland Southwestern Australia ....................................................... 1 1.1 Introduction ............................................................................ 1 1.2 Archaeological Research Bias in the Southwest ................... 1 1.3 Study Area, Aims and Research Questions ........................... 5 1.4 Method of Addressing Research Questions .......................... 7 1.4.1 Resource and Occupation Model ............................... 7 1.4.2 Archaeological Investigations ..................................... 10 1.5 Significance of this Research ................................................. 12 1.6 Organisation of Thesis ........................................................... 14 Chapter 2 Theoretical Framework: Mobility, Subsistence and Technology .. 16 2.1 Introduction .......................................................................... 16 2.2 Forager Theory .................................................................... 16 2.3 Technological Organisation ................................................. 20 2.3.1 Mobility ..................................................................... 21 2.3.2 Raw Material Availability........................................... 23 2.3.3 Risk........................................................................... 26 2.3.4 Summary .................................................................. 28 2.4 Technological Provisioning Systems ................................... 29 2.4.1 Individual Provisioning ............................................. 30 2.4.2 Place Provisioning .................................................... 31 2.4.3 Summary .................................................................. 34 2.5 Conclusion ........................................................................... 35 Chapter 3 Natural Environment and landscape of the Study Area ............... 36 3.1 Introduction .......................................................................... 36 3.2 Geology and Landform ........................................................ 36 3.3 Hydrology ............................................................................. 40 3.4 Climate ................................................................................. 44 3.4.1 Palaeoclimate of Southwestern Australia ................. 46 3.4.2 Modern Climate of the Study Area ........................... 50 3.5 Vegetation ............................................................................ 52 3.5.1 Botanical Provinces of Western Australia ................ 52 3.5.2 Vegetation Communities in the Study Area .............. 54 3.6 Fauna .................................................................................. 57 3.7 Conclusion ........................................................................... 58 vii Chapter 4 Cultural Context ........................................................................... 59 4.1 Introduction .......................................................................... 59 4.2 Archaeological Background ................................................. 59 4.2.1 Archaeological Sites in the Study Area .................... 60 4.2.2 Other Relevant Research ......................................... 65 4.3 Ethnohistorical Background ................................................. 70 4.3.1 Aboriginal Social Boundaries ................................... 71 4.3.2 Ethnohistory of the Study Area ................................ 73 4.4 Conclusion ........................................................................... 75 SECTION TWO: RESOURCE AND OCCUPATION MODEL Chapter 5 Constructing the Resource Model ................................................ 76 5.1 Introduction .......................................................................... 76 5.2 Landscape Divisions ............................................................ 76 5.3 Quantifying Plant and Animal Foods .................................... 80 5.3.1 Compiling Floral and Faunal Lists ............................ 81 5.3.2 Identifying and Classifying Food Species ................. 82 5.3.3 Determining Seasonal Availability ............................ 83 5.3.4 Benefits and Limitations of the Approach ................. 85 5.4 Quantifying Water Availability .............................................. 87 5.4.1 Representing Rainfall Variability............................... 87 5.4.2 Soil Water ................................................................. 88 5.4.3 Rock Structures ........................................................ 94 5.5 Formulating the Occupation Model ...................................... 98 5.6 Conclusion ......................................................................... 100 Chapter 6 Plant and Animal Foods ............................................................. 101 6.1 Introduction ........................................................................ 101 6.2 Plant Food .......................................................................... 101 6.2.1 Gum ........................................................................ 101 6.2.2 Lerps/Manna........................................................... 106 6.3.3 Storage Organs ...................................................... 106 6.3.4 Fruit ........................................................................ 108 6.3.5 Flowers ................................................................... 109 6.3.6 Leaves/Roots/Stems .............................................. 110 6.3.7 Seeds ..................................................................... 111 6.3.8 Summary ................................................................ 111 6.3 Animal Foods ..................................................................... 114 6.3.1 Birds and their Eggs ............................................... 114 6.3.2 Reptiles and their Eggs .......................................... 119 6.3.3 Mammals ................................................................ 121 6.3.4 Amphibians ............................................................. 121 6.3.5 Summary ................................................................ 122 6.4 Conclusion ......................................................................... 124 viii Chapter 7 Water Availability ........................................................................ 125 7.1 Introduction ........................................................................ 125 7.2 Soil Water .......................................................................... 125 7.2.1 Low Winter Rainfall – 2010, 2012........................... 125 7.2.2 Average Winter Rainfall – 2009, 2011 .................... 128 7.2.3 High Winter Rainfall – 1992, 1998 .......................... 131 7.2.4 Summary ................................................................ 134 7.3 Rock Structures.................................................................. 136 7.3.1 Low Winter Rainfall – 2010, 2012........................... 136 7.3.2 Average Winter Rainfall – 2009, 2011 .................... 138 7.3.3 High Winter Rainfall – 1992, 1998 .......................... 142 7.3.4 Summary ................................................................ 144 7.4 Summary of Water Availability in the Study Area .............. 146 7.5 Conclusion ......................................................................... 150 Chapter 8 Occupation Model for the Study Area ........................................ 151 8.1 Introduction ........................................................................ 151 8.2 Viability of Year-Round Occupation ................................... 151 8.3 Seasonal Route ................................................................. 154 8.3.1 Regular Seasonal Route ........................................ 154 8.3.2 Impact of Rainfall Variation .................................... 163 8.4 Inter-Group Aggregation .................................................... 164 8.5 Archaeological and Technological Signatures ................... 167 8.5.1 Northern and Central Study Area ........................... 168 8.5.2 Southern Study Area (Palaeovalley) ..................... 169 8.5 Conclusion ......................................................................... 171 SECTION THREE: ARCHAEOLOGICAL INVESTIGATIONS Chapter 9 Archaeological Methods ............................................................. 173 9.1 Introduction ........................................................................ 173 9.2 Field Methods ................................................................... 173 9.2.1 Site Selection ......................................................... 173 9.2.2 Survey and Excavation Methods ............................ 176 9.3 Lithic Analysis .................................................................... 177 9.3.1 ‘Horrid Little Bits of Stone’ – the Problem with Quartz Artefact Analysis ......................................... 178 9.3.2 Stone Artefact Classification................................... 180 9.3.3 Attribute Analysis .................................................... 184 9.4 Other Analysed Materials ................................................... 188 9.4.1 Charcoal ................................................................. 188 9.4.2 Sediments............................................................... 188 9.4.3 Ochre ...................................................................... 189 9.4.3 Glass ...................................................................... 189 9.5 Data Analysis ..................................................................... 192 9.5.1 Quantifying Site-Specific Water Sources................ 192 9.5.2 Constructing a Temporal Framework ..................... 194 9.5.3 Identifying Technological Strategies ....................... 195 9.5.4 Quantifying Occupation Intensity ............................ 199 9.6 Conclusion ......................................................................... 200 ix Chapter 10 Anderson Rocks ......................................................................... 202 10.1 Introduction ....................................................................... 202 10.2 Site Description ................................................................ 202 10.2.1 Water Availability .................................................. 204 10.3 Survey and Excavation Methods ..................................... 205 10.4 Organics and Dating ........................................................ 209 10.5 Stone Artefact Assemblage ............................................. 211 10.5.1 Surface Collections ............................................... 212 10.5.2 AR1 ....................................................................... 214 10.5.3 AR2-3 ................................................................... 216 10.6 Other Cultural Material ..................................................... 223 10.7 Conclusion ....................................................................... 223 Chapter 11 Gibb Rock ................................................................................... 225 11.1 Introduction ....................................................................... 225 11.2 Site Description ................................................................ 225 11.2.1 Water Availability .................................................. 225 11.3 Survey and Excavation Methods ..................................... 226 11.4 Organics and Dating ........................................................ 229 11.5 Stone Artefact Assemblage ............................................. 230 11.5.1 Surface Collections ............................................... 230 11.5.2 GR1 ...................................................................... 232 11.6 Conclusion ....................................................................... 232 Chapter 12 Mulka’s Cave .............................................................................. 234 12.1 Introduction ....................................................................... 234 12.2 Site Description ................................................................ 234 12.2.1 Water Availability .................................................. 236 12.3 Previous Research ........................................................... 239 12.4 Survey and Excavation Methods ..................................... 241 12.4.1 Main Cave Area .................................................... 241 12.4.2 The Humps ........................................................... 243 12.4.3 Camping Area ....................................................... 244 12.5 Organics and Dating ........................................................ 245 12.6 Stone Artefact Assemblage ............................................. 251 12.6.1 Entrance Platforms ............................................... 251 12.6.2 Column Sample .................................................... 252 12.6.3 MC1 ...................................................................... 253 12.6.4 The Humps Scatter ............................................... 255 12.6.5 TH1 ....................................................................... 256 12.6.6 Camping Area Scatter .......................................... 258 12.6.7 CA1 ....................................................................... 262 12.6.8 CA2 ....................................................................... 267 12.7 Other Cultural Material ..................................................... 272 12.8 Conclusion ....................................................................... 273 x Chapter 13 Identifying Technological Provisioning Systems......................... 275 13.1 Introduction ....................................................................... 275 13.2 Gibb Rock ........................................................................ 275 13.3 Anderson Rocks ............................................................... 277 13.3.1 Surface Collections ............................................... 277 13.3.2 AR1 ....................................................................... 278 13.3.3 AR2-3 ................................................................... 280 13.4 Mulka’s Cave .................................................................... 286 13.4.1 Main Cave Area .................................................... 286 13.4.2 The Humps ........................................................... 288 13.4.3 Camping Area ....................................................... 290 13.5 Conclusion ....................................................................... 295 SECTION FOUR: MERGING THEORY WITH ARCHAEOLOGY Chapter 14 Mobility, Site Use and Occupation Intensity ............................... 297 14.1 Introduction ....................................................................... 297 14.2 Gibb Rock ........................................................................ 297 14.3 Anderson Rocks ............................................................... 298 14.3.1 Mobility and Site Use ............................................ 298 14.3.2 Occupation Intensity Over Time ........................... 302 14.4 Mulka’s Cave .................................................................... 304 14.4.1 Mobility and Site Use ............................................ 304 14.4.2 Occupation Intensity Over Time ........................... 310 14.5 Conclusion ....................................................................... 312 Chapter 15 Conclusion: Occupation History of the Study Area..................... 313 15.1 Introduction ....................................................................... 313 15.2 Addressing the Research Questions ............................... 313 15.2.1 Theoretical Resource and Occupation Model ...... 314 15.2.2 Timescale of Aboriginal Occupation ..................... 321 15.2.3 Mobility and Occupation Patterns .......................... 324 15.2.4 Occupation Intensity and Water Availability ......... 331 15.3 Future Research Directions ............................................. 337 15.4 Conclusion ....................................................................... 339 References ................................................................................................... 343 xi APPENDICES Appendix A Flora Species Present in the Study Area ................................... 376 Appendix B Distribution of Plant Foods in the Study Area ............................ 399 B.1 Monthly Availability of All Plant Foods .................. 401 B.2 Monthly Availability of Plant Foods in Granite ....... 405 B.3 Monthly Availability of Plant Foods in Heath ......... 408 B.4 Monthly Availability of Plant Foods in Mallee ........ 410 B.5 Monthly Availability of Plant Foods in Saline ......... 412 B.6 Monthly Availability of Plant Foods in Thicket ....... 414 B.7 Monthly Availability of Plant Foods in Woodland .. 416 Appendix C Distribution of Animal Foods in the Study Area ......................... 418 Appendix D Measuring Artefact Attributes .................................................... 425 D.1 Metric Attributes .................................................... 425 D.2 Retouch ................................................................. 426 D.3 Breakage ............................................................... 427 D.4 Flake Scar Orientation ........................................... 428 D.5 Bipolar Reduction .................................................. 429 Appendix E Anderson Rocks Stone Artefact Assemblage ............................ 431 E.1 Cores ..................................................................... 432 E.2 Small Flaked Fragments ....................................... 433 E.3 Large Flaked Fragments ....................................... 434 E.4 Small Flakes .......................................................... 435 E.5 Large Flakes ......................................................... 437 E.6 Retouched Flakes ................................................. 443 Appendix F Gibb Rock Stone Artefact Assemblage ...................................... 445 F.1 Cores ..................................................................... 446 F.2 Small Flaked Fragments ....................................... 446 F.3 Large Flaked Fragments ....................................... 447 F.4 Small Flakes .......................................................... 447 F.5 Large Flakes .......................................................... 448 Appendix G Mulka’s Cave Stone Artefact Assemblage ................................ 450 G.1 Cores .................................................................... 451 G.2 Small Flaked Fragments ....................................... 452 G.3 Large Flaked Fragments ....................................... 453 G.4 Small Flakes ......................................................... 456 G.5 Large Flakes ......................................................... 458 G.6 Retouched Flakes ................................................. 469 xii LIST OF TABLES Table 2.1 Forager models and theories for achieving optimisation, based on the references listed and summary data from Bettinger et al. (2015), Clarkson (2007), Kelly (2013) and Winterhalder (1981). References denote the first iteration of a model or significant subsequent contribution. 18 Table 2.2 Behavioural characteristics arising from different resource structures. Re-tabulated from Ambrose and Lorenz (1990:10). 19 Table 3.1 Summary of main palaeoenvironmental data from southwestern Australia – research focussing on the recent past has been omitted (e.g. Cullen and Grierson 2009; Treble et al. 2003). Temperature data from Gouramanis et al. (2012) as it does not span the entire record. Dates are cited in the format provided by the author. * = Perth metro area, not marked on Figure 3.7. 48 Table 3.2 Vegetation communities found in the study area. Community names, descriptions and classifications are based on Beard (1972, 1975, 1976, 1979, 1980). Note that these descriptions apply only to the study area (i.e. the Hyden, Muntadgin and Skeleton Rock systems), and may not reflect community characteristics at a broader scale. 55 Table 4.1 Archaeological features in the study area, excluding those analysed for this research. ID, name and classification taken from DPLH’s Aboriginal Heritage Inquiry System (AHIS), which separates Registered sites (RS) from ‘Other Heritage Places’ (OHP). Unless noted otherwise, information derives solely from DPLH site files. The map code indicates the position of a particular feature or set of features on Figure 4.1—features are numbered from west to east. 62 Table 4.2 Archaeological data from Smith’s (1993:271) interior zone, showing the number of sites in different topographic units, and the number of sites falling within different assemblage size categories. 70 Table 5.1 Landscape Divisions and their associated vegetation communities (see descriptions in Table 3.2 and distribution in Figure 3.14) and landform. Vegetation communities are defined following Beard (1972, 1975, 1976, 1979, 1980). 77 xiii Table 5.2 Factors governing seasonal availability for plant and animal foods, and generalisations used to determine periods of availability. Where no governing factors were present the items would be available year-round. Generalisations were employed where data were lacking. For details, see Appendices B–C. 85 Table 5.3 Years chosen to represent rainfall variability, showing annual and winter rainfall totals (mm) and classification, and the number of isolated falls > 15 mm, occurring outside winter (from Hyden station, BoM ID 010568). 88 Table 5.4 Soil Water Tool (DPIRD 2016a) data fields and inputs used. A soil profile was selected to represent each Landscape Division (provided in parentheses after profile) based on generalised soil characteristics (Beard 1979, 1990; DPIRD 2016b; Murphy-White 2007; Schoknecht and Pathan 2013). Model years were chosen to represent a variety of rainfall quantities and distributions, as per Chapter 5.4.1. The 1992 and 1998 model periods were abridged due to unusually high Plant Available Water (PAW) balances occurring at germination, resulting from model mechanics whereby plants begin to draw on a store of water accumulated under preceding fallow conditions. This does not accurately represent land under native vegetation, as plants would be drawing water yearround. The April break-season rule was voided to ensure that germination occurs during winter, so model periods were more directly comparable. A single crop coefficient was used throughout, to simulate mature native vegetation. 91 Table 5.5 Ranked plant food categories, based on return rates (cal/hr) determined by handling time (hr/kg) and nutritional value (cal/kg). Handling time and nutritional content are given only where return rates were calculated manually. Asterisks indicate inferred handling times. 99 Table 6.1 Total number of plant food products found in the study area. The number of plant species providing a particular food item is listed, so the overall total indicates the total number of plant food items in the study area rather than the number of food-bearing species. A species may yield more than one edible product. Plant uses are based on Bindon (1996), Low (1991), Meagher (1974) and Webb (2000); not all uses correspond to each food item and/or species. Food items are ranked from highest to lowest, as per Table 5.5. 102 xiv Table 6.2 Spatio-temporal availability of all plant food categories within the study area, showing the number of species providing specific products during particular months. Foods are ordered by rank, from highest to lowest. 103 Table 6.3 Spatio-temporal availability of all animal food categories within the study area, showing the number of species providing a certain product during each month, across all Landscape Divisions. 116 Table 7.1 Longevity of PAW (days) in modelled years, across all six Landscape Divisions. 134 Table 7.2 Winter and annual rainfall totals (mm) for each of the modelled years, and the longevity of water (days) in each rock structure. Years have been arranged by annual rainfall total (low to high). Note that runoff gnammas frequently held water from the end of one year and into the next – this is not reflected in the longevity figures. 145 Table 7.3 Longevity of water (days) in various soils and rock structures, as a result of differently sized isolated falls. Durations are calculated assuming that falls occurred in December and were received by a dry structure or profile. Daily loss for soils is based on average December figures (4.4 mm for Granite, 4.8 mm for Thicket), and recharge amounts as per Chapter 7.2.2. Rainfall data sourced from Bureau of Meteorology, for Hyden (010568) for 1929– 2017 (no data for 1947). Asterisks indicate that a structure or profile has reached its storage capacity, limiting longevity. 149 Table 8.1 The volume of water held by various capped rock structures and the longevity of water supply (people days). People days = number of people x number of days, so ten people days may be one person consuming water for ten days, or ten people consuming water for a single day. Daily consumption was estimated at 1.5 L per person. 152 Table 8.2 Temporal distribution of ranked plant food items (following Table 5.5), and unranked animal foods in each Landscape Division. 157 Table 9.1 Artefact type definitions and the attributes recorded for each. Flaked fragments were separated based on their maximum dimension, and flakes by their length. 182 xv Table 9.2 Stone artefact attributes – further information regarding methods can be found in Appendix D. Negative flake scars < 5 mm in maximum/percussion length have been omitted as they more likely result from retouch or use, and single scars can be very difficult to separate (and therefore quantify) on all but very high-quality quartz. The references provided point to a definition of a particular attribute, and its method of measurement. 183 Table 9.3 Features used to identify modified and unmodified glass, and the attributes recorded for each. To be classified as modified glass, the piece needed to meet at least one of the criteria listed – all other pieces were classified as unmodified. Criteria are based on Allen and Jones (1980), Goward (2011:97), Harrison (2000), Holdaway and Stern (2004:30), Paterson (1999), Runnels (1976), Veth and O'Connor (2005) and Williamson (2004). Maximum dimensions were recorded via the same method used for stone artefacts (see Table 9.2 and Appendix D). Edge damage is only considered evidence of modification only where the site is relatively undisturbed, i.e. receives little foot or vehicular traffic. 191 Table 9.4 Artefact attribute indicators for dominant and supplementary technological provisioning systems, using the pairing system established in Figure 9.5. The indicators are only of value within paired systems (Individual Provisioning and Gearing Up, Place Provisioning and Short-term Expedient), since they describe the relative position of each strategy on a continuum, whose range of values will be different to that of the alternate pair. 198 Table 10.1 Dimensions, capacity, shape, evaporation coefficient (EC) and modelled runoff conditions for the three gnammas at Anderson Rocks. Depth and radius were measured for each structure and used to calculate capacity following the standard mathematical formula for the volume of a cone (see Figure 9.4). 205 Table 10.2 Soil pH values for selected XUs within AR1 and AR2-3. 209 Table 10.3 Charcoal samples and their returned radiocarbon ages, from pit AR2-3. Dates are arranged by XU depth. Asterisk beside sample weight indicates in-situ sample – all others derived from sieve residue. Note that no 13C measurements were provided with results. 211 Table 10.4 Number of stone artefacts from Anderson Rocks pits and surface collections. 212 xvi Table 10.5 Artefact types in surface collections at Anderson Rocks. Figures in parentheses refer to high-quality materials. Coll. = collection. Sc. = Scatter. OSM = other surface material, from outside defined scatters. 212 Table 10.6 Stone artefact types by XU in AR1. Figures in parentheses refer to high-quality raw materials. 215 Table 10.7 Stone artefact types by XU in AR2-3. Figures in parentheses refer to high-quality raw materials. 217 Table 10.8 Location, weight and colour of ochre samples from AR23. 223 Table 11.1 Number of stone artefacts from the Gibb Rock pit and surface collections. 230 Table 11.2 Frequency of artefact types found at Scatter 1 and Scatter 2. 230 Table 11.3 Frequency of artefact types by XU in GR1. 232 Table 12.1 Surface dimensions, capacity, shape, evaporation coefficient (EC) and modelled runoff conditions for the six gnammas at Mulka's Cave. Dimensions of five traditional gnammas after Webb and Rossi (2008), cleft rockhole from Gunn (2004, 2006). Volumetric capacities were calculated using standard mathematical formulae provided in Figure 9.4. 237 Table 12.2 Radiocarbon dates from inside Mulka's Cave. Asterisk denotes AMS date. CS = Column Sample (cut in 2006), P88 = Bowdler et al. (1989) pit, excavated in 1988. Dates have been arranged by depth below the 2006 ground surface using Figure 12.8. Note that Bowdler et al. (1989) excavated by stratigraphic unit, so the thickness of each unit varied in different parts of the pit. The maximum depths have been cited below. All CS samples came from sieve residue, while the location of P88 samples are unknown. 249 Table 12.3 Radiocarbon dates for the Humps and Camping Area pits. Asterisk denotes AMS date. All samples were collected from sieve residue. 250 Table 12.4 Number of stone artefacts from Mulka’s Cave pits and surface collections. Assemblages have been grouped by site area. 251 xvii Table 12.5 Stone artefact types by XU from the Column Sample inside Mulka’s Cave. Figures in parentheses refer to highquality quartz. 252 Table 12.6 Stone artefact types by XU in MC1. Figures in parentheses refer to high-quality quartz. 254 Table 12.7 Stone artefact types by XU in TH1. Figures in parentheses refer to high-quality quartz. 256 Table 12.8 Stone artefact types in the Camping Area Scatter. Figures in parentheses refer to high-quality materials. Note that ‘other CA’ refers to the low-density material found outside the clusters, but within the Camping Area Scatter. 258 Table 12.9 Stone artefact types by XU in CA1. Figures in parentheses refer to high-quality materials. 262 Table 12.10 Stone artefact types by XU in CA2. Figures in parentheses refer to high-quality materials. 267 Table 12.11 Location, weight and colour of ochre samples from Mulka's Cave. 272 xviii LIST OF FIGURES Figure 1.1 Map of dated sites in Australia, after Williams et al. (2014: Figure 1). The Holocene sites pre-date 10,000 BP. Black box encases all sites in Noongar country; the specific boundary of the Single Noongar Native Title Claim is shown in Figure 1.2. 3 Figure 1.2 Left: Location of the Single Noongar Native Title Claim (grey) and the study area boundary (blue). Right: Enlarged view of the study area, showing the location of archaeological sites mentioned in the text. The eastern boundary follows the State Barrier Fence. Base imagery from Google earth. 6 Figure 2.1 Optimal and suboptimal strategies for different resource structures (Ambrose and Lorenz 1990:15). 19 Figure 2.2 The relationship between quality and abundance of local stone, and the use of formal and informal tool production (Andrefsky 1994:30). 26 Figure 2.3 Impact of mobility on provisioning strategies: A (left): frequency of moves and occupation duration; B (right): length of logistical forays. (Graf 2010, after Kuhn 1995). 30 Figure 3.1 Simplified geology of Western Australia, showing extent of the Yilgarn Craton, also known as the Yilgarn Block (Beard 1990:41). 37 Figure 3.2 Major regions within the Yilgarn Craton (Anand and Butt 2010:1018). 38 Figure 3.3 Geology of characteristic Wheatbelt valleys (Commander et al. 2001, reproduced from Department of Environment 2005b). 39 Figure 3.4 Geology of selected parts of the study area (A) and beyond (B), showing the occurrence and typical distribution of potential tool-stone. Geological imagery (including legend) from Geological Survey of Western Australia (1984). Note that geological maps depict the surface and near-surface geology, so raw material sources may not have all been accessible at groundlevel. 41 Figure 3.5 Groundwater salinity across the Yilgarn Craton (Anand and Butt 2010:1078). 43 xix Figure 3.6 Recently dead vegetation near Lake Gounter, Hyden, in October 2016 (Photograph: A M Rossi). Note the replacement of large woody shrubs by much smaller plants with a greater salt tolerance. 44 Figure 3.7 Palaeoenvironmental sites mentioned in the text. Those in and around Perth are not marked. 47 Figure 3.8 Major climatic trends identified in southwestern Australian palaeoenvironmental studies. Periods were classified in relation to the preceding period; blue =wetter, red = drier, black = no change, grey = uncertain. All dates have been plotted as cited, although Churchill’s (1968) dates have been converted from BC/AD to cal. BP and plotted on the secondary axis with the data obtained by Gouramanis et al. (2012). 47 Figure 3.9 Harrison’s (1993:219) reconstruction of lake status at 1000-year intervals, from 6000–1000 BP. 49 Figure 3.10 Bioclimatic zones of Western Australia (Beard 1990:39). 51 Figure 3.11 Annual rainfall data from Hyden weather station (no. 10568), 1929–2017 in relation to average rainfall (blue dashed line). Data are not available for 1947 or 1974. 51 Figure 3.12 A: Natural regions of Western Australia (Beard 1990:34); B: IRBA 7 regions of Western Australia (Australian Government Department of Environment and Energy 2012). Abbreviations largely reflect Beard’s (1990) subregion names except: AVW = Avon Wheatbelt; COO = Coolgardie, GES = Geraldton Sandplains, JAF = Jarrah Forest, SWA = Swan Coastal Plain, VIB = Victoria Bonaparte, YAL = Yalgoo. 53 Figure 3.13 Catenary sequence for the Mallee Region (top) and Avon Wheatbelt Region (bottom), showing the relationship between topography and vegetation communities (Beard 1990:116, 129). 54 Figure 3.14 Pre-European distribution of vegetation communities in the study area. Spatial data extracted from NRInfo (DPIRD 2016b), based on Beard’s (1972, 1979, 1980) data. Continuous black line separates the two IBRA regions that meet in the study area: the Avon Wheatbelt (north) and Mallee (south). Dashed line represents the boundary of the study area. 56 xx Figure 4.1 Registered sites and Other Heritage Places in the study area. White squares denote approximate site positions, enlarged in inset maps to show their immediate surroundings. Locational data were retrieved from Aboriginal Heritage Inquiry System (AHIS), maintained by the Department of Planning, Lands and Heritage. The central point of sites/place marked on insets. Site details can be found in Table 4.1. 61 Figure 4.2 Locations mentioned in the text. 65 Figure 4.3 Area covered by South West Native Title Settlement, showing subdivisions (Source: National Native Title Tribunal). ILUA = Indigenous Land Use Agreement. 72 Figure 4.4 Aboriginal social boundaries in southwestern Australia according to various sources (redrawn from original maps by A.M. Rossi). A: Curr (1886,1887) separated groups into the non-circumcising Western Division (green) and the circumcising Central Division (beige). Group names are given in italics, where known. Red crosses indicate locations where circumcision was practised, but recorded locations are often vague and their placement on the map may be somewhat arbitrary. E.g. the cross within Kokor territory is noted in text as ‘east of York’. Original image provided courtesy of Fisher Rare and Special Collections, University of Sydney. B: Bates (White 1985) placed the division for her non-circumcising group (the Bibbulmum) further inland. Her eastern boundary locations are shown in blue. At the time of her research, Karlgarin was the most inland town established in the vicinity of the study area, so was the only real point of reference. The western boundary for the circumcising group (Karatjibbin – specific locations shown in red) is considerably further east. C: Davidson (1938). D: Tindale (1974). Groups east of his circumcision/subincision boundary are not shown. 73 xxi Figure 5.1 Vegetation and landform characteristic of each Landscape Division – all photographs and species information from Beard (1990:119, 121, 123, 131–132, 134, 136). A–B: Saline, dominated by short, salt-tolerant shrubs (Atriplex spp., Tecticornia spp.). Note the taller Eucalyptus sp. trees on slightly higher ground in B. C: Eucalyptus spp. Woodland, note the generally open nature of the canopy and the relatively sparse understorey. D: Mallee, note the multi-stemmed Eucalypts and the thicker understorey than in C. E–F: Heath, dominated by a diverse range of low shrubs including Grevillea spp. and Verticordia spp. Note the Mallee in the background of E, indicating the mosaic nature of Landscape Divisions. G: Thicket of Acacia, Allocasuarina and Melaleuca species – note the closely spaced shrubs. H: Granite. Note the thicker vegetation at the base of the slope, due to water runoff. The rock is mostly bare except where pockets of soil support shrub and small tree growth. 78 Figure 5.2 Distribution of Landscape Divisions within the study area – note that the Saline Landscape Division comprises both saline valleys and salt lakes. Spatial data extracted from NRInfo (DPIRD 2016b). Examples of each Landscape Division are shown in Figure 5.1. 79 Figure 5.3 The relative water content of various soil types, showing available and unavailable water (Moore 2001:82). Available water can be extracted by plants, referred to here as Plant Available Water (PAW). Note that a soil may be holding water, even if none is available for plant use (unavailable water); coarser soils (those with larger particle sizes) have lower water-storage capacity than finer soils but require less input before plants are able to extract water. 89 Figure 5.4 Example output from the Soil Water Tool (DPIRD 2016a), showing soil water (PAW) under fallow (black line) and cropped conditions (green line), plus daily rainfall (blue bar). Fallow and cropped values will be in-sync until germination date ('break'); thereafter, modelled crops draw on stored water accumulated under preceding fallow conditions. An increase in the PAW balance is registered only when daily input (rainfall) exceeds daily loss (through evaporation/evapotranspiration). Daily data values (box) are displayed when the mouse is hovered over the relevant point on the chart output (black filled circle). 92 xxii Figure 6.1 Examples of plant foods found in the study area – photographs and information sources as cited. A: Acacia acuminata, 3–5 m tall, bearing edible seeds and gum (Bindon 1996:251). B: Edible gum exuded from Acacia spp. (Low 1991:152). C: Typical Acacia seed pod (Sweedman and Merritt 2006:176). D: Woody fruits (2–4 cm long) containing edible seeds of Allocasuarina humilis (Erickson et al. 1979:21). E–F: Eucalyptus redunca and its flowers; tree to 6 m tall (Bindon 1996:131). G–H: Lerps on Eucalyptus spp. leaves (Low 1991:153). I: Chamelaucium megalopetalum, edible flowers 10–15 mm wide (Erickson et al. 1979:88). J: Enchylaena tomentosa fruits, approximately 5 mm wide (Low 1991:167). K: Thysanotus patersonii tubers, generally 10–30 mm long and 5 mm wide (Bindon 1996:251). L: Carpobrotus virescens, fruits and leaves (up to 65 mm long) are both eaten (Daw et al. 1997:4–5). M: Atriplex vescaria shrub (≤ 1 m high) which bears edible leaves (Bindon 1996:46). N: Edible fruits of Santalum acuminatum, tree grows to 5 m, fruit 20–40 mm wide (Daw et al. 1997:52–53). 105 Figure 6.2 Plant food categories demonstrating strong temporal variation in availability, showing monthly species diversity within each Landscape Division. The same colour key applies to each chart. 107 Figure 6.3 Total number of different plant foods within each Landscape Division. 112 Figure 6.4 Total number of plant food items (top) and rank (bottom) available in each month across the entire study area. 113 Figure 6.5 Examples of birds found in the study area. Measurements refer to the body length (from crown to tip of the tail) unless noted otherwise. Common names follow species name. Photographs and measurements from Nevill (2014:52–55, 78–81, 90–91, 106–111). A–C: Passeriformes (perching birds): Phylidonyris niger, whitecheeked honeyeater, 18 cm (A); Pomatostomus superciliosus, white-browed babbler, 21 cm (B); Petroica goodenovii, red-capped robin, 12 cm (C). D–G: Waterbirds: Himantopus himantopus, black-winged stilt, 37 cm (D); Tadorna tadornoides, Australian shelduck, 56– 72 cm (E); Charadrius ruficapillus, red-capped plover, 15 cm (F); Vanellus tricolor, banded lapwing, 27 cm (G). H: Dromaius novaehollandiae, emu, 160–190 cm tall. I: Leipoa ocellata, malleefowl, 55–60 cm. J: Barnardius zonarius, Australian ringneck, 36 cm. 115 xxiii Figure 6.6 Monthly availability of birds (top) and birds’ eggs (bottom) in each Landscape Division. The same colour key applies to each chart. 118 Figure 6.7 Example of native non-avian fauna present in the study area. Common names are provided after species name, measurements cited are maximum total length (TL – including tail) or body length (BL – excluding tail) Photographs and measurements from Nevill (2014:124– 129, 136–137, 142–143, 151, 156–157, 162–163, 184– 185). A–D: Reptiles: Ctenophorus cristatus, bicycle dragon, TL 370 mm = (A); Liopholis multiscutata, bull skink, TL 250 mm (B); Underwoodisaurus milii, barking gecko, TL 165 mm (C); Varanus gouldii, sand monitor, TL 1.6 m (D). E–K: Mammals: Sminthopsis crassicaudata (family Dasyuridae), fat-tailed dunnart, BL 75 mm (E); Dasyurus geoffroii (family Dasyuridae), chuditch, BL 310– 360 mm (F); Macropus fuliginosus, western grey kangaroo, BL 0.95–2.2 m (G); Chalinolobus gouldii, Gould’s wattled bat, BL 65–75 mm (H); Isoodon obesulus, southern brown bandicoot, BL 340 mm (I); Trichosurus vulpecula, common brushtail possum, BL 380 mm (J); Tachyglossus aculeatus, short-beaked echidna, BL 400 mm (K). L: Myobatrachus gouldii, turtle frog, BL 50 mm. 120 Figure 6.8 Total number of animal food items available in each month across the entire study area. 123 Figure 6.9 Number of different animal food items within each Landscape Division. Note that Heath and Thicket both preserve a single amphibian species, not visible at the scale of this figure. 123 Figure 7.1 Daily moisture balance of soils in each Landscape Division (mm of plant available water – PAW – coloured lines) and rainfall (black bars) for years with low winter rainfall. Note that the same colour key applies to both. 127 Figure 7.2 Daily moisture balance of soils in each Landscape Division (mm of plant available water – PAW – coloured lines) and rainfall (black bars) for years with average winter rainfall. Note that the same colour key applies to both. 129 Figure 7.3 Daily moisture balance of soils in each Landscape Division (mm of plant available water – PAW – coloured lines) and rainfall (black bars) for years with high winter rainfall. Note that the same colour key applies to both. 133 xxiv Figure 7.4 Daily water balance (mm) in rock structures (coloured lines) and rainfall (black bars) for years with low winter rainfall. Note that the same colour key applies to both. 137 Figure 7.5 Daily water balance (mm) in rock structures (coloured lines) and rainfall (black bars) for years with average winter rainfall. Note that the same colour key applies to both. 141 Figure 7.6 Daily water balance (mm) in rock structures (coloured lines) and rainfall (black bars) for years with high winter rainfall. Note that the same colour key applies to both. 143 Figure 7.7 Frequency of different rainfall scenarios in the study area from 1929–2017. A (left): number of low, average and high winter rainfall years. B (right): frequency of variously sized isolated falls. Based on Bureau of Meteorology data for the Hyden station (010568); note no data were recorded in 1947. 147 Figure 7.8 Generalised water availability in soils and rock structures in each Landscape Division, in average (top), low (middle) and high (bottom) winter rainfall years. Note that Heath, Mallee, Saline, Thicket and Woodland areas only provide soil water. Solid lines indicate the generally consistent presence of water over the specified period, i.e. being dry for no more than a few days at a time. Dashed lines indicate when water may be available – its presence/absence depends on the timing of rainfall events. Unusually large pre- or post-winter falls have been voided (see Chapter 5.4.3), as they mask the normal longevity of water from winter rain. Water-holding periods for average (2009, 2011) and high (1992, 1996) winter rainfall years are based on the mean of both years. For low winter rainfall years (2010, 2012), only 2012 data was used as they were deemed more broadly representative of low rainfall conditions. In winter 2010 a few large falls were interrupted by three weeks of limited rainfall, causing soils to dry out when they normally would not have. Mallee soils are not pictured for low rainfall conditions as they held water for just a single day. 148 Figure 8.1 Occupation model for the study area under average rainfall conditions, showing the primary water sources (blue), primary foraging (green) and residential areas (yellow) and the optimal portion of the study area (orange) that should be occupied at particular times of the year. For division location see Figure 8.2. 155 xxv Figure 8.2 Location of the northern, central, southern parts of the study area and the optimal months to occupy each. Note that these divisions are not firm boundaries, but merely indicative of the broad areas referred to in the text and Figure 8.1. 156 Figure 9.1 Location of archaeological sites excavated by the author. 174 Figure 9.2 Procedural key used to classify flaked stone via the materialist approach developed by Hiscock (2007). Unflaked stone preserves no evidence of flake removals or having been detached from a larger piece. Note that ventral surfaces were identified via any of the normal landmarks, such as bulb of percussion, ripple marks, termination etc, following Holdaway and Stern (2004:108). Retouch scars must be initiated from or modify the ventral surface – see Table 9.1 and Appendix D.2 for more information. Note the amendments for quartz-dominated assemblages, including the use of impact points or shatter facets to identify flaked quartz, as well as flaked fragments incorporating those pieces where breakages or poor-quality material make it difficult to confirm the presence/absence of a ventral surface. Similarly, poor-quality material likely masked retouch in several instances – where such uncertainty was encountered, pieces were classified as in the lower category, as flakes. 181 Figure 9.3 Variable quality of quartz within the study area. Quality increases from left to right. Top row: very poor. Second row: poor–moderate. Third row: moderate–good. Bottom row: high quality. White bars are 1 cm. 186 Figure 9.4 Shapes used to represent morphological variation between individual gnammas, and the mathematical formulae used to determine their total capacity and the volume of water with depth. 193 Figure 9.5 Major differences in technological organisation between the four provisioning systems: Place Provisioning (PP), Individual Provisioning (IP), Gearing Up (GU), and Shortterm Expedient (EXP). 196 xxvi Figure 10.1 Features seen at Anderson Rocks – all photographs by A.M. Rossi. Individual pictured is 1.64 m tall. A–B: Northfacing view of quartz vein striking north-south up the outcrop. C: Large pan (10.9 m wide) on top of outcrop, facing east. Note the outflow channel leading downslope. D: Deep channels created by water flow – these link many pans together in an interconnected system. E–F: Pans linked by channels. Note the higher elevation of some pans that, when full, would overflow into the channels directing water to lower lying structures. G: South-east facing view of large pan that overflows downslope to gnammas (behind small circular patch of shrubs in top right of photo). This pan receives runoff from upper slopes and channel systems. H: North-west facing view of the two gnammas that receive runoff from upslope pan and channel system (which continues to flow after rain has ceased), as well as the surrounding slope. Note that the well-defined flow of water does not reach the third gnamma, which lies to the left of those pictured. I: South-east facing view of all three gnammas – the black staining indicating inflow channels to the left and central gnamma. The smallest gnamma (right) receives runoff only from its immediate surrounds. Largest gnamma (left) has diameter of 1.43 m. 203 Figure 10.2 Total volume of water stored within Anderson Rocks gnammas during high (top), average (middle) and low (bottom) winter rainfall years. 206 Figure 10.3 A: Aerial photograph of Anderson Rocks. White square indicates the area magnified in image B. B: Main surveyed area of Anderson Rocks, showing the location of gnammas, artefact scatters and pits. Note the thicker vegetation separating the eastern and western exposures. Features too small to outline at the scale of this figure have been indicated by arrows. Base imagery from Google Earth. 207 Figure 10.4 A: Vegetated area near Scatter 1 and pit AR1, facing north-east. Note that the scatter and pit were immediately downslope of the thick vegetation pictured, where only low ground cover was present. B: Scatter 2, located within a small pan bisected by a quartz vein, facing north. Pan measures 1.63 m long by 0.91 m wide (maximum dimensions). C: close-up of the quartz vein in Scatter 2. D: Location of Scatter 3 and pit AR2-3 within open vegetated area southeast of the gnammas, facing southwest. Note the granite slab marks the position of the backfilled pit. All photographs by A.M. Rossi. 208 xxvii Figure 10.5 Charcoal weight with depth in AR1 (left) and AR2-3 (right). 210 Figure 10.6 Depth-age curve for AR2-3. Solid line joins radiocarbon age determinations, while linear trendline (dotted line) was used to estimate the age of the lowest cultural material, from XU32 (-825 mm). 211 Figure 10.7 Artefact discard rates in AR2-3. The rate has been adjusted to artefacts/0.5 m2/100 years to allow for the difference in excavated area in the lower deposits, as only pit AR3 continued below 600 mm. 218 Figure 10.8 Proportion of small flakes on high-quality material (left), and breakage rate for all small flakes (right) from AR2-3. 220 Figure 10.9 Selected attributes recorded on large quartz flakes, by XU, in AR2-3. A-B (top): Average negative flake scar (blue bars) and platform counts (orange bars) for highquality materials (A, left) and lower-quality materials (B, right). C-D: average mm/scar values (C, left) and proportion of broken flakes (D, right) for high-quality materials (black bars) and lower-quality materials (grey bars). 221 Figure 11.1 Shallow pans on eastern side of Gibb Rock; B: Fenced off wall to capture runoff from the outcrop, facing west. Both photographs by A.M. Rossi. 226 Figure 11.2 A: Aerial photograph of Gibb Rock. White rectangle indicates area magnified in image B. Note the extensive wall constructed to filter runoff to the large circular tank, near the car park on the eastern side of the outcrop. B: Location of Scatter 1 and Scatter 2 (within white boundaries), and pit GR1 (red arrow). Note the thicker vegetation north of the car park area. Base imagery from Google Earth. 227 Figure 11.3 A: Area around Scatter 1, facing south. B: Area around Scatter 2, facing east. C: quartz weathering from parent granite, in Scatter 1. All photographs by A.M. Rossi. 228 Figure 11.4 Weight of charcoal with depth in pit GR1. 229 xxviii Figure 12.1 Cultural and natural features present at The Humps. All photographs by A.M. Rossi. A: the main entrance to Mulka's Cave, facing west. Note the raised platforms either side of the entrance, created when erosion removed the sediments below. B: Two paintings inside the entrance to Mulka's Cave, note the yellow handstencil inside the circular motif. Lefthand painting measures 127 x 36 cm, righthand painting 83 cm wide (Gunn 2004:26). C: Raised walkway outside Mulka's Cave, facing south. D: Red and white pigmented handstencils and handprints inside Mulka's Cave. E: View of raised walkway from the top of Mulka's Cave. F: Gnamma 1, measuring 1.4 m long and 0.85 m wide. Note the shadows cast by the gnamma walls, protecting the water from evaporation. 235 Figure 12.2 A: Aerial photograph of The Humps. White rectangle indicates the area magnified in image B. B: Location of natural and cultural features around the Humps, including Mulka's Cave itself, the Humps Scatter, Camping Area Scatter, lizard trap and the main water sources. Features too small to outline at the scale of this figure are indicated by arrows. Pit locations for Mulka's Cave and the Camping Area are illustrated in Figure 12.4 and Figure 12.6. Locations for cleft rockhole and Gnamma 5 are approximations, based on Figure 3 from Gunn (2006:22). Base imagery from Google Earth. 236 Figure 12.3 Total volume of water available at Mulka's Cave from the gnammas and cleft rockhole (solid line) and the gnammas only (dashed line), during high (top), average (middle) and low (bottom) winter rainfall years. 238 Figure 12.4 Main Cave Area, facing west, showing the location of the Entrance Platforms (flat areas indicated by white arrows), Column Sample and pit MC1 (red arrows). Photograph: A.M. Rossi. 243 Figure 12.5 A: View of The Humps outcrop taken from the area around Gnammas 1 and 2, facing south west. Note the sparsely vegetated shoulder on which the Humps Scatter is located. B: The Humps Scatter, showing pit TH1, facing north east. Gnammas 1 and 2 occur on the flat granite exposure visible in the top right of the image. C: A portion of the Camping Area, facing west. Note the relatively clear area surrounded by thicker border vegetation – no artefacts were present in the latter area. 244 xxix Figure 12.6 Camping Area artefact scatter, showing the extent of surface artefacts (light grey shading), the location of the high-density surface clusters (Clusters A–F, dashed circles) and pits CA1 and CA2. Note the scatter is surrounded by thick border vegetation (dark grey shading). 245 Figure 12.7 Charcoal weight by XU within pits CA1 (top), CA2 (middle) and TH1 (bottom). 246 Figure 12.8 Stratigraphic section (Bowdler et al. 1989: Figure 4.2a) superimposed with their XU levels (blue), sediment column (orange) and the approximate position of the 2006 Column Sample (pink). XUs numbered by each author. Rossi (2014a) after Bowdler et al. (1989). 248 Figure 12.9 Depth-age curves for inside Mulka’s Cave (left) and other parts of the site (right). Solid lines join radiocarbon age determinations, while linear trendlines (dotted lines) were used to estimate the age of the lowest cultural material, represented by the end point of those lines. Inside Mulka’s Cave, Wk-27113 (P88 XU15) and Wk-29188 (P88 XU16) were statistically indistinguishable. The lower sample was selected as it was derived from a larger charcoal sample. 249 Figure 12.10 Small core from MC1 XU4. Note the presence of four original crystal facets (indicated by blue arrows). 254 Figure 12.11 Artefact discard rates for pit TH1. 256 Figure 12.12 Selected attributes recorded on large quartz flakes within the Camping Area Scatter. Cl. = Cluster. A–B (top): Average negative flake scar (blue bars) and platform counts (orange bars) for high-quality materials (A – left) and lower quality materials (B – right). C–D: average mm/scar values (C – left) and proportion of broken flakes (D – right) for high-quality materials (black bars) and lower quality materials (grey bars). 260 Figure 12.13 Artefact discard rates for pit CA1. 262 Figure 12.14 Proportion of small flakes on high-quality material (left), and breakage rate for all small flakes (right) from pit CA1. 264 xxx Figure 12.15 Selected attributes recorded on large quartz flakes, by XU, in CA1. A–B (top): Average negative flake scar (blue bars) and platform counts (orange bars) for high-quality materials (A – left) and lower quality materials (B – right). C–D: average mm/scar values (C – left) and proportion of broken flakes (D – right) for high-quality materials (black bars) and lower quality materials (grey bars). 265 Figure 12.16 Dorsal surface of three quartz flakes from CA1 XU2. Note the brown crust that covers most of the dorsal surface (including negative flake scars) and has been removed in other areas; this crust could not be removed by washing or rubbing. Crust was absent on ventral surfaces. White bar = 1 cm. 266 Figure 12.17 Artefact discard rate in pit CA2. 267 Figure 12.18 Proportion of small flakes on high-quality material (left), and breakage rate for all small flakes (right) from pit CA2. 269 Figure 12.19 Selected attributes recorded on large quartz flakes, by XU, in CA2. A–B (top): Average negative flake scar (blue bars) and platform counts (orange bars) for high-quality materials (A – left) and lower quality materials (B – right). C–D: average mm/scar values (C – left) and proportion of broken flakes (D – right) for high-quality materials (black bars) and lower quality materials (grey bars). 270 Figure 12.20 A (left): quartz flake from CA2 XU4. Lateral margins preserve notches formed by previous flake removals. Black bar = 1 cm. B (right): an idealised core that could produce notched flakes. Note the long, parallel flake scars, and the scalloped edge evident in cross section. If the platform or base of the core was removed, flakes would exhibit similar scalloping. 271 Figure 12.21 Flaked glass from Entrance Platforms at Mulka's Cave. Note the small negative flake scars concentrated along the right distal margin, while the other remains sharp and intact. Stippling represents unmodified surface, black bar = 1 cm. 273 Figure 14.1 Technological provisioning systems (top) and occupation intensity (bottom) at Anderson Rocks, based on dated archaeological content from AR2-3 and surface material; the latter was presumed to date from the last 50 years or so. Where provisioning systems could not be confidently identified, these have been omitted. 301 xxxi Figure 14.2 Technological provisioning systems (top) and occupation intensity (bottom) at Mulka’s Cave. Provisioning systems are shown for all dated pits and the combined surface assemblages, which were presumed to date from the last 50 years. CS = Column Sample. Where provisioning systems could not be confidently identified, these have been omitted. Occupation intensity is based on assemblages from the Camping Area. 308 Figure 15.1 Occupation periods represented by subsurface archaeological deposits at Anderson Rocks (dark blue), Gibb Rock (red) and Mulka's Cave (light blue). CS = Column Sample, P88 = 1988 pit excavated by Bowdler et al. (1989). GR1 was not dated but based on the short depth of deposit the assemblage is probably no more than a few hundred years old. Note the presence of cultural material throughout the occupation periods, indicating no significant hiatuses in site visitation. 322 Figure 15.2 Occupation intensity (top) for each site and effective rainfall conditions for dated sites (middle) and the study area as a whole (bottom). Note the Gibb Rock assemblages were presumed to date to the last few hundred years, and all surface assemblages to the last 50 years. 334 xxxii CHAPTER 1 INTRODUCTION: ADDRESSING THE SCARCITY OF ARCHAEOLOGICAL DATA FROM INLAND SOUTHWESTERN AUSTRALIA 1.1 INTRODUCTION This chapter provides the background to the author's research into Aboriginal occupation of inland southwestern Australia. To begin, the archaeology of the Noongar country, southwestern Australia, is considered, with specific reference to the research bias that has focussed on coastal areas and neglected the inland environs. The archaeological and ethnohistorical data sources are then summarised for the coastal southwest and the Western Desert. These provide a stark contrast to the limited information available for the inland southwest but cannot be applied due to considerably different environmental conditions. The study area is then defined, with reference to Tindale's (1974) boundary that separates the Noongar from the desert groups, as well as a distinct natural boundary that it replicates. The major research aim is outlined, and four specific research questions are devised to address it. The methodology forms a two-pronged approach, whereby a theoretical occupation model based on resource distribution is integrated with archaeological data. The latter includes that collected by the author and the fragmentary remains documented in the wider area. The methods associated with each approach are outlined, and the limitations discussed. The significance of this research is then evaluated, both in the results it aims to provide, and the novel approach employed. Finally, the organisation of the thesis is outlined, with reference to the four distinct sections; each address a particular topic throughout several chapters. 1.2 ARCHAEOLOGICAL RESEARCH BIAS IN THE SOUTHWEST When compared to other parts of the continent, Western Australia is not particularly well represented in the national database of dated archaeological sites. However, Noongar country – defined here as the area covered by the Single Noongar Native Title Claim – preserves a particularly problematic record. While sites are present throughout southwestern Australia, most excavated and dated sites are located within 100 km of the present coastline (Figure 1.1). The area yields some of the 1 oldest sites in Australia – Devil's Lair and Upper Swan both preserve occupation sequences dating to around 45 ka, and the former was arguably visited from 50 ka (Allen and O'Connell 2014; Balme 2014; Pearce and Barbetti 1981; Turney et al. 2001). All other sites post-date 22,000 cal. BP, and many show occupation throughout the Last Glacial Maximum, 21 ± 3 ka (Dortch 1996; Reeves et al. 2013; Williams and Smith 2013). Various site types are represented, including rockshelters, quarries, open sites, rock art sites, and shell middens. While the specifics of these sites are not relevant here, the diversity of this archaeological sample is noteworthy. The ethnohistorical record is similarly varied, unsurprising since early European activity was focussed on the coast, particularly around Perth and Albany. There is an extensive collection of letters, diaries and field notebooks that contain detailed observations of Aboriginal people in the 19th century (e.g. Grey 1841; Hammond 1933; Hassell 1975; Hercock 2014; Moore 1884 – see list in Haebich and Tilbrook 1981). This substantial body of archaeological and ethnohistorical data has provided a wealth of knowledge on various facets of Aboriginal life, including occupation patterns (e.g. Anderson 1984; Bird 1985; Smith 1993), land and resource management practices (e.g. Hallam 1975, 1989; Lullfitz et al. 2017), and food sources (e.g. Meagher 1974). The inland southwest presents a stark contrast to the rich data sources described above. Only one archaeological site has been dated – Mulka's Cave, a decorated rockshelter that was the focus of the author's MA research (Rossi 2010, 2014a). While Rossi (2014a) demonstrated that initial occupation occurred around 8000 cal. BP, the stone artefact assemblage was of little interpretive value due to the lack of comparative archaeological data. Ethnohistorical data was also scarce, so the author was unable to form any strong conclusions about how the site was used, or how it fit into the broader context of southwestern Australia. 2 Figure 1.1 Map of dated sites in Australia, after Williams et al. (2014: Figure 1). The Holocene sites pre-date 10,000 BP. Black box encases all sites in Noongar country; the specific boundary of the Single Noongar Native Title Claim is shown in Figure 1.2. Several archaeological locations are known in the area surrounding Mulka's Cave, but most were identified during cultural heritage surveys for development works, so limited data were recorded and made available. Bird (1985) and Smith (1993) both considered some inland areas in their large-scale analyses, but the data sources were optimised for their individual research questions, so were of little comparative value. While several ethnohistorical sources referred to the Aboriginal group that lived around Hyden (e.g. Davidson 1938, Tindale 1974), other observations were limited to a few brief statements by Roe (1836, 1852), Landor and Lefroy (1843) and Hammond (1933), from which little information could be drawn. This problem persists today and typifies the scarcity of archaeological and ethnohistorical data for the entire inland southwest. This same research bias is not evident elsewhere in the state – there is an extensive suite of sites in the Western Desert, which also stretches into the Northern Territory and South Australia. As in the coastal southwest, site types are varied, and range from rockshelters and rock art sites to ochre mines and stone arrangements 3 (Williams et al. 2008). Several Pleistocene sites are present, including Karnatukul (formerly Serpent's Glen), which was occupied from 47,830 cal. BP (McDonald et al. 2018); several other sites date to around 35,000 cal. BP (Smith 2013:79). The ethnohistorical record is also fairly rich, but for a different reason to the coastal southwest. Due to the remote location, European settlement occurred much later, and the process of cultural dislocation proceeded much more slowly (Haebich and Tilbrook 1981:2). Therefore, by the time Tonkinson and Gould began their fieldwork in the 1960s, they were able to observe many facets of traditional life that had been lost elsewhere. Both published various accounts and addressed several topics, including mobility and occupation patterns, subsistence behaviour, and aggregation events (e.g. Gould 1968, 1969a, 1969b, 1971a, 1971b, 1984, 1991; Tonkinson 1978). While both the coastal southwest and the Western Desert yield a relative abundance of archaeological and ethnohistorical data, it is of limited relevance to the inland southwest. The coastal regions receive more rainfall, and preserve freshwater lakes, rivers and potable groundwater; rainfall is more limited in the inland southwest, and more variable year-to-year (Bureau of Meteorology data). Other differences in the climate, hydrology and geology of each area has created distinct vegetation communities, which harbour different faunal assemblages. Therefore, there is no reason to expect a similar antiquity of occupation, nor that demographic or mobility patterns would be comparable between the coastal and inland southwest. Similarly, the Western Desert data cannot be applied. All Pleistocene sites were associated with a reliable water source of some variety, while none are found in the inland southwest (Bird et al. 2016). Further, in the Western Desert, rainfall is erratic, both within and between years, and can be extremely localised (Gould 1991; Tonkinson 1978:7). As a result, mobility patterns would have been vastly different than in the inland southwest where rainfall is highly seasonal. In contrast to the coastal southwest and Western Desert, then, little is known about the antiquity or nature of Aboriginal occupation of the inland southwest. It is this extensive knowledge gap that the author wishes to address with the current research. 4 1.3 STUDY AREA, AIMS AND RESEARCH QUESTIONS While the research gap identified above affects most of inland southwestern Australia, it was necessary to define a more manageable-sized area for this research to focus on. The position of the study area was primarily dictated by Mulka's Cave, the only excavated and dated site in the area. A targeted sample of granite outcrops identified three other sites with archaeological potential. Therefore, the study area was defined as a square measuring approx. 100 x 100 km, centred on the point between the sites (Figure 1.2); the area east of the State Barrier Fence (formerly No. 1 Rabbit Proof Fence) was excluded, as it broadly coincides with a significant cultural and natural boundary. Near the Hyden township, the fence-line closely resembles the position of Tindale's (1974) frontier that divides circumcising/subincising desert groups from the Noongar, who did neither; this boundary defines the eastern extent of the Single Noongar Native Title Claim. While the circumcision boundary was not solid, nor necessarily static (Gibbs and Veth 2002), it is reasonable to assume that, as Noongar, the study area occupants may not have had regular access to the land further east. The State Barrier Fence also separates the Eastern Goldfields region from other parts of southwestern Australia. At a broad scale, the geology, drainage and climate of the study area has more in common with the Eastern Goldfields than the areas to the west (Anand and Butt 2010 – see Chapter 3.2). At a finer scale, however, significant differences emerge. Greenstone belts (bands of metavolcanic and metasedimentary rock) are common in the Eastern Goldfields, while in the study area basement rock is almost exclusively granite. These geological variations create distinct soil profiles, each of which have different water retention properties; this strongly influences vegetation formations, which in turn affects faunal assemblages. These differences were significant enough that the Eastern Goldfields was separated from southwestern Australia for the purposes of biodiversity surveys (e.g. Biological Surveys Committee 1984). Considering the importance of soil, flora and fauna in the resource and occupation model (see Chapter 1.4), the author chose to exclude the area east of the State Barrier Fence. 5 Figure 1.2 Left: Location of the Single Noongar Native Title Claim (grey) and the study area boundary (blue). Right: Enlarged view of the study area, showing the location of archaeological sites mentioned in the text. The eastern boundary follows the State Barrier Fence. Base imagery from Google earth. The main aim of this research is to address the knowledge gap for inland southwestern Australia by producing an occupation history for the study area. Specifically, the main aim is to answer one particular question: when and how did Aboriginal people use the study area, and did the nature and/or intensity change over time? This broader aim will be achieved by considering four specific research questions: 1. How were food and water distributed across the study area and, following Optimal Foraging Theory, how would seasonal and interannual resource availability influence human occupation patterns? 2. What is the timescale of Aboriginal occupation in the study area, and how does this compare to other parts of Australia? 3. What do the archaeological remains reveal about mobility and occupation patterns within the study area, and how do these results fit within the theoretical model? 4. How did occupation intensity vary over time, and can this variation be explained with reference to long-term shifts in water availability? 6 The methods of answering these questions, and their limitations, are discussed below. 1.4 METHOD OF ADDRESSING RESEARCH QUESTIONS While no human society should be viewed through a framework of environmental determinism, there is no doubt that the natural environment and landscape places severe limitations on how humans can utilise an area. In the context of Indigenous occupation of Australia, the most important factors were the availability of water, food plants and animals. Therefore, this research employs a two-pronged approach. The first requires construction of a theoretical occupation model based on critical resource distribution, while the second involves archaeological investigations. These complimentary data sources will be merged to provide an occupation history for the study area. The methods for each stage of the approach, and the associated limitations, are discussed briefly below; more detailed information is available in Chapters 5 and 9. 1.4.1 Resource and Occupation Model Before the occupation model can be constructed, it is necessary to quantify food and water distribution throughout study area, allowing for variation over space and time. To address spatial variation, several Landscape Divisions will be defined, based on dominant landform and vegetation associations in the study area as recorded by Beard (1972, 1979, 1980); food and water resources will be quantified for each Division. Lists of native flora and fauna will be compiled from the biodiversity surveys conducted on various nature reserves in the study area; vegetation and landform descriptions will be used to identify the relevant Landscape Division/s represented therein. Plant foods will be identified using published lists (e.g. Bindon 1996; Bird and Beeck 1988; Meagher 1974), allowing for unlisted specimens that are known to be edible; all animals will be considered edible, and any associated eggs. Monthly availability of each food item will then be defined using the relevant biological data – this will provide a monthly record of every plant and animal food item available in each Landscape Division. 7 The study area has saline groundwater and no freshwater surface drainage, so the only potable water is rainwater, temporarily preserved in soils and rock structures. As a result, it is important to consider temporal variation not only within a single year, but between years. Therefore, water availability will be considered during low, average and high winter rainfall years, defined with reference to modern rainfall data from Hyden (BoM station no. 010568). Soil water will be evaluated using the Soil Water Tool (Department of Primary Industries and Regional Development 2016a), which was designed to model the water-holding period of various agricultural soils under different conditions, using local rainfall data. A representative soil profile will be chosen for each Landscape Division, based on generalised soil characteristics outlined in the literature (Beard 1979, 1990; Department of Primary Industries and Regional Development 2016b; Murphy-White 2007; Schoknecht and Pathan 2013). Rock structures (pans, gnammas) will be modelled manually, using local rainfall data in conjunction with evaporation and runoff coefficients; this will produce a daily water balance for different structures under various rainfall conditions. The distribution of critical resources undoubtedly places constraints on when and how intensively people can occupy certain parts of the landscape. Considering this, Optimal Foraging Theory (OFT) provides a number of rules that quantify the optimal response to varying resource distributions. The OFT family of models considers which resources should be targeted first, when diets should diversify, and which resources can be profitably transported longer distances (Macarthur and Pianka 1966; Orians and Pearson 1979). Similarly, they consider which foraging patches (represented here by Landscape Divisions) should be searched first, and when groups should move to a new patch (Charnov 1976; Macarthur and Pianka 1966). To use these rules, plant foods will be ranked by their return rates (cal./hr), based on published data. Animal foods will not be ranked as the calculations are more complex for mobile prey, and the data are unavailable. Indeed, fixed resources are more likely to dictate occupation patterns (Bird 1985:131, 140). The spatio-temporal availability of differently ranked resources, as well as unranked resources and water, will be considered in light of the OFT tenets, to determine the optimal seasonal occupation pattern for the study area. 8 Method-specific limitations are evaluated in Chapter 5, but it is important to identify the limitations and assumptions that impact the resource and occupation model at a broader scale. These are listed below: 1. Landscape Divisions are presumed to have consistent water-holding properties and plant/animal foods resources throughout the study area. However, it is conceivable that this may vary at a smaller scale, and that each iteration of a particular Division may differ in their resource base; 2. The effect of annual rainfall on plant and animal foods cannot be considered, as the relevant data are not available. However, foods were unlikely to have been equally abundant under all rainfall conditions; 3. The model will consider the distribution of critical resources as the sole factor dictating occupation patterns. Shelter requirements are not considered, since thick vegetation, rockshelters or hollowed boulders provided natural protection from the elements. Further, shelters could be constructed from bark, branches and other plant matter – Roe (1836) noted the remnants of many such structures in the inland southwest; 4. Human landscape management/modification is not considered. The most widespread technique was firing, which can be used to increase access to prey and to promote plant growth (Bird et al. 2005, 2008, 2013; Gould 1971b; Hallam 1975, 1989; Jones 1969). However, the inland southwest was fired less frequently than coastal regions, and the vegetation communities characteristic of the study area were infrequently burnt when encountered in other parts of the southwest with more intensive firing regimes (Abbott 2003; Lullfitz et al. 2017; Prober et al. 2016). Therefore, fire regimes are unlikely to have significantly impacted the density or distribution of resources in the study area; 5. Population density and territory size will not be considered, as the data are lacking. The study area is assumed to represent a distinct territory occupied by a single group, and the optimal seasonal route is defined with reference to this area only. It is somewhat misleading to consider an area in isolation, since the group may have had access to some Noongar land outside the boundary. Similarly, the study area may have supported more than one group, each with distinct portions of land. 9 1.4.2 Archaeological Investigations This study is distinct from a regional analysis (e.g. Attenbrow 2004; McBryde 1974) as it must necessarily omit a defining feature of these: extensive site survey of the study area or a representative sample of it (McBryde 1986; White 2011). This cannot be achieved for the study area, since most of the land has been cleared for agriculture and is privately owned. Known archaeological site locations are therefore skewed towards those areas of the landscape that are less affected by such factors – granite outcrops and salt lakes. Nevertheless, it is still important to consider the archaeological record from a number of sites to construct a meaningful occupation history, as: Any sequence based on the evidence of one site, however rich the site, may be unreliable, due to local specialized features of that site not readily recognized on first investigation. Variations may be found from site to site, not only in the quantity of material left behind by their prehistoric occupants, but also in the types of implements represented. Assemblages from sites of approximately the same geographical area may differ in their components...It is only on the basis of excavated, stratified, assemblages from a number of sites within the same region that a sound reconstruction of the cultural sequence for that region’s prehistory may be made’ (McBryde 1974:16–17). Therefore, this research will consider archaeological remains from several sites. Those from Anderson Rocks, Gibb Rock and Mulka's Cave will be considered in detail, as these granite outcrops were the focus of the author's excavations and surface collections (Figure 1.2). Another outcrop, Twine Reserve, will be considered more briefly, for reasons outlined in Chapter 9.2.1; discussion will be limited to observations made by the author during a site visit, and their relevance within in the context of the occupation model. Where sufficient details are available, other archaeological remains from the study area will be considered in conjunction with the model and the author's data. Stone artefacts from Anderson Rocks, Gibb Rock and Mulka's Cave are classified using a materialist approach, based on Hiscock (2007). In contrast to typological systems that rely on artefact form, and often inferred function, a materialist approach is less subjective as it classifies artefacts based on their position within the reduction sequence. This places more importance on artefact attributes than the frequency of different artefact types. To address questions of mobility and occupation patterns, 10 assemblage data will be considered with references to Kuhn's (1995) provisioning systems. The strength of this framework lies in the integration of mobility, raw material availability and risk into a single model – one that also follows the tenets of ecological/forager theory. Kuhn (1995:20) accepts that people move to access critical resources (food and water) but, following Bamforth (1986), acknowledges that this changes spatial relationship between people and stone sources, which are rarely truly ubiquitous. Kuhn (1995) proposed two contrasting solutions to ensure that technological needs were consistently met: Individual Provisioning and Place Provisioning. Individual Provisioning arises under increased residential mobility, when unpredictable access to replacement stone is managed by the manufacture of a portable, multifunctional toolkit. This kit is heavily maintained and recycled; artefacts are only discarded when they cannot be repaired or recycled. In contrast, Place Provisioning involves a particular site being furnished with toolmaking potential, in the form of raw material stockpiles. Artefacts are produced in a more expedient manner, and discarded after use; when another task arises, new flakes are struck. This strategy is only optimal when residential mobility is decreased, and the duration and predictability of occupation increases. Each strategy produces distinct assemblages due to differences in raw material requirements, discard behaviour and reduction sequences. Occupation intensity is considered with reference to these provisioning systems, as well as other commonly employed proxies such as reduction intensity, artefact discard rates, and post-depositional breakage. The main limitations in the archaeological investigations relate to the sample, specifically the number of sites investigated, the size and distribution of pits, and how representative these samples are at a broader scale; the issue is not unique to this research, as all archaeological excavation involves some level of sampling. Three sites were targeted for more detailed investigation. While an occupation history would ideally derive from a larger sample, this is partially addressed by investigating different areas of each site. This increases the number of comparative assemblages available and allows the antiquity of occupation to be defined based on several different dated deposits. However, the author's pits are small, 500 x 500 mm, to permit different site areas to be investigated while still remaining within the one 11 square metre excavation limit imposed by Section 16 of the Aboriginal Heritage Act (1972). Small pits are more susceptible to sampling biases, particularly in terms of less frequently encountered artefact types (e.g. retouched pieces). It is therefore necessary to consider the possibility that such pieces were not truly absent, but simply outside the pit boundary. Nevertheless, Hiscock's (2007) and Kuhn's (1995) approaches both rely more heavily on artefact attributes than types, so small pit size should have a minor impact. It is probably more influential in terms of yielding smaller overall assemblages, in which certain attribute patterns may be harder to identify. 1.5 SIGNIFICANCE OF THIS RESEARCH This study is significant as it aims to address the knowledge gap for inland southwestern Australia, typified by the lack of excavated and dated sites compared to the other parts of the country (Williams et al. 2014). At present, there are insufficient archaeological or ethnohistorical data to indicate when and how the study area was used by Aboriginal people; these questions will be directly addressed by this research. The study area may be of additional interest due to its location in something of an intermediate area climatically, environmentally and culturally – it lies between the well-watered coastal southwest and the arid interior, and adjoins the boundary separating southwestern Australia from the Eastern Goldfields, and the Noongar from desert groups. As a result, the study area should preserve its own unique occupation history, unlike those from the coastal southwest or Western Desert, where data are more abundant. The occupation history will allow the Mulka's Cave assemblage to be interpreted, so the site can be placed in its regional context. This is particularly important since it is the most profusely decorated Noongar rock art site, and receives more than 80,000 visitors annually, partly due to its vicinity to Wave Rock (Gunn 2004, 2006). The theoretical occupation model will allow fragmentary archaeological remains or particular site types (e.g. granite outcrops) to be interpreted, and predict the type of occupation sequences that might be found in different parts of the study area. These results may not be applicable to the entire inland southwest, but will certainly be of some broader relevance beyond all but the eastern boundary of the study area. 12 While the importance of increasing the body of archaeological data for the inland southwest cannot be overstated, the significance of this research is also in the approach itself, particularly with regard to quantifying resource distribution. This is not in itself a novel idea, but the scale and integration of analysis is noteworthy. Here, the theoretical occupation model and archaeological assemblages are treated as complementary data sources, with neither taking precedence over the other. Too often, resource data are considered supplementary to archaeological findings, and provide background or contextual information rather than being fully integrated into analyses (e.g. Bird 1985; Smith 1993). Plant and animal food lists are generally brief, incomplete, and compiled for large regions; as a result, these underestimate resource diversity and spatial variability. In contrast, the author aims to compile an extensive list of all plant and animal foods likely present within each Landscape Division, and to define their availability on a monthly basis. Further, the focus is on organism-specific data that dictates when an edible product was available, rather than ethnohistorical records detailing the season that particular food sources were used. The latter may be unreliable where items were available for longer periods, or year-round, since season of use will be partially dictated by the relative abundance of other food products. This method will provide a more complete, less subjective record of food resources in the study area, while also addressing spatio-temporal variation at a finer scale than previously attempted in southwestern Australia. As with food resources, water availability is often considered more broadly, by simply noting the types of sources that may have been used during certain seasons (e.g. Bird 1985:176). Ephemeral water sources are rarely considered in detail, probably because they were less integral to human survival, as most areas provide potable groundwater, or freshwater lakes or swamps. In the study area, however, this is not the case. As a result, new methods are devised to quantify the longevity of rainwater stored in soil profiles and rock structures under low, average and high rainfall conditions. These data will provide a high-resolution representation of how water availability varied over space and time, and indicate how this would have constrained human occupation. The author is not aware of any previous attempts to quantify ephemeral water sources at all, let alone in this manner; it may be of use in other areas where water is similarly limited. 13 1.6 ORGANISATION OF THESIS Section One (Chapters 1–4) provides the background and contextual information of relevance to this research. Chapter 1 has outlined the research bias in the Noongar country, southwestern Australia, which has resulted in little knowledge on when or how the inland environs were used. It detailed the research questions formulated, and the methods devised to address this knowledge gap, and the significance of doing so. Chapter 2 details the theoretical background, focussing on forager theory and technological organisation, which are united through Kuhn's (1995) technological provisioning systems. Chapter 3 focusses on the natural environment of the study area. The geology and hydrology are described, noting the impact of clearing on hydrological conditions. The Quaternary and modern climate is evaluated, with particular focus on the conflicting nature of the Holocene palaeoenvironmental records for southwestern Australia. Finally, the vegetation systems and fauna distribution are described. Chapter 4 describes the cultural context of the study area and beyond. The earliest archaeological evidence in Australia is briefly evaluated, followed by a discussion of the archaeological locations in the study area, and the few relevant studies from further afield. The fragmentary ethnohistorical record is then evaluated. Section Two (Chapters 5–8) describes the various facets of the resource and occupation model. Chapter 5 details the methods used to quantify resource availability and formulate the model. Six Landscape Divisions are defined to address spatial variation within the study area. Discussion then turns methods of quantifying food and water availability in each – while the former relies on biodiversity surveys and published data, the latter involves some more unique methods, including the use of an agricultural model. The means of converting these data into an occupation model are then outlined. Chapter 6–7 provide the descriptive results; the former outlines the spatio-temporal availability of plant and animal foods, while the latter details the longevity of ephemeral water sources under low, average and high rainfall conditions. Chapter 8 presents the occupation model, based on these data. It defines the optimal seasonal route that the study area occupants may have followed, and notes which technological provisioning systems should characterise certain sites or occupation periods. It also considers whether aggregation events could have occurred in the study area, and where and when these might have been held. 14 Section Three (Chapters 9–13) focusses on the archaeological investigations. Chapter 9 outlines the methodology, beginning with field methods, before outlining the materialist approach used to analyse the quartz-dominated assemblage, and the reasons for its selection. Artefact attributes are then discussed, and archaeological indicators formulated for each technological provisioning system, as well as different levels of occupation intensity. Chapters 10–12 focus on each of the excavated sites – Anderson Rocks (Chapter 10), Gibb Rock (Chapter 11) and Mulka's Cave (Chapter 12). The sites are described, and their specific water resources quantified. The organics and any radiocarbon dates are discussed, before the stone artefact assemblages, and the attributes collected from them, are detailed. Other cultural material, such as ochre or flaked glass, are then briefly considered. Chapter 13 considers the assemblage from each site in Kuhn's (1995) framework, to identify the technological provisioning systems represented within each pit or surface collection, allowing for temporal shifts within the former. Section Four (Chapters 14–15) unites the theoretical occupation model with the archaeological data. Chapter 14 considers how Anderson Rocks, Gibb Rock and Mulka's Cave were used, with reference to site function and overall mobility patterns. Where possible, changing intensity over time is then considered, and linked to shifts in water availability. Finally, Chapter 15 summarises the research to provide an occupation history for the study area. Each of the four main research questions is answered and the broader implications of those results considered. Future research directions are then identified, before the concluding remarks are made. 15 CHAPTER 2 THEORETICAL FRAMEWORK: MOBILITY, SUBSISTENCE AND TECHNOLOGY 2.1 INTRODUCTION This chapter discusses the theoretical framework of this study, outlining how occupation patterns are influenced by ecological conditions, and how such patterns may be detected within the archaeological record. Forager models are briefly discussed, as they provide solid predictions about how people should move around the landscape and alter their subsistence strategies and settlement patterns to optimise returns under varying resource distribution. Technological organisation is then evaluated, focussing on how behavioural and physical attributes of technology can be affected by mobility, raw material availability, and risk. Finally, Kuhn’s (1995) provisioning systems are described; this concept unites forager theory with technology by predicting the different ways that people can ensure their technological needs are met under various conditions. 2.2 FORAGER THEORY Forager Theory falls under the umbrella of Human Behavioural Ecology (HBE), a field that explains human behavioural diversity using models grounded in evolutionary ecology (Winterhalder and Smith 2000). While not the sole determining factor, the environment is a major element constraining human occupation – HBE considers how human behaviour can be affected by various environmental conditions (Nettle et al. 2013). A major subset of HBE is Optimal Foraging Theory (OFT), a family of models that relate the resource base to certain behaviours or choices, with the overall goal of optimisation (utility increase), i.e. to maximise returns (Kelly 2013:33–34; Mithen 1989). Foraging efficiency is often considered to be a proxy for reproductive fitness so, following evolutionary theory, optimal foraging should be the goal, whether conscious or not (Kelly 2013:34; Smith 1979). There is a vast body of research dedicated to HBE and forager theory, so only relevant contributions are discussed further; more comprehensive treatments can be found in Bettinger et al. (2015), Kelly (2013), Nettle et al. (2013) and Winterhalder (1981). 16 Individual OFT models may identify which available resources should be pursued, the best places for foraging, when resources should be processed before transport, and when groups should move to new foraging grounds. HBE and OFT have their roots in ecological science, so many models were initially developed for animals (e.g. Charnov 1976; Emlen 1966; Horn 1968; MacArthur & Pianka 1966; Orians and Pearson 1979), and subsequently modified for, or applied to, hunter-gatherer societies (e.g. Dyson-Hudson and Smith 1978; Harpending and Davis 1977; Hawkes & O'Connell 1992; Wilmsen 1973). Individual models are often complex mathematical computations (Kelly 2013:46), involving the following: o Goal: generally, to maximise caloric energy return for time spent foraging (including search, handling, transport and processing); o Currency: calories, or energy; o Constraints: e.g. the amount of time that can be dedicated to foraging; o Options: various potential food sources or foraging patches. Applying a foraging model requires intimate knowledge of the landscape and the various resources found therein – their nutritional value, handling time, transport costs and distribution. That such detailed knowledge is required to generate arguably simple models is a common criticism of forager theory (Bettinger et al. 2015:133; Grøn 2012; Smith 1983). Models also do not allow for choices unrelated to energy requirements or costs, for example items selected or omitted due to preference or taboo, but as Kelly (2013:76) noted, these models are not intended to replicate reality, but to model expected outcomes if certain conditions are met. While detailed knowledge is required to apply individual forager models, their underlying rules (Table 2.1) are robust enough to be successfully applied in isolation, or incorporated into broader analyses (e.g. Clarkson 2004, 2007). Ambrose and Lorenz (1990) used OFT models as the foundation of their resource structure scheme. They characterised resources according to their predictability and abundance, creating four extremes of the resource-structure continuum: predictable and dense (e.g. well-watered woodlands), predictable and scarce (e.g. semi-arid woodland, scrub), unpredictable and dense (e.g. grasslands with seasonal herds), unpredictable and scarce (e.g. deserts). Using several OFT models, they characterised residential mobility, settlement patterns, and social organisation for 17 groups inhabiting each distinct resource zone (Figure 2.1; Table 2.2). While Ambrose and Lorenz (1990) merely summarised findings from existing forager models, their data provide a useful synthesis while also permitting OFT models to be more widely applied by relating predictions from individual models to specific resource structures. Table 2.1 Forager models and theories for achieving optimisation, based on the references listed and summary data from Bettinger et al. (2015), Clarkson (2007), Kelly (2013) and Winterhalder (1981). References denote the first iteration of a model or a significant subsequent contribution. Model name /subject Method to achieve optimisation Reference Dietbreadth model Foods are ranked according to returns (energy value for time spent procuring and processing). Higher ranked foods are pursued first, and lower ranked resources are utilised when higher ranked items are unavailable or used to supplement the diet when higher ranked items are less abundant. Macarthur and Pianka 1966 Patch choice model Patches (areas of land that contain resources, which are distributed unequally across landscape) are ranked according to return rates. Patches with higher returns are searched first, moving to lower ranked patches as higher ones are depleted. Macarthur and Pianka 1966 Group size and dispersion Where resources are stable and evenly distributed, and variance in returns is low, foragers should be dispersed in small units; where resources are mobile and clumped (e.g. greater spatial variation in returns), people should be aggregated in a larger group. Residential mobility increases as resources become scarcer and less predictable. Moves will be opportunistic where resources are unpredictable and scheduled where they are predictable. Harpending and Davis 1977; Horn 1968; Wilmsen 1973 Marginal Value Theorem Patches are not exploited to exhaustion. People will move to a fresh patch when the return rate in the current patch reaches the average for all patches, allowing for travel time. Patches are abandoned sooner when the environment is of better quality overall. This model may not hold true when occupation is tethered to certain resources, e.g. a stable water source. Charnov 1976 Territory size / range Range (e.g. territory size) should be greater in poorer environments, where resources are scarcer or patchily distributed. Territories will be smaller where resources are dense and more evenly distributed. Harpending and Davis 1977 Economic Where resources are dense and predictable, territories are small, so Defensibility the cost of defending territory is exceeded by benefits. Elsewhere Model costs are too great, leading to passive territories (e.g. undefended, semi-permeable) or shifting territories/high mobility. Central Place Foraging Model Dyson-Hudson & Smith 1978 Allows for food being transported to a central place rather than being Orians and consumed at the point of extraction. Unprocessed resources should Pearson 1979 be transported to camp until the return rate (including travel time) is greater for processed vs unprocessed foods – processing will become more profitable as transport distance increases. Distance to camp determines what resources are worth collecting, e.g. lowranked foods can't be transported far as costs outweigh returns, so the range of viable resources narrows with distance from camp. 18 Figure 2.1 Optimal and suboptimal strategies for different resource structures (Ambrose and Lorenz 1990:15). Table 2.2 Behavioural characteristics arising from different resource structures. Re-tabulated from Ambrose and Lorenz (1990:10). Resource structure Territorial strategy Information exchange Residential mobility Group size Population density Diet breadth A: predictable and dense Territorial defence Low Low, scheduled Small High Moderate B: predictable and scarce Home range, semipermeable Undefended, very permeable Undefended, very permeable Medium Medium, scheduled Small Medium High Very high High, opportunistic Large Medium Very low High Very high, opportunistic Very small Very low Very high C: unpredictable and dense D: unpredictable and scarce 19 2.3 TECHNOLOGICAL ORGANISATION Technological organisation is 'the study of the selection and integration of strategies for making, using, transporting, and discarding tools and the materials needed for their manufacture and maintenance' (Nelson 1991:57). Previously, technological variation had been attributed to stylistic or functional concerns (see discussion in Rolland and Dibble 1990), but it was later recognised that technology had the power to reveal elements of human behaviour, as technology responds to social and economic factors. Hence, the ‘need to acquire resources in different locations, to move around the landscape, to remain settled at a place, to transport different kinds of resources and material needs, and many other variables condition the technological strategies employed at a particular time and place' (Nelson 1991:88) Technology, then, can be regarded as a problem-solving strategy and a physical manifestation of how people interact with the environment (Kuhn 1995:19). As such, it is vital to integrate technology into theoretical concerns, not only because stone artefacts dominate archaeological assemblages, but to generate a more complete picture of past human behaviour (Torrence 1983, 1989). Several factors influence technological organisation, including specific functional requirements (e.g. Ferris 2015), manufacture costs and resource returns/tool efficiency (e.g. Clarkson et al. 2015; Garvey 2015; Kuhn and Miller 2015), raw material availability (e.g. Andrefsky 1994; Bamforth 1986; Manninen and Knutsson 2014), demographic changes (e.g. Henrich 2004; Shennan 2001), mobility (e.g. Hiscock 1996; Parry and Kelly 1987; Shott 1986) and risk (e.g. Collard et al. 2005, 2013; Hiscock 1994; Torrence 1983, 1989, 2001). The resulting variation can be behavioural – relating to the timing or location of stone-working activities – or physical, relating to the design of individual artefacts, the composition of assemblages, or the reduction techniques used to produce them (e.g. Binford 1979; Hiscock 1994; Nelson 1991; Parry and Kelly 1987; Shott 1986; Torrence 1983, 1989, 2001). A thorough discussion of these contributing factors is well outside the scope of this thesis, so the focus here is centred on three major factors: mobility, raw material availability, and risk. Arguments predicated on assemblage diversity or implement design (e.g. Bleed 1986; Collard et al. 2005, 2013; Shott 1986; Torrence 1983, 1989, 2001) are not considered below, as the principles are difficult to apply to amorphous, debris-dominated assemblages, such as those frequently encountered 20 in Australia (Hiscock and Clarkson 2000; Kuhn 1995:19–20). Instead, the discussion is limited to those case studies yielding results relevant within an Australian context. 2.3.1 Mobility Generally, two types of mobility are recognised: residential and logistical. Residential mobility occurs when an entire group moves to a new location to exploit particular resources, while logistical mobility involves small task groups leaving a central location to procure resources from further afield, before transporting them back for the entire group to consume (Binford 1980; Kelly 2013:78–79). As discussed above, the magnitude of residential mobility is dictated by the abundance and predictability of resources, while increased logistical mobility can arise when residential mobility is impeded, often due to the incongruent or patchy distribution of critical resources – for example where permanent water is present but food is scarce (Binford 1980). Neither residential nor logistical mobility, however, should be treated as a single variable, as each comprises several parameters, including frequency of moves, distance per move and annual distance travelled (Shott 1986). Discussion here focusses on the relationship between residential mobility and stone reduction techniques – specifically standardised versus unstandardised techniques, and the use of bipolar flaking, as these features are easily identified in Australian stone artefact assemblages. Parry and Kelly (1987) relied heavily on Binford's (1973, 1977) concept of curated and expedient technology to define their standardised and unstandardised reduction techniques, but directly relate the use of these techniques to the level of residential mobility. They note that unstandardised (expedient) technologies require minimal time, effort or skill to execute, as no attempt is made to control flake form. Instead, the most suitable piece for a particular task is chosen from the flaking products; pieces are not reworked, conserved or retained after initial use. Unstandardised techniques consume a lot of stone, so raw material selection is dictated by what is locally available in sufficient quantity, though bipolar flaking may be used to conserve material where it is less abundant (Parry and Kelly 1987). In contrast, standardised (curated) technologies require considerable time, energy and skill to execute, as well as good quality raw material that is used in a much more efficient manner. Tools are portable, formal and multifunctional, and are repaired and reused many times. 21 However, standardised technologies are much costlier to manufacture, maintain and even use, since generalised tools are less efficient than specialised tools and retouched edges are blunter than fresh ones (Parry and Kelly 1987; Torrence 1983). In their archaeological example, Parry and Kelly (1987) identified a shift to standardised reduction that could not be related to local conditions – including the availability of raw materials – but instead coincided with the occurrence of large, permanent villages, indicating a shift to a sedentary lifestyle (Parry and Kelly 1987). They argued that for highly mobile groups, the greater cost of standardised technologies is outweighed by the benefits conferred through the portability of items, conservation of raw material (since access may be limited) and the likelihood of most tasks being achievable with a multifunctional toolkit. Where mobility is reduced, the cost of standardised technologies is too high and many of the benefits evident under greater mobility are not required (e.g. portability) – therefore, expedient technologies are produced (Parry and Kelly 1987). These expedient technologies may also be associated with greater mobility where raw material is ubiquitous and abundant, negating the transport of tools between residences (Bamforth 1986; Parry and Kelly 1987). Hiscock (1996) also evaluated the influence of residential mobility on technological organisation, adopting a narrower focus than Parry and Kelly (1987), by linking the level of bipolar reduction to residential mobility. Parry and Kelly (1987) noted that bipolar reduction was evident in their expedient assemblages – which reflected more sedentary groups – but only linked it to situations of reduced raw material availability. However, Hiscock (1996) considered that conservation may arise independent of raw material availability and increase in concert with occupation duration. He studied several assemblages from the South Alligator River area of the Northern Territory, focussing on sites in woodland areas and comparing them to those on the floodplain margin. At woodland sites, bipolar cores became more common as the distance to raw material sources increased, indicating the type of raw material conservation noted elsewhere (Bamforth 1986 – see Chapter 2.3.2). However, all floodplain sites were at, or extremely close to, raw material sources, yet bipolar cores were more than twice as frequent there than at any woodland sites, even those furthest from raw material sources (Hiscock 1996: Figure 1). Therefore, variation in bipolar frequency could not be explained with reference to source location alone. Instead, 22 Hiscock (1996) noted that bipolar knapping often occurs near the end of a reduction sequence, when a core would otherwise be abandoned due to low inertia; it would be beneficial for less mobile groups to extend core use-lives, and thereby reduce raw material procurement and transport costs. Therefore, he concluded that the floodplain sites, characterised by the greater frequency of bipolar cores, represented longer occupations (Hiscock 1996). It may seem unnecessary for groups to conserve stone when sources are nearby, but the benefits are evident when considered in the context of utility increase. Since foragers seek to maximise returns while minimising time and energy expenditure, there would be little benefit to incurring the costs associated with stone procurement – minimal as they may be – when existing supplies could be further reduced using a different technique, one that requires little time or skill to execute and produces more cutting edge per core than freehand flaking (Pargeter and de la Pena 2017; Pargeter and Elen 2017; Parry and Kelly 1987). 2.3.2 Raw Material Availability Binford's (1979) concept of embedded procurement arose from his fieldwork in Alaska and Central Australia, where he observed that people rarely made trips specifically to procure stone. Instead, raw material collection was embedded within subsistence tasks – people brought home stone when foraging or hunting returns were low to maximise transport potential. Binford (1979) argued that embedded procurement could drastically reduce, and even entirely void, direct procurement costs. As stone collection occurred during regular food gathering activities, Binford (1979) posited that raw materials used for tool manufacture are, therefore, a direct reflection of the foraging range accessed from a particular site. Garvey (2015) noted that the widespread acceptance of Binford's (1979) concept places more importance on the means of conserving stone, once procured, than on wholly evaluating the mechanism of procurement itself. Regardless of whether material is procured directly or via an embedded strategy, access to raw material can impact technological organisation independent of other factors. Here, the focus is on the distribution of raw material sources and the physical properties of the stone itself. 23 The distribution of raw materials can influence whether they are included within the technological system and, if they are, how intensively they are reduced. Bamforth (1986) found that where a single raw material source was utilised, periods of greater residential mobility encouraged higher levels of recycling and maintenance. He argued, however, that this was not because mobility inherently required the use of a more formal toolkit, but because changing mobility restricted access to the stone source, and therefore encouraged conservation; conversely, when access improved, there was less need to recycle and maintain tools. Where multiple sources existed – at varying distances from an occupation site – Bamforth (1986) found that local, lower quality materials were retouched less and were frequently discarded before they were broken, while higher quality material from more distant sources was subjected to increased levels of maintenance and recycling. In addition, pieces were more likely to be broken before being discarded. If this high-quality material was collected during subsistence activities (embedded procurement) it may reflect a portion of territory that was visited long ago, or where the next visitation date is uncertain (Clarkson 2007:25). However, reduction intensity may also be related to transport and procurement costs if the material was targeted directly. More distant materials incur greater procurement and transport costs, so the resultant artefacts must be used for longer periods (and therefore retouched, maintained and recycled) to offset costs and confer some benefit over using lower quality, more local materials (Garvey 2015; Kuhn and Miller 2015). Where shifts in mobility bring people closer to a higher quality source, however, procurement costs are reduced. This may make it profitable to use the higher quality material or, where it was already in use, permit less intensive reduction and more frequent artefact replacement (Garvey 2015). It is also important to consider the physical properties of various raw materials, particularly those that are local and abundant, as they are more likely to dictate technological decisions. Not all raw materials are equally suited to artefact manufacture – higher quality materials are generally more homogenous and isotropic, so they not only flake more predictably than lower quality stone but are more suited to reworking and retouching, and may retain their sharp edges for longer (Bleed 1986; Cotterell and Kamminga 1979; Whittaker 1994:65–66). 24 Andrefsky (1994) devised a framework for predicting how the abundance and quality of local stone impacts technological organisation, specifically the use of formal or informal tool production (Figure 2.2). He found that where local stone was abundant it dominated assemblages, regardless of the quality of the material; however, the technological system adopted varied in response to the quality of the stone. Where it was abundant, high-quality material was reduced both formally and informally, though low-quality stone was always associated with informal reduction, often supplemented with small quantities of formal tools made of higher quality material (Andrefsky 1994). These results were echoed by Manninen and Knutsson (2014), who identified an extreme case whereby local stone was so abundant but of such a poor quality that the entire technological system was simplified to incorporate it, rather than rationing high-quality stone as expected (Andrefsky 1994; Bamforth 1986). Andrefsky (1994) found similar levels of formal and informal reduction in both short- and long-term occupations (i.e. higher and lower residential mobility), where the raw material source was constant. Elsewhere, he found evidence for informal technology associated with higher levels of residential mobility since, as others had argued, this strategy could be adopted where raw material availability permitted (Bamforth 1986; Parry and Kelly 1987). Similarly, Manninen and Knutsson (2014) identified that their observed switch to informal technology was associated with increased mobility, which took people further from the source of good-quality raw material. These results appear to address Bamforth's (1986) question of why people would make and transport formal tools if material was abundant by answering, ‘they didn’t’. 25 Figure 2.2 The relationship between quality and abundance of local stone, and the use of formal and informal tool production (Andrefsky 1994:30). 2.3.3 Risk In the context of subsistence, risk relates to the probability of failing to procure resources, or not meeting dietary requirements (Hiscock 1994; Torrence 1989). Risk can vary in timing and severity (the cost of failure), which can be relatively minor, whereby additional time or energy costs are incurred, or severe enough that it endangers the lives of individuals. Risk can be caused by several factors, including uncertainty of resource distribution (when people move into unfamiliar areas), climatic shifts, resource depletion/scarcity/mobility, and the limited spatio-temporal availability of prey (Hiscock 1994; Torrence 1989). Torrence (1989, 2001) argued that risk is greatest where diet breadth is lowest, especially where people rely on a few species of large game that are highly mobile and only available during a certain season. While this particular scenario is not relevant to Australian research, other risk factors would certainly have been influential. To a certain extent, risk can be minimised by utility increase, i.e. altering diet-breadth, mobility and aggregation patterns to emulate the optimal solution for local conditions (see Chapter 2.2; Clarkson 2007:12–14; Gould 1991; Hiscock 1994). However, technology can also 26 reduce risk through both behavioural and physical characteristics – technological activities can be scheduled to reduce time/energy costs and ensure that tools are ready for use when required, and risk can also be mitigated by the physical attributes of artefacts themselves; these strategies are discussed in more detail below. Initially an independent concept, Torrence's (1983) time-budgeting model was later subsumed into her broader risk reduction concept (Torrence 1989, 2001). Her central idea was that, to reduce risk, people must schedule technological tasks so they do not interfere (either through time or energy consumption) with more important subsistence tasks, while simultaneously ensuring that tools are ready for use. These technological tasks can incorporate raw material procurement as well as artefact manufacture, maintenance and repair. She assumed that raw materials were collected via embedded procurement, since Binford (1979) demonstrated that this was the lowest cost method of procurement, but this requires that usable stone is found within the foraging range. Manufacture and maintenance tasks should be performed during down-time, either after dark or during other activities, for example while waiting for game (Binford 1977, 1979). Where risk is more frequent or the cost of failure is greater, manufacture and maintenance should be anticipatory to ensure that tools are ready before they are required; when risk is more limited, tools can be made or repaired as they are needed (Torrence 1989, 2001). As Clarkson (2007:15) summarised, the archaeological implications for risk reduction should be that raw materials reflect foraging range, and that more complex assemblages – representing manufacture, maintenance and discard – would occur at residential sites, where down-time is predictable and abundant. As noted earlier, design elements can be difficult to observe in Australian lithic assemblages that are often dominated by debris and other amorphous material. However, Hiscock (1994, 2002a) related the mid-Holocene introduction of the Australian Small Tool Tradition (ASTT) – identified by its small, formal, highly standardised implements – to risk reduction required by resource uncertainty that arose due to climatic shifts, possibly related to El Niño Southern Oscillation (ENSO), and/or people moving into unfamiliar parts of the landscape. Despite having generalised knowledge of the wider area, for example the types of plant and animal foods and stone that may be available, people would be unfamiliar with local 27 resource distribution, so would have trouble predicting when artefacts may be required and when/where raw materials may be encountered (Hiscock 1994). Torrence (1989, 2001) noted that well-organised technology is required to successfully reduce risk. ASTT implements certainly conform to this definition and provide clear advantages to people wishing to reduce risk. The small, standardised tool forms such as points, backed artefacts and tulas – each of which has a distinct geographical distribution (Hiscock 1994: Figure 2) – have long been argued to be the only surviving components of composite tools used for a variety of tasks (Fullagar et al. 2009; Kamminga 1980; Robertson 2009; Robertson et al. 2009). These tools were portable, multifunctional, reliable, maintainable, and had redundant features built in to guard against component failure, making them advantageous under risky conditions that may promote increased mobility (Bleed 1986; Hiscock 1994, 2002a; Shott 1986; Torrence 1983, 1989, 2001). Standardisation reduces repair time, conserves raw material and normalises performance, desirable characteristics where risk is heightened (Clarkson 2007:18–19; Hiscock 2006). Some regions exhibit a decline in these formal tools during the late Holocene – Hiscock (1994, 2002a) attributed this to less risky conditions, as effective precipitation had possibly increased, and resource structures were well-known. Groups were possibly less mobile, so portable, multifunctional toolkits were no longer required, although some formal elements were retained where they were functionally necessary (Hiscock 1994). 2.3.4 Summary The discussion above indicates that several factors influence technological organisation, and that their effects can be manifested in various ways. Further, the same technological strategy can arise under different conditions. For example, expedient technologies may arise as a result of reduced mobility, abundant raw material, or low-risk conditions. Similarly, each of these factors is interrelated – mobility shifts may occur as a response to risk and subsequently alter access to raw material sources. In each case study, the relevant author identified the factor that they believed exerted the strongest influence over technological organisation when related to their own dataset, but it is important to acknowledge that the dominant influence may differ according to the unique characteristics of a site or area, i.e. 28 there is no single factor – whether it be mobility, raw material availability, or risk – that will override all others in every scenario. It is therefore necessary to consider the local conditions, including raw material distribution, mobility and subsistence strategies, as well as risks that may arise through environmental change or altered settlement patterns. All of these factors have the potential to influence technological organisation, but the dominant mechanisms, as well as expressions of variation, will differ over space and time. 2.4 TECHNOLOGICAL PROVISIONING SYSTEMS Kuhn's (1995) provisioning model successfully united the two previously discussed bodies of theory: forager models and technological organisation, incorporating the major themes of mobility, raw material availability, and risk. Following ecological theory, Kuhn accepted that mobility is primarily dictated by the abundance and predictability of critical resources: food and water. Movement alters the location of people relative to raw material sources – which are rarely truly ubiquitous – so people may find themselves in an area with an abundance of food, but no usable stone (Kuhn 1995:20). Therefore, Kuhn saw two ways for people to ensure that they have access to the technology they need, when they need it: Individual Provisioning involves people carrying their technological requirements with them in the form of portable, multi-functional toolkits; in contrast, Place Provisioning involves furnishing a specific location with toolmaking potential (raw material) or finished implements. Following Binford (1977, 1979) and Torrence (1983), Kuhn (1995:25) believed that manufacture and maintenance tasks occur during down-time, which is most predictable and abundant at residential locations. As a result, the provisioning strategy employed should be evident from the archaeological debris generated at these sites. This in turn yields information on mobility, since the scale, frequency and predictability of residential moves dictates the optimal technological provisioning system (Figure 2.3a). The contrasting provisioning systems are discussed in more detail below. 29 Figure 2.3 Impact of mobility on provisioning strategies: A (left): frequency of moves and occupation duration; B (right): length of logistical forays. (Graf 2010, after Kuhn 1995). 2.4.1 Individual Provisioning In situations of high residential mobility, often required where resources are scarce or unpredictable (Ambrose and Lorenz 1990), the opportunity to reprovision with stone may be poorly known or difficult to predict. In those cases, Kuhn (1995) argued that individuals should provision themselves with their technological needs – in the form of finished tools, blanks and small standardised cores – to guard against the risk of being unprepared. However, this preparedness must be weighed against the cost of constantly carrying around one's technological requirements. Transport costs are arguably the overriding factor constraining technology for highly mobile groups, so the toolkit must be portable but also versatile and multifunctional, since specialised tools can rarely be accommodated (Gould 1969a:76; Kuhn 1995; Nelson 1991; Shott 1986). Since future access to raw material sources is uncertain, and mobile toolkits leave little room for spares, the use-life of individual implements should be extended as far as possible (Kuhn 1995). Items should be designed for durability and be reworked and recycled when they become blunt or damaged, and discarded only when they cannot be modified for continued use (Bleed 1986; Binford 1979; Kuhn 1995). Replacements must be manufactured from blanks, or flakes detached from transported cores. The greater use-life of such tools favours the use of a higher quality material, so stone types should be carefully selected, prioritising those materials that flake predictably, retain sharp edges for longer, are more resistant to breakage, and can be easily reworked (Bleed 1986; Garvey 2015; Goodyear 1989; Graf 2010; Mackay 2005; Torrence 1989). 30 Invariably, unanticipated tasks will arise that cannot be completed with existing tools. In these cases, tools must be made by modifying components of the toolkit or, where possible, manufacturing more expedient tools from locally available raw material (Kuhn 1995; Mackay 2005). Clarkson (2007:16) noted that this type of strategy will be especially important under increased residential mobility, since frequent moves constrain the diversity of transported toolkits (Shott 1986). The above data permit a prediction regarding what sort of assemblages should be generated at residential sites where Individual Provisioning was the dominant system. Elements of the mobile toolkit have extended use-lives so should be infrequently encountered in the archaeological record and – where they are present – show evidence of extensive modification through retouch and reworking (Kuhn 1995; Mackay 2005). While tools may be infrequently discarded, sites should still preserve some traces of heavily modified pieces through debris generated by retouch, reworking and converting blanks into implements (Kuhn 1995:26). This evidence should comprise small flakes of good quality raw material. Some level of primary reduction may also occur by the use of small, portable cores, also of higher quality material – this flaking should be standardised and show attempts at raw material conservation, through the use of blade cores or attempts at rejuvenation by overhang removal (Graf 2010; Kuhn 1995; Mackay 2005; Mitchell 2016:159; Parry and Kelly 1987). As with elements of mobile toolkits, cores themselves may be rare within archaeological deposits (Mackay 2005), but evidence of their use should be preserved in flaking debris. Lower quality materials would be used more expediently and generate a similar signature to Place Provisioning, i.e. the artefacts would show no evidence of standardisation, retouch, or raw material conservation, and pieces would be discarded after light use (Kuhn 1995; Mackay 2005). Presumably, a more limited quantity of debris would accumulate via the use of expedient technology as a short-term risk reduction strategy, when compared to its widespread use within Place Provisioning. 2.4.2 Place Provisioning As residential mobility decreases, the duration and predictability of occupation increases (Figure 2.3a; Kuhn 1995). It is therefore possible for people to identify areas in the landscape where lengthy visits can occur at regular intervals in the 31 future, and to provision these places by stockpiling raw material. Stockpiling material negates the ongoing transport costs associated with mobile toolkits, while simultaneously avoiding direct costs of raw material transport by embedding collection within foraging tasks – raw material selection therefore reflects the foraging range (Bamforth 1986; Binford 1979; Kuhn 1995). The abundance of stone means that people can adopt a less formal, unstandardised, expedient technology, whereby tools are made as required and discarded once they are no longer functional or after a task has been completed (Binford 1977; Kuhn 1995). Stone consumption is much higher than that associated with formalised reduction techniques, so expedient technology can only be widely adopted if stone is abundant, either by stockpiling or through natural abundance (Bamforth 1986; Kuhn 1995; Nelson 1991; Parry and Kelly 1987). Nevertheless, informal, expedient technologies have several important benefits over the more standardised, formal reduction techniques that characterise Individual Provisioning: o Unmodified raw material can be converted into whatever type of tool is required, without being constrained by the size and form of blanks or small transported cores (Kuhn 1995:24); o Unstandardised reduction is faster and easier than more formal reduction techniques (Parry and Kelly 1987); o Fewer costs are entailed by discarding a blunt or broken tool and making a new one, when compared to reworking or recycling an existing tool (Bamforth 1986); o A fresh edge (i.e. newly detached flake) is always sharper than a retouched edge (Parry and Kelly 1987); o A specialised tool (suited to a narrower range of tasks) will always be more efficient than a generalised, multifunctional tool (Torrence 1983); o Raw material quality is less of a constraint – lower quality material can be incorporated since tools are discarded after use or breakage and do not need to be reworked (Graf 2010; Manninen and Knutsson 2014; Mitchell 2016:152). Increased stone consumption means that expedient strategies will create much more archaeological debris than Individual Provisioning and may overwhelm the latter 32 where both systems were in use at a single location (Binford 1977; Kuhn 1995:26). As the focus is on primary reduction, cores should be a frequent component of the stone artefact assemblage, and both cores and flakes will be much larger than under Individual Provisioning (Mackay 2005; Speal 2009; Sullivan and Rozen 1985). Flakes should be unretouched and many will show little or no macroscopic evidence of use, while broken or worn tools will be discarded without further modification – no efforts were made to extend the use-life of individual artefacts (Kuhn 1995; Mackay 2005); the lack of retouch should limit the number of very small flakes. There should be little evidence of standardised flaking, manifested by a wide array of artefact sizes and shapes. Raw material choices will be indicative of the foraging range, but may trend towards the best material available locally – it should show little pre-processing before transport, evidenced by larger fragment sizes and the presence of cortex (Binford 1979; Clarkson 2007:24–25). Despite the abundance of raw material, lengthy visits and the time since last raw material procurement may encourage some level of raw material conservation and economisation (Clarkson 2007:25), i.e. by using bipolar reduction that allows small cores to be reduced further than would be possible using other techniques. Hiscock (1996) found that even where raw material was abundant, the incidence of bipolar flaking increased with occupation duration, while Parry and Kelly (1987) found that bipolar artefacts were common among expedient assemblages. Therefore, it seems likely that bipolar flaking would be characteristic of assemblages generated by Place Provisioning. While it may seem contradictory to employ a conservation strategy within a more wasteful, expedient technology (one characterised by the short-term use of artefacts and lack of reworking/recycling), it is possibly due to costs relative to benefits. Since it only alters the mechanism of primary reduction, the bipolar technique represents a simple, low-cost means of conserving raw material while retaining the benefits of an expedient technology. Reduced residential mobility may, in some cases, be accompanied by increased logistical mobility, especially where environments are patchy and returns are variable (Binford 1980; Clarkson 2007:24; Kuhn 1995:27). Under these conditions, people travel greater distances from the central residential location in search of resources, a characteristic of Binford's (1980) collectors. Specific tools would be required for long 33 logistical forays, beyond those expediently manufactured items used for tasks occurring at the residential site. People would therefore 'gear up' at the residential location by manufacturing the required tools (Binford 1979). This is a form of Individual Provisioning, but it occurs in concert with, rather than in opposition to, Place Provisioning (Figure 2.3). Longer logistical forays generally target specific resources, since these must be highly ranked enough to offset transport and fieldprocessing costs (Binford 1980; Orians and Pearson 1979). A narrower and more predictable range of tasks would be conducted, so tools could be more specialised – and therefore more efficient – than the generalised toolkit required under conditions of high residential mobility. Portability and durability would still be integral components, however (Clarkson 2007:24; Torrence 1983). Therefore, while expedient technologies can accommodate lower quality materials, more careful selection of stone would presumably characterise tools required for long logistical forays. Evidence will be scarcer than the debris arising from expedient tool manufacture but – if present – it should be distinguishable from traditional Individual Provisioning. During the process of 'gearing up', tools are made from cached raw materials, rather than the transported blanks or small cores used under traditional Individual Provisioning. Therefore, evidence of primary reduction should dominate, and debris should be more abundant than under traditional Individual Provisioning and also show a greater variety of sizes, including larger flakes and cores; smaller flakes arising from retouch may occur where a tool’s function required a retouched edge. Standardised reduction and raw material conservation should be far less evident than under Individual Provisioning. 2.4.3 Summary Perhaps the greatest success of Kuhn's (1995) scheme is its reliance on flaking methodology and retouch indices to identify signatures characteristic of Individual and Place Provisioning. In contrast to models that use assemblage diversity or design criteria, Kuhn's can be applied to fairly amorphous lithic collections dominated by debris with few formal tools and little typological variation. This perhaps explains why his model is commonly invoked when interpreting Australian lithic assemblages (e.g. Clarkson 2004, 2007; Mackay 2005; Mitchell 2016) that are well-known for their amorphous nature (Hiscock and Clarkson 2000). While Individual and Place Provisioning have distinct technological signatures, Kuhn (1995:25–26) noted that 34 these strategies are not mutually exclusive. All mobile groups employ some form of Individual Provisioning, in that there are some items that people always transport with them, and some level of Place Provisioning. The dominant system simply varies according to the circumstances at a given time or place, as a group's mobility fluctuates in response to resource availability. Shifts in provisioning systems will be most evident, then, from broad, inter-assemblage analyses rather than considering lithics from an isolated site (Kuhn 1995:29). 2.5 CONCLUSION Forager models demonstrate that optimisation can be achieved under varying resource distributions by altering several components of hunter-gatherer life, including which resources should be targeted, when one foraging patch should be rejected for another, when resources should be processed before transport, whether people should aggregate in a large group or disperse into small units, and how frequently they should move their residential base. These findings are robust enough to use without reference to the underlying mathematical models and are frequently incorporated into archaeological analyses. Technological organisation can be influenced by several factors, including mobility, raw material availability, and risk, which are interrelated, to a certain degree. These factors dictate where and when technological activities occur, which raw materials are used and to what extent they are reduced, whether formal or informal technologies are adopted, and the reduction techniques used to produce them. However, technological organisation must be considered via a multifaceted approach geared towards local conditions, as no single factor is universally dominant. Significantly, Kuhn's (1995) provisioning systems unite forager theory with technological organisation and define optimal technological solutions under different levels of residential mobility; more importantly, his scheme can be applied to amorphous Australian assemblages. Using his predictions in concert with local resource distribution, it should be possible to identify those parts of the landscape where more frequent and less predictable moves would favour Individual Provisioning, in contrast to more scheduled, predictable moves that permit Place Provisioning. This provides a valuable context for analysing and interpreting fragmentary archaeological remains. 35 CHAPTER 3 NATURAL ENVIRONMENT AND LANDSCAPE OF THE STUDY AREA 3.1 INTRODUCTION This chapter outlines the natural environment of the study area, focussing on those aspects that would dictate the availability of resources for its human occupants. The geology, landform, and hydrology are described, as well as the impact of vegetation clearing on hydrological conditions. The Quaternary climate of the continent is evaluated, focussing on those periods relevant to human occupation in Australia – the late Pleistocene and the Holocene. The Holocene palaeoenvironmental records of southwestern Australia are summarised, with particular focus on the few records from inland areas; the modern climate of the study area is also quantified, as it forms a framework upon which to superimpose past climatic changes. Finally, the preEuropean distribution of flora and fauna is assessed after considering the impact caused by clearing native vegetation for agriculture. 3.2 GEOLOGY AND LANDFORM The study area, like much of southwestern Australia, is underlain by the Yilgarn Craton (Figure 3.1). At 657,000 km2, this is the largest of the continent’s three cratons, each of which represents a fragment of the Earth's crust that has avoided significant deformation through tectonic processes for more than a billion years (Anand and Butt 2010; Blewett et al. 2012; Johnson 2009:59). The rocks that make up the Yilgarn – mainly granitoids and gneiss with north-northwest oriented bands of greenstone (heterogeneous deposits of metavolcanic and metasedimentary rock, more common in the eastern portion of the craton) – were formed during the Archaean period, mostly between 3000 and 2600 million years ago (Ma), though some parts date to 3700 Ma (Anand and Butt 2010; Anand and Paine 2002; Blewett et al. 2012). There is considerable variation in the geochronology, geological history, drainage and soil types across the craton, so it should not be considered a homogenous unit. 36 Figure 3.1 Simplified geology of Western Australia, showing extent of the Yilgarn Craton, also known as the Yilgarn Block (Beard 1990:41). Anand and Butt (2010) divided the Yilgarn craton into three regions: Southwestern, Southern and Northern (Figure 3.2), based on a combination of geological, drainage, climatic, soil and groundwater characteristics. The Southern region incorporates the study area and is characterised by low relief and gently undulating terrain, while the adjacent Southwestern region comprises a plateau with high relief and a much more dissected landscape (Anand and Butt 2010); these topographic differences result from drainage histories that diverged around 16 million years ago. Large palaeorivers had once snaked across much of southwestern Australia, but regular flow ceased after the onset of aridity in the mid-Miocene between 16 and 11 Ma (Martin 2006). In the Eocene, uplift created the Darling Range and rejuvenated 37 drainage west of the Meckering line (Commander et al. 2001). New river systems were created in the Southwestern region, but east of the Meckering line (the Zone of Ancient Drainage, which broadly corresponds to the Wheatbelt region) drainage remains uncoordinated, and connected flow occurs only after periods of exceptionally heavy rainfall (Anand and Butt 2010; Beard 1999; Bettenay 1962; Hatton et al. 2003). Figure 3.2 Major regions within the Yilgarn Craton (Anand and Butt 2010:1018). As a result of the inactive drainage and tectonic stability noted above, the topography of the Wheatbelt is, as a whole, flat and subdued. The lowest points in the landscape are the broad, flat valley floors up to 15 km wide – relicts of palaeodrainage systems at least 45 million years old (Commander et al. 2001). The oldest of these deposits are the narrow palaeochannel sands and clays (Figure 3.3) that are concealed by younger, Quaternary deposits that infill the valley. Within these valley floors are a characteristic feature of the inland southwest – chains of salt lakes and playas. Tertiary laterite formations are common in small pockets, while the remnant sandplains derived from this cover a large portion of the landscape (Figure 3.4; Murphy-White and Leoni 2006). Archaean granite outcrops exist throughout the landscape – these are surface expressions of more resilient portions of the 38 underlying bedrock that have resisted the weathering to which the surrounding rock has succumbed (Campbell 1997). Gneisses are largely confined to the east, outside the study area (Figure 3.4). As a whole, the topography is unimposing, and the slopes gradual; only the occasional granite outcrop breaks the horizon. Such outcrops were valued by Aboriginal people as they capture and trap rainfall in crevices and depressions, provide a range of plant and animal foods, and act as natural lookouts (Bayly 1999; Bindon 1997; Smith 1993:20). Figure 3.3 Geology of characteristic Wheatbelt valleys (Commander et al. 2001, reproduced from Department of Environment 2005b). The study area yields three raw materials that are suitable for artefact manufacture: quartz, silcrete and dolerite. Silcrete (Figure 3.4: Czb) occurs within lateritic profiles, generally above deeply weathered bedrock (Czo). It becomes exposed as the overlying duricrust (Czl) erodes or where the silcrete layer is otherwise exposed, as in breakaways (Chin et al. 1984). Silcrete occurs in small patches, predominantly south of the main palaeodrainage channel, but is much more common in the greenstone belt east of the study area (Figure 3.4a). Dolerite is available in the large discontinuous dolerite-gabbro dykes (Figure 3.4: Pd) striking ENE along the main palaeochannel; smaller dykes are fairly common throughout the study area. Quartz occurs as veins in granite bedrock, formed by hydrothermal processes whereby fluid is forced into a void and subsequently cools into a solid – the oldest, coarsest quartz 39 is found in the core of the vein, surrounded by younger, finer textured material (Abeysinghe 2003:6); a single source can therefore yield quartz of varying quality (de Lombera Hermida 2009). Quartz veins are especially common in the Southwestern region and can outcrop at the surface where the surrounding rock has weathered away (Abeysinghe 2003:172, Figure 111). One vein is large enough to appear on geological maps of the study area (Figure 3.4a), but smaller veins could be encountered in any area of exposed bedrock. Pegmatite bodies can yield high-quality quartz from their cores (often 10–15 m wide), as crystals have the space to form unimpeded, so can reach considerable sizes (Abeysinghe 2003:174; Fetherston et al. 2017:91). A single pegmatite vein occurs in the study area (Figure 3.4a) but the cores are generally encased by several metres of pegmatite, so the quartz may be inaccessible unless the core has become exposed (Abeysinghe 2003:164). Quartzite (Alm) and banded chert (Aic) are present in the greenstone belt outside the study area immediately east of the vermin-proof fence (Figure 3.4b) – both are well-suited to artefact manufacture (Chin et al. 1984; Glover 1984). 3.3 HYDROLOGY The study area lies within the Lockhart River catchment, one of four sub-catchments in the Avon Basin. While the Avon is one of Australia's major river systems, most of the 120,000 km2 basin (including the Lockhart catchment) drains internally, meaning connected flows are rare except after exceptionally wet periods (Bettenay 1962; Department of Water 2009; Galloway 2004). This lack of flow means that weathering products and, more importantly, salts, are not removed from the system and instead accumulate over long periods (Anand and Butt 2010). Salt has been naturally accumulating in the soils for many thousands of years, transported by westerly winds from the Indian Ocean, but it is also present in rainfall and in the regolith (Blewett et al. 2012; Clarke et al. 1998; De Dekker 1983; Hatton et al. 2003; Salama 1994). These salts are concentrated in areas where water pools and subsequently evaporates (salt lakes and playas), leaving behind ever-increasing quantities of salt that active drainage would otherwise remove. Salinity is, therefore, a natural and prominent part of the landscape (Bettenay et al. 1964; Department of Water 2009). 40 Figure 3.4 Geology of selected parts of the study area (A) and beyond (B), showing the occurrence and typical distribution of potential tool-stone. Geological imagery (including legend) from Geological Survey of Western Australia (1984). Note that geological maps depict the surface and near-surface geology, so raw material sources may not have all been accessible at ground-level. 41 This natural salinity defines the landscape of the study area, which is characterised by saline groundwater, saline surface features, and a lack of freshwater lakes or rivers. Saline groundwater results from the interplay of several factors including increasing aridity, long residence times, low hydraulic gradients, and evaporative concentration of salts around playas and discharge features (Anand and Butt 2010; Bettenay et al. 1964; George 1992; Magee 2009). Aside from on the coastal plain, most groundwater is above the safe salinity level for human consumption, around 2100–1600 mg/L (Figure 3.5; NHMRC, NRMMC 2011:973–974). The salinity levels are even greater in the study area, commonly > 20,000 mg/L, and up to 55,000 mg/L in palaeodrainage channels (Bowen and Benison 2009; Ghauri 2004; Hollick 2004). The saline and hypersaline groundwater combined with low rainfall accounts for the lack of freshwater surface features in the study area. In contrast, saline surface features have been part of the landscape for millions of years, acting as discharge points for naturally occurring saline groundwater (Bettenay et al. 1964; Bowler 1986; Department of Water 2009; George and Coleman 2001; Salama 1994). Nevertheless, salt lakes and playas are not static features, but instead respond to climate change. George and Coleman (2001) argued that current salt lakes in the Wheatbelt are presently half the size they were 20,000 years ago, during the last dune-building phase. Salt lakes are also dynamic across shorter time-scales, as water-level and salinity vary throughout the year in response to rainfall (Bowler 1986; Halse et al. 2003). Much of the landscape has also been affected by secondary (dryland) salinity, an unfortunate by-product of agricultural development, specifically the clearing of native vegetation for introduced crops. In the study area, less than 13% of native vegetation remains (Department of Water 2009). Native plants are deep-rooted, perennial, and will effectively use all water available to them over the course of a year (Mitchell et al. 2009; Wildy et al. 2004). In contrast, introduced crops are shallow-rooted annual plants that are inactive for several months of the year, and therefore have lower water requirements. Hatton et al. (2003:344) concluded that before the Wheatbelt was cleared, 'very little rainfall was discharged in liquid form'. Instead, most rain infiltrated the soil where it fell and was available for plant use or evaporated from the ground on which it landed; runoff was limited, as was recharge to the regional aquifer 42 system (Farrington and Salama 1996; Hatton et al. 2003). Rainwater that would previously have been intercepted and utilised by native vegetation is now surplus to crop needs and is redistributed as runoff to lower lying areas (e.g. valley floors) where it infiltrates the soil and accumulates. This increased recharge can significantly raise the water table in just decades (Farrington and Salama 1996; George et al. 1997; Hatton et al. 2003). Figure 3.5 Groundwater salinity across the Yilgarn Craton (Anand and Butt 2010:1078). As the water table rises, it dissolves and mobilises the salt present in the deeper parts of the soil profile below the major rooting zone for vegetation and brings it closer to the surface, where it is detrimental to plant growth (Clarke et al. 2002; Hatton et al. 2003). This process creates newly salinised land and increases the salt concentration in naturally saline land. In many cases, Melaleuca shrubs (and associated species) that inhabit naturally saline areas cannot adapt to increased salt loads, so these plants die and are replaced by those with a greater salt tolerance, such as samphires (Department of Water 2009). Stands of dead vegetation are common in these areas, indicating the very short timeframe across which these processes operate (Figure 3.6). Salt lakes have also been drastically affected – under native vegetation conditions, they would have held water for a few months 43 after winter rains and been dry by summer, except where significant summer rains were generated by cyclones that occasionally arrived from the North West (Department of Water 2009; Halse et al. 2003). The rising water table, in conjunction with additional input from surface runoff, means that these lakes now hold water for far longer and, in many cases, have lost their natural wetting and drying cycles to the detriment of the local plants and animals (Department of Environment 2005a; Department of Water 2009; George et al. 2008; Halse et al. 2003). The water is becoming saltier, and salinity levels now frequently exceed 200,000 mg/L, although salinity and water chemistry vary between water bodies (Benison et al. 2007; Bowen and Benison 2008; Cale et al. 2004). Figure 3.6 Recently dead vegetation near Lake Gounter, Hyden, in October 2016 (Photograph: A M Rossi). Note the replacement of large woody shrubs by much smaller plants with a greater salt tolerance. 3.4 CLIMATE Today, Australia has several climatic zones – from the temperate south, the arid deserts of the red centre, to the tropical north. This variation is unsurprising considering the continent spans 33 degrees of latitude and 40 degrees of longitude; past climates would have varied similarly, but conditions were very different from those at present. Climatic variation is driven by the interplay of many factors, such as palaeogeography, volcanic activity, insolation levels, as well as shifts in the Earth’s orbit and rotational axis (Hays et al. 1976; Laskar et al. 2004; Wanner et al. 2008). The characteristics of this past climate can be determined through a variety of methods, including the analysis of ice- and sediment-cores, pollen, alluvial/lacustrine deposits, and speleothems (e.g. Churchill 1968; Galloway 1965; Harrison 1993; Moros et al. 2009; Pickett et al. 2004; Quigley et al. 2010). Most techniques involve 44 the use proxy data such as lake levels, for example, which respond to the amount of effective precipitation received during a specific period. Effective precipitation is dictated by the combination of total precipitation, temperature, evaporation, wind, and vegetation cover (Reeves et al. 2013). Hence, it can be very difficult to quantify specific climatic conditions, so periods are often characterised only in a relative sense (i.e. wetter vs. drier). The late Quaternary, which incorporates the Pleistocene and Holocene epochs, is the only period relevant to human occupation of Australia. Changes in the Earth’s orbit and rotation (the Milankovitch cycle), are largely responsible for the characteristic feature of the late Quaternary global climate, where periods of aridity – glacial periods – are interspersed with interglacials, when conditions were wetter and more favourable to humans (Bowler et al. 1976; Galloway 1965; Kershaw et al. 2003; Martin 2006). The changing sea-level, which resulted from ice sheet expansion and retraction, would have intensified the variations caused by astronomical cycles. During the Last Glacial Maximum (21 ± 3 ka: Reeves et al. 2013) continent-wide climatic data were most synchronous: sea-surface temperatures were 3–6 °C colder than present; sea-levels were approximately 100– 130 m lower than today, exposing around 3 million sq. km of continental shelf; aeolian/dune-building processes were at their height; and drier conditions are evident from many terrestrial records (Barrows et al. 2007; Brooke et al. 2017; Johnson 2009:180; Lewis et al. 2013; Nanson et al. 1992; Reeves et al. 2013; Zheng et al. 2002). However, effective precipitation may have reached its minimum a few thousand years later, between 15 and 11 ka, but the timing varied significantly across the continent (Harrison 1993; Kershaw and Nanson 1993; Reeves et al. 2013; Turney et al. 2006; Williams et al. 2009). Worldwide, Holocene climatic variation was far less severe than during the Pleistocene (Mayewski et al. 2004; Wanner et al. 2008); this pattern is echoed in Holocene Australian records (Bowler et al. 1976; Chappell 1991; Dodson 2001). Broadly, the early Holocene (12–8 ka) was characterised by warmer and wetter conditions across the continent, but the timing of onset varied regionally (Reeves et al. 2013). Seas reached their current levels between 8 and 7.5 ka (Lewis et al. 2013). The late Holocene (5–0 ka) climate was more variable, a characteristic 45 attributed to the El Niño Southern Oscillation (ENSO) in the north and east of the continent, and a redistribution of westerly wind belts in the south (Reeves et al. 2013). The varying distribution and influence of these systems resulted in another characteristic of the late Holocene – distinct spatial variation in climate – which means that it is important to consider southwestern Australia separately. 3.4.1 Palaeoclimate of Southwestern Australia It is well known that southwestern Australia is poorly represented in palaeoenvironmental records (Gouramanis et al. 2012; Hesse et al. 2004; Petherick et al. 2013; Reeves et al. 2013); long-term Holocene data are available from just a few sites (Figure 3.7). While the specifics of most studies are irrelevant here, the discordance between their results is noteworthy (Figure 3.8; Table 3.1). Opinion is divided about whether the current climate is wetter (Churchill 1968; Dodson and Lu 2000; Kendrick 1977; Semeniuk 1986; Yassini and Kendrick 1988), drier (Gouramanis et al. 2012; Harrison 1993; Zheng et al. 2002), or the same (Newsome and Pickett 1993) as that in the mid- to late Holocene – scholarly opinion is equally divided concerning the timing and length of perceived wet/dry periods (Figure 3.8). Hence, the discordance between the palaeoenvironmental studies ‘raises the question of how much reliance should be placed on the results’ (Newsome and Pickett 1993:246). Due to the location of the study area, inland records (Harrison 1993; Zheng et al. 2002) are most relevant, and are discussed below. These data are more coherent, but this may be a product of an extremely small sample size. Harrison’s (1993) continental lake-level analysis includes four sites from southwestern Australia: Lake Grace, Lake King, Myalup Swamp and Storeys Lake, but the latter only preserves an isolated record at 17,000 BP. The remaining three sites provide a record from 6000 BP to the present, but data are never available for all three sites for any specific millennium (Figure 3.9). Lake-levels were inferred from lithology or stratigraphy, and multiple radiocarbon (14C) dates were available for each site except Myalup Swamp, for which only one age determination was obtained. The data are extremely limited, but the overall pattern indicates higher lake-levels at 6000 and 5000 BP, while low-lake states dominate thereafter until the present (Harrison 1993). 46 Figure 3.7 Palaeoenvironmental sites mentioned in the text. Those in and around Perth are not marked. Figure 3.8 Major climatic trends identified in southwestern Australian palaeoenvironmental studies. Periods were classified in relation to the preceding period; blue = wetter, red = drier, black = no change, grey = uncertain. All dates have been plotted as cited, although Churchill’s (1968) dates have been converted from BC/AD to cal. BP and plotted on the secondary axis with the data obtained by Gouramanis et al. (2012). 47 Table 3.1 Summary of main palaeoenvironmental data from southwestern Australia – research focussing on the recent past has been omitted (e.g. Cullen and Grierson 2009; Treble et al. 2003). Temperature data from Gouramanis et al. (2012) as it does not span the entire record. Dates are cited in the format provided by the author. * = Perth metro area, not marked on Figure 3.7. Author Sites/study area Proxy Major palaeoenvironmental results Churchill (1968) 1. Boggy Lake 2. Flinders Bay Swamp 3. Fremantle* 4. Myalup Swamp 5. Perth* 6. Rottnest Island 7. Scott River Swamp 8. Weld Swamp 9. West Lake Muir Pollen Several shifts from wet to dry conditions, as follows: o 5000–2600 BC: wetter o 2600–500 BC: drier (maximum at 1200 BC) o 500 BC – AD 400: wetter o AD 400–1400: drier o AD 1400–present: wetter Kendrick (1977) Swan River near Guildford* Molluscs Drier conditions from before 6700 BP to around 4500 BP. Wetter conditions thereafter. Semeniuk (1986) 1. 2. 3. 4. Lancelin Whitfords* Rockingham Leschenault Peninsula Calcrete 7000–3500 BP mostly semi-arid, small band along west coast with subhumid-humid. After 2800 BP regional trend to increased humidity. Yassini and Kendrick (1988) Point Waylen* Ostracods Drier conditions 6000– 4100 BP. Wetter conditions thereafter. Harrison (1993) 1. Lake Grace 2. Lake King 3. Myalup Swamp Lake-levels Higher lake levels 6000–5000 BP. Low-lake states almost exclusive thereafter. Newsome and Pickett (1993) 1. Loch McNess Pollen 2. Boggy Lake Largely stable climate 9000 BP to present; possibility of shorter arid periods throughout. Dodson and Lu (2000) Byenup Lagoon Pollen Record begins around 4800 BP, indicating regional increase in effective precipitation. Zheng et al. (2002) 1. Lake Cowan 2. Lake Lefroy Aeolian and lacustrine sediments High lake event 10,000– 4000 BP. Reactivation of aeolian activity indicating increased aridity 4000– 2000 BP, possibly into the present. Gouramanis et al. (2012) 3. Barker Swamp Pollen, charcoal, gastropods, ostracods Record begins 7.4ka, indicating regional increase in effective precipitation. Several smaller shifts: o 7.4–7.2 ka: wetter o 7.2–6.7 ka: drier o 6.7–6.2 ka: wetter o 6.2–5.8 ka: drier o 5.8–4.5 ka: wetter o 4.2–2.3 ka: drier o 2.3–1.4 ka: wetter o 1.4 ka – present: dry/climate as today. 48 Figure 3.9 Harrison’s (1993:219) reconstruction of lake status at 1000-year intervals, from 6000–1000 BP. 49 Finer grained data are available from the Lake Cowan/Lake Lefroy system more than 600 km east of the present coastline. Based on optically-stimulated luminescence (OSL) and radiocarbon dates from aeolian and lacustrine sediments, Zheng et al. (2002) proposed that the Holocene climate transitioned through two major phases: 1. Early to mid-Holocene: High lake event 10,000–4000 BP; 2. Late Holocene: Reactivation of aeolian activity 4000–2000 BP, possibly continuing into the present. If aeolian activity and lake-levels are indeed accurate proxies for effective precipitation, then conditions at Lakes Cowan and Lefroy were wetter from 10,000 to 4000 BP and became drier thereafter. 3.4.2 Modern Climate of the Study Area To understand the variation in effective precipitation discussed above, it is necessary to consider the baseline environmental conditions upon which these variations are superimposed. The study area is within Beard’s (1981, 1990) Extra Dry Mediterranean climatic zone (Figure 3.10), where maximum rainfall occurs in winter and there are seven or eight ‘dry’ months (where evaporation exceeds precipitation) each year. In Hyden, maximum summer temperatures average 34° C but may reach 48° C. Winter minima average 5° C, but may fall as low as -6° C at night (Bureau of Meteorology data). Average annual rainfall is approximately 340 mm (based on 1929–2017 data), but in many years this average is not reached; every five to ten years falls are significantly above average (Figure 3.11); this highly variable rainfall pattern is characteristic of Mediterranean climates (Dodson 2001). If the limited palaeoenvironmental data for inland southwestern Australia (Harrison 1993; Zheng et al. 2002) are applicable, then this climatic pattern has been in place for at least 4000 years. 50 Figure 3.10 Bioclimatic zones of Western Australia (Beard 1990:39). Figure 3.11 Annual rainfall data from Hyden weather station (no. 10568), 1929–2017 in relation to average rainfall (blue dashed line). Data are not available for 1947 or 1974. 51 3.5 VEGETATION Palaeovegetation data for southwestern Australia are rare, but those available data indicate little shift in vegetation patterns during the Holocene (Dodson 2001; Dodson and Lu 2000; Itzstein-Davey 2004; Newsome and Pickett 1993; Pickett 1997; Pickett et al. 2004). The one site with a longer record spanning the last 31,000 years (Bibra Lake, in Perth), indicates that the biome around Perth at 18,000 BP was the same as that of the present-day (temperate sclerophyll woodland and shrubland) but different species may have dominated (Pickett 1997; Pickett et al. 2004). Regardless of whether floral stability indicates a largely unchanging climate or the resistance of vegetation communities to climate change, the vegetation patterns evident before agricultural development had probably been in place for millennia. 3.5.1 Botanical Provinces of Western Australia Beard (1990) divided Western Australia into three main provinces (Figure 3.12a): South-West (SWP), Eremaean (EP) and Northern Province (NP). The boundaries of these major provinces are broadly congruent with the climatic zones shown in Figure 3.10, indicating that, as Beard (1990:50) stated, the vegetation of Australia has adapted to thrive in the nutrient-poor conditions and is instead constrained only by the amount of moisture in the soil. The study area lies within Beard’s (1990) SouthWest Province, whose eastern boundary roughly separates the Extra-Dry and SemiDesert Mediterranean zones (Figure 3.10). Here, deeply weathered soil profiles allow plant roots access to the pallid zone (weathered rock) where moisture accumulates, enabling trees to survive through the dry season. This unique feature of the landscape allows much more densely wooded areas, such as those in the lower southwest, to be found within Mediterranean-type climates (Beard 1990:50, 55). The SWP is a biodiversity hot-spot, known for its high proportion of endemic species, around 3000 of the 5710 identified (Beard et al. 2000). Clearing has had a disastrous effect on the natural environment of the study area, but the floral species themselves endure. In the Wheatbelt (approx. 75% of the SWP), almost 350 species are now considered rare or endangered, but by the 1980s only 24 had been rendered extinct (Hobbs 2003; Hopper et al. 1990; Leigh et al. 1984). Beard’s (1990) provinces have been separated into a number of regions (Figure 3.12a), most of which have been retained in IBRA 7 (Interim Biogeographic 52 Regionalisation for Australia, Version 7 – Figure 3.12b). Most of the study area falls within the Mallee region, while the northern extremity lies in the Avon Wheatbelt region. The Mallee Region is so-named for the dominance of the mallee vegetation structure, comprising small multi-stemmed eucalypts. Mallee originally covered around 75% of the region, while woodlands, scrub-heath and thicket were comparatively rare (Beard 1990:130 & 135). In contrast, the Avon Wheatbelt only preserves small isolated patches of mallee, and is instead dominated by woodlands and thicket (Beard 1990:118). Beard (1990) related the distribution of these different vegetation units to topography, and therefore soil type and depth (Figure 3.13), but there is some argument concerning exactly how topography and associated variables influence vegetation formations on a more general level (Dirnböck et al. 2002). The broad catenary sequence is similar in both regions, with hypersaline areas inhabited by halophytes, heavy lowland soils by woodlands, duplex slopes by mallee, and sandplains by thicket and/or scrub-heath (Figure 3.13). Figure 3.12 A: Natural regions of Western Australia (Beard 1990:34); B: IRBA 7 regions of Western Australia (Australian Government Department of Environment and Energy 2012). Abbreviations largely reflect Beard’s (1990) subregion names except: AVW = Avon Wheatbelt; COO = Coolgardie, GES = Geraldton Sandplains, JAF = Jarrah Forest, SWA = Swan Coastal Plain, VIB = Victoria Bonaparte, YAL = Yalgoo. 53 Figure 3.13 Catenary sequence for the Mallee Region (top) and Avon Wheatbelt Region (bottom), showing the relationship between topography and vegetation communities (Beard 1990:116, 129). 3.5.2 Vegetation Communities in the Study Area Beard (1975) mapped the entire South West Botanical Province at 1:250 000 scale, the highest resolution vegetation data available for the entire region. He aimed to reconstruct the distribution of native vegetation before European arrival, using remnant vegetation, aerial photography and soil-vegetation catenae. Vegetation was classified based on the nature, size and density (canopy/foliage coverage) of the dominant stratum, plus any diagnostic plant species. The spatial data are preserved NRInfo (DPIRD 2016b), as well as published maps and accompanying memos (Beard 1972, 1976, 1979, 1980). Where particular plant communities recurred following the same catenary sequence, they were grouped into vegetation systems – effectively subdivisions of a botanical province (Beard 1976). Three vegetation systems are present in the study area: the Hyden, Muntadgin and Skeleton Rock systems. The same catenary sequences are evident in each, although they preserve different proportions of vegetation communities due to local 54 variation in landform and soils. These communities are summarised in Table 3.2 and their distribution mapped in Figure 3.14. Table 3.2 Vegetation communities found in the study area. Community names, descriptions and classifications are based on Beard (1972, 1975, 1976, 1979, 1980). Note that these descriptions apply only to the study area (i.e. the Hyden, Muntadgin and Skeleton Rock systems), and may not reflect community characteristics at a broader scale. Vegetation community Description Broombush Thicket Very dense one- or two-layered shrub community. Dense upper layer dominated by Allocasuarina, Acacia and Melaleuca > 1 m, generally 2–2.5 m tall, 30–70% canopy coverage. Lower layer may be absent or suppressed, comprising sparse layer of shrubs < 1 m tall. Granite rocks Outcrop bare or lichen-clad; small trees and shrubs may occur where sufficient soil has accumulated in clefts and depressions. Fringe vegetation diverse and varied, frequent thickets and woodlands. Mallee Open Eucalypt shrub assemblage, > 1 m but < 10 m tall – canopy coverage generally 10–30%, but up to 70%. Eucalyptus eremophila almost always present, associated with several other eucalypts. Variable low shrub understorey, often dominated by Melaleuca. Melaleuca thicket On swampy (saline) ground. Closed (30–70%) community of Melaleuca shrubs > 1m tall. Scrub-heath Mixed, stratified, partly open shrub assemblage. Lower closed layer of shrubs < 1 m tall, 30–70% canopy coverage. Open (10–30%) upper layer of shrubs > 1m tall. No dominant species. Succulent steppe On saline soils. Samphire communities comprising low succulent shrubs. May form ground layer only or be associated with teatree shrubs and/or Eucalypt woodland. Teatree scrub On saline soil. Irregular open community of shrubs 1–6 m tall, commonly 2.5–3.5 m. Single-layered formation, canopy coverage 10–30%. Several species, all Melaleuca, with occasional admixture of small eucalypts. Ground vegetation absent, or scattered grass and samphire. Woodland On the heaviest soils, lower in the landscape. Woodland comprising Eucalyptus trees 10–25 m tall, with 10–30% canopy coverage; dominated by E. salmonophloia, often in association with E. longicornis. Understorey dominated by Melaleuca pauperiflora, irregularly scattered and reaching up to 3.5 m tall, plus various eucalypts. Lower shrub layer varies depending on proximity to salt. 55 Figure 3.14 Pre-European distribution of vegetation communities in the study area. Spatial data extracted from NRInfo (DPIRD 2016b), based on Beard’s (1972, 1979, 1980) data. Continuous black line separates the two IBRA regions that meet in the study area: the Avon Wheatbelt (north) and Mallee (south). Dashed line represents the boundary of the study area. Mallee is dominant in the south of the study area, reflecting its local prevalence within the Mallee region; it has a much more restricted distribution further north, within the Avon Wheatbelt region, but there are sizeable areas of mallee with patches of woodland. True woodlands are much more limited in extent, but substantial areas occur around larger salt lakes in the south-eastern portion of the study area where soils are heavier. While structurally different, similar species are often found in woodland and mallee (Beard 1972, 1979). Thicket and scrub-heath have a mutually exclusive distribution: the former is prevalent in the north of the study area and the latter restricted to the south. The structure of these formations is also inverted; thickets comprise a dense upper layer of taller shrubs, sometimes accompanied by a sparse understorey of small shrubs, while heath exhibits a dense lower shrub layer and a sparse upper layer (Table 3.2). Similar species may be present in both formations, though thickets often have a less diverse assemblage where the understorey is absent (Beard 1972, 1979). Teatree scrub and Melaleuca 56 thicket are similar, both found on saline soils in palaeochannels, but the shrubs are more densely packed in thickets. Succulent steppe is much more limited in distribution, confined to the southwestern portion of the study area in the main palaeochannel. Finally, granite rocks outcrop throughout the study area. These preserve a multitude of microhabitats, ranging from eternally dry soils to permanently shaded and moist cave deposits, identifying granite outcrops as local biodiversity hotspots (Hopper et al. 1997). Outcrops channel water, nutrients and soil to the base of the rock, permitting denser vegetation than would otherwise be possible while also providing valuable water sources for humans (Bayly 1999; Bindon 1997; Schut et al. 2014). Each of the vegetation communities described above would offer a unique suite of plant foods for human consumption, from eucalypt-derived lerps and manna, to Acacia and other tree seeds, as well as various fruits, roots and flowers (see Chapters 5–6). 3.6 FAUNA The native fauna of the Wheatbelt has been significantly impacted by extensive landclearing. Of the 43 native ground-dwelling mammal species that originally inhabited the Wheatbelt, 13 have disappeared since European arrival, and just 12 species are considered abundant or common (Kitchener et al. 1980). A total of 195 bird species have been identified, of which two are now extinct and 95 have suffered a decline in range or abundance (Saunders and Ingram 1995). At least 22 species have become more widespread and numerous following agricultural development (Saunders 1989); predatory birds are now more common due to introduced prey including pests (mice, rabbits) and stock animals, while other birds have been attracted by new plant foods such as cereal crops (Storr 1991). A total of 1000 wetland invertebrates, 665 ground-dwelling spiders, 86 reptile and 21 frog species have been identified in the Wheatbelt (McKenzie et al. 2004; Pinder et al. 2004), but the impact of clearing on these populations is unknown. Kitchener et al. (1980) argue that few lizards have become extinct since populations persist in small patches where their habitat is preserved, such as granite outcrops. The species mentioned above would not all be present in the study area, as the Wheatbelt incorporates an area of almost 25 million hectares (Hobbs 2003). While 57 smaller scale faunal data are available, they are limited to surveys of isolated reserves. For example, McKenzie et al. (1973) studied the faunal assemblage at Dragon Rocks, 30 km southeast of Hyden, and identified just 15 native mammals (including macropods, possums, and the echidna – Tachyglossus aculeatus), 59 birds (including the emu – Dromaius novaehollandiae), 19 reptiles (including two dragons) and four amphibian species. This list likely underestimates pre-clearing faunal diversity but indicates the variety of animal foods available to Aboriginal people, who are known to have eaten almost any animal they could catch (Hassell 1975:10). 3.7 CONCLUSION Long periods of tectonic stability and inactive drainage have created a characteristically subdued landscape in the study area, dominated by gentle slopes and low relief; salt lakes are common in the relict palaeodrainage systems. Clearing of native vegetation has altered the landscape by raising the naturally saline water table and creating salinised and waterlogged soils. Salt lakes have been a natural part of the landscape for thousands of years, but before clearing they probably only held water for a few months each year. While the pre-clearing landscape can be quantified there is, at present, no coherent palaeoenvironmental reconstruction; opinion is divided about whether the current climate is wetter or drier than that of the preceding period. The limited data from inland southwestern Australia indicate that drier conditions have been in place for the last 4000 years. Prior to clearing, mallee covered most of the study area, with scrub-heath in the south transitioning to thicket in the north; woodlands and salt-tolerant vegetation was also present. The fauna of the study area suffered devastating species loss following the introduction of agriculture, but several bird species have benefitted from the introduced food sources and more abundant water. Hence, modern faunal surveys do not necessarily accurately represent pre-European distribution. 58 CHAPTER 4 CULTURAL CONTEXT 4.1 INTRODUCTION This chapter provides the cultural context, both archaeological and ethnohistorical, for the study area. The archaeological background is evaluated, including the earliest evidence for Aboriginal people in Australia, in marginal environments, and in the southwest. The few southwestern studies with some inland component are then discussed. Finally, the known archaeological locations within the study area are identified – they are few in number, and none have been investigated in detail or radiometrically dated. The ethnohistorical sources are then analysed, though these are few, fragmentary, and fraught with biases. Ethnographic information collected during recent cultural heritage surveys provides a valuable supplement to this limited dataset. 4.2 ARCHAEOLOGICAL BACKGROUND It is generally accepted that Aboriginal people first moved from island South East Asia to Sahul (the landmass comprising mainland Australia, Tasmania and New Guinea, which were joined at times of lower sea level), sometime around 60,000 years ago. Two sites with occupation evidence from 60–50 ka, Nauwalabila and Madjedbebe/Malakunanja II, were deemed problematic by some due the questioned association between artefacts and dated materials (see contrasting debates in Hiscock 2013 and Allen and O'Connell 2014). However, Clarkson et al. (2017) reexcavated and re-dated Madjedbebe and demonstrated compelling evidence in the form of diverse stone artefact assemblages, ochre, and evidence of plant processing for occupation by 65,000 cal. BP. Nevertheless, the main body of archaeological evidence in Australia post-dates 47,000 cal. BP (e.g. David et al. 2013; Morse et al. 2014; Slack et al. 2004; see list in Allen and O’Connell 2014). Following colonisation, cross-continent movement was fairly rapid, and most parts of the continent, including the present arid zone, have sites pre-dating 40,000 cal. BP (e.g. Allen and O’Connell 2014; Smith 2013:79–80). 59 Aboriginal people were in southwestern Australia relatively soon after initial colonisation – while 50,000 cal. BP dates for Devil's Lair (Turney et al. 2001) are disputed, the 45.44 ± 2.57 ka date is widely accepted (Allen and O'Connell 2014; Balme 2014); Upper Swan dates to around the same time, 44.77 ± 3.92 ka (Pearce and Barbetti 1981, calibrated in Allen and O'Connell 2014). Despite the more coastal location of these sites, there is no reason to presume that inland areas could not have been occupied this early – it would simply have depended on the local landscape, specifically the distribution of critical resources, particularly water. Early sites in the arid zone elsewhere in Australia are all associated with a reliable water source of some kind (Bird et al. 2016; Smith 2013:78–91). 4.2.1 Archaeological Sites in the Study Area While several archaeological sites have been identified around the study area and in the wider region, little is known about them, and none have been radiometrically dated. A total of 16 registered sites and ‘other heritage places’ (as defined by the Department of Planning, Lands and Heritage – DPLH) are within the study area, excluding those analysed as part of this research (Figure 4.1, Table 4.1); expanding the area would yield more sites but add little comprehensive data. Sites are mostly found on granite outcrops and valley floors around salt lakes, echoing Bird’s (1985) and Smith’s (1993) results (see Chapter 4.2.2). Corsini (1993) noted that he visited several unnamed granite outcrops in the area and found artefacts in most places where water pools after rain. Nevertheless, the number and distribution of known archaeological locations should not be presumed representative of the area as a whole – preservation, survey and reporting biases must be considered. Sites are more likely preserved in landforms unaffected by agriculture, such as granite outcrops and salt lakes. Site locations skew towards areas surveyed for development (roads, pipelines, infrastructure) or places reported by locals/landowners, rather than from the results of systematic research surveys. This likely explains the relatively high number of rock art sites, as these are easily identified by people with no archaeological training. 60 Figure 4.1 Registered sites and Other Heritage Places in the study area. White squares denote approximate site positions, enlarged in inset maps to show their immediate surroundings. Locational data were retrieved from Aboriginal Heritage Inquiry System (AHIS), maintained by the Department of Planning, Lands and Heritage. The central point of sites/place marked on insets. Site details can be found in Table 4.1. 61 Table 4.1 Archaeological features in the study area, excluding those analysed for this research. ID, name and classification taken from DPLH’s Aboriginal Heritage Inquiry System (AHIS), which separates Registered sites (RS) from ‘Other Heritage Places’ (OHP). Unless noted otherwise, information derives solely from DPLH site files. The map code indicates the position of a particular feature or set of features on Figure 4.1— features are numbered from west to east. ID Name DPLH class Map code Archaeological content Reference 16787 Kondinin Pipeline 1 OHP 1 Artefacts Staub and Taylor 1999 18508 Corrigin-Hyden Road OHP 2 Artefacts 16791 Kondinin Pipeline 5 RS 3 Artefacts Staub and Taylor 1999 16788 Kondinin Pipeline 2 RS 4 Artefacts Staub and Taylor 1999 16789 Kondinin Pipeline 3 OHP 5 Artefacts Staub and Taylor 1999 5839 Camel Peaks RS 6 Rock art Davidson 1952 4438 Wave Rock Scarred Tree RS 7 Modified tree, artefacts Corsini 1993 4661 Hippo’s Yawn OHP 7 Rock art 5840 Hyden Rock RS 7 Rock art, artefacts Davidson 1952 Glendenning 2004 Peck 1982 21384 Wave Rock Scarred Tree OHP 7 Modified tree Glendenning 2004 21385 Wave Rock Rockholes OHP 7 Gnammas Glendenning 2004 21386 Wave Rock Isolated Find OHP 7 Artefact Glendenning 2004 21387 Wave Rock RS 7 5844 Hyden OHP 8 Gnamma, artefacts 5610 Graham Rock OHP 9 Gnammas 5611 Hyden-Hyfield RS 10 Rock art Glendenning 2004 The sites listed in Table 4.1 are described below; unless cited, information derives from DPLH site files. The sites clustered on or around Wave Rock are discussed as one unit since they potentially represent a single site complex. No effort was made to visit the listed sites (some no longer exist), rather information is reported as it appears in the site files or heritage survey reports. Kondinin Pipeline 1: A low-density artefact scatter on and around a salt lakebed, with two higher-density clusters, the first measuring 551 m2 yielded 300 artefacts and the second approximately 315 m2 had 90 artefacts. A sample of 20 artefacts 62 comprised 18 quartz – predominantly debris (n=13), with four complete flakes and one broken flake – along with a quartzite flake and flake fragment. Most artefacts were small, < 10 mm in maximum dimension; no pieces measured > 30 mm (Staub and Taylor 1999). Corrigin-Hyden Road: An artefact scatter adjacent to a granite exposure. Six brown bottle glass artefacts (five scrapers and one retouched piece) were found in a 10 x 15 m area. Isolated finds along the same road alignment comprised two quartz flaked pieces and a quartz flake on salt pans/flats, a quartz flaked piece on a creekline, and a worked brown glass piece on the granite itself. Kondinin Pipeline 5: A low-density artefact scatter measuring 85 x 36 m on a lunette, adjacent to a salt lake. Quartz debris measuring 15–30 mm dominated the assemblage. A studied sample comprised eight quartzite artefacts (three flakes, two pieces of debris, plus a single flake fragment, core and broken flake) and 156 quartz artefacts (92 pieces of debris, 23 broken flakes, 22 flakes, 14 flake fragments and five cores); a disturbed area west of the scatter also yielded nine artefacts (Staub and Taylor 1999). Kondinin Pipeline 2: A low-density artefact scatter on and around the bed of a saltlake. Two higher density clusters were found on the western shore: one 25 m2 with 50 artefacts, and a two square-metre area yielding 20 artefacts. Quartz debris dominated both assemblages, mostly ≤ 20 mm in maximum dimension. A sample comprised three quartzite artefacts (a broken flake, a flake and flake fragment) and 27 quartz artefacts (20 pieces of debris, four flake fragments, two flakes and a broken flake) (Staub and Taylor 1999). Nearby, Quartermaine (2000) found three quartz flaked pieces, a quartz flake and a worked glass piece. Kondinin Pipeline 3: An artefact scatter on and around a salt lake. Some artefacts were found on the lakebed, but most were concentrated in a band 2 m wide around the lake edge, approximately 1–2 artefacts/m2. Artefacts were mostly quartz debris, ≤ 20 mm in maximum dimension. A sample of 27 artefacts (all quartz) was classified as nine broken flakes, nine pieces of debris, six flakes, plus a flake fragment, core, and a flake subsequently used as a core (Staub and Taylor 1999). 63 Camel Peaks: A rockshelter on granite outcrop, housing red handstencils and a painted rake-like motif with five tines (Davidson 1952). Wave Rock: Site 21387 does not refer to specific archaeological content, rather encompasses the entire Wave Rock area, which Glendenning’s (2004) informants identified as an important ethnographic site that was used as a camp, hunting place, and water source. Two rock art sites, Hyden Rock (5840) and Hippo’s Yawn (4661), occur on the eastern edge of the Wave Rock outcrop. Hyden Rock comprises two small rockshelters housing red handstencils and a white circle (Davidson 1952); both shelters have since been damaged by fire and/or graffiti, but motifs are still visible (Glendenning 2004; Peck 1982). A few white quartz flakes and a quartzite grindstone were also found nearby. Hippo’s Yawn originally preserved at least three handstencils, but they were inadvertently removed with overlying graffiti; no artefacts are present. Other archaeological features around Wave Rock include an isolated broken quartz flake (site 21386), a series of gnammas (site 21385) and two scarred/modified trees (sites 4438 and 21384), the first of which was associated with 13 quartz pieces, four of which had clearly been flaked. Corsini (1993) surveyed the Wave Rock area and located five artefact scatters and 23 isolated artefacts, totalling 60 quartz pieces (23 of which had definitely been flaked) and a possible hammerstone made on basalt. Only one registered site (4438) could be identified from his limited information. The remaining cultural material he noted may be incorporated within or additional to the sites described above. Hyden: Comprises a deep gnamma (> 1.5 m deep) and some quartz fragments on a granite exposure. Graham Rock: Two gnammas were identified on the granite outcrop, but no artefacts were present on Graham Rock or the neighbouring outcrop, even in rockshelters where sediment had accumulated. Hyden-Hyfield: Paintings (no stencils) located in a breakaway rockshelter, on granite. 64 4.2.2 Other Relevant Research So little archaeological data exists for the study area and its surrounds that it is necessary to go further afield for comparative material. Significant research projects have been conducted in southwestern Australia (e.g. C. Dortch 2000; J. Dortch 2000, 2004; Ferguson 1985; Hallam 1972, 1975), but they are focussed on the coastal plain or the heavily-forested southwestern corner. Due to the vastly different environment of the study area – specifically the comparative scarcity of water – occupation intensity and mobility patterns would not be comparable, and hence a different archaeological signature should be expected. Therefore, the only studies of relevance to this research are those that include an inland component, in areas with similar climate, geology and/or hydrology (e.g. Anderson 1984; Bird 1985; Smith 1993). Figure 4.2 Locations mentioned in the text. 65 Swan Coastal Plain–Darling Plateau Anderson (1984) investigated how Aboriginal occupation varied between three distinct landscapes: the Swan Coastal Plain, the jarrah forests immediately east of the Darling Scarp, and the Darling Plateau, up to 180 km from the present coastline, near Yealering (Figure 4.2). She analysed data from almost 450 sites and found that the size of sites (based on the number of artefacts found, excluding amorphous debris) and artefact densities varied significantly between the three environmental zones. On the coastal plain, ‘major’ sites (with > 500 artefacts) were more common and sites generally had more artefacts than those in the other zones. Assemblages included a wider range of artefact and raw material types, as well as grinding stones and greater amounts of debitage (Anderson 1984:34). In contrast, no major sites were found in the jarrah forests, where minor (< 50 artefacts) sites instead dominated, but a few intermediate sites (50–500 artefacts) were also present. Inland sites (all near water sources) fell between these two extremes – major sites were rare but present, and artefact numbers were consistently lower than at coastal sites. Artefact assemblages contained a more restricted range of raw materials, and were quartz-dominated, but dolerite was also fairly common (Anderson 1984:19); more detailed artefact data were not provided for inland sites. Anderson (1984:34, 37) related differential site distribution to resource availability and human mobility and concluded that the Swan Coastal Plain and the inland Darling Plateau were both inhabited by distinct groups, while the forested area acted as a transitional zone between them. She argued that the people who lived on the coastal plain had a regular seasonal round. They aggregated on the coast, or around larger waterbodies during summer and autumn, to exploit the wealth of aquatic resources. Food was scarcer in winter, so people broke into smaller groups and some moved into the forested area – explaining the predominance of small sites there – while the others ranged more widely among the plain; by late spring, coastward movement recommenced. In contrast, the more limited resource base meant that groups living on the Darling Plateau had a less defined seasonal round. These groups were more mobile, so ranged over longer distances, but also utilised the forest area more heavily as its swamps held water for longer periods. Due to the general lack of dated sites, Anderson (1984) did not discuss how long these occupation patterns had been in place. 66 Bremer Bay–Lake Grace Bird (1985) studied a transect running from Bremer Bay to Lake Grace (Figure 4.2), incorporating 180 sites across four environmental zones: coast (n=37), coastal plain (n=32), intermediate (n=58) and inland salt lake (n=53). Her research aim focussed on lithic utilisation strategies, specifically the reduction intensity of chert, which was of limited availability compared to the more ubiquitous quartz. Like Bamforth (1986), she demonstrated that chert was conserved as distance from source increased, while quartz reduction was largely unchanged. While many of Bird’s conclusions are largely irrelevant here, her site distribution data – specifically the sizeable inland sample – warrants discussion. Bird (1985:96, 150) classified surface artefact scatters based on artefact density, allowing for the size of the area over which artefacts were distributed. Her study area yielded 23 major sites (> 1 artefact/m2), 90 minor sites (< 1 artefact/m2), 12 diffuse scatters (< 1 artefact/m2 but over 100,000 m2 in area), 33 isolated artefacts and 22 other sites (burials, quarries, stone arrangements and shell scatters). She argued that her study area was populated at a density of four people per 100 square kilometres (Bird 1985:129). Based on those data, and the assumption that water availability constrained patterns of aggregation and dispersal, Bird (1985:176) proposed the following seasonal route for her study area: o Feb–May: People would cluster in a few large camps, based around permanent, reliable water sources (gnammas, soaks, coastal dunes); o June–September: People would disperse into smaller groups and use more ephemeral water sources (rock pools, areas with standing water after winter rains); o October–January: Camps varied in size, with people using intermediate water sources (soaks, small swamps, etc.). Bird (1985:150) identified 53 sites in the Inland Salt Lake zone: 30 minor, nine major, six diffuse scatters, seven isolated artefacts, and one quarry/stone arrangement. Based on geomorphological evidence, she considered that surface artefact scatters probably post-dated 5000 BP, but none were directly dated (Bird 1985:115). Detailed artefact data are unavailable, but the assemblages were heavily quartz-dominated, with small quantities of chert and silcrete (Bird 1985:340–344). Most sites were in open Eucalypt woodland on valley floors near salt lakes – just one granite outcrop 67 site was represented (Bird 1985:161, 163). Nevertheless, she cautioned that this does not necessarily represent occupation patterns but rather sampling bias, as she did not record a representative sample of sites, but instead initially targeted those areas with known or potential water sources, or where artefacts had been reported previously (Bird 1995:96). Sites were more common in the southern part of the Inland Salt Lake zone, where freshwater lakes and swamps occur (Bird 1985:131, 162). Nevertheless, Bird (1985:161) noted that there was no strong association between site locations and water sources. Indeed, based on Bird’s (1985:340–344) data, the average distance between a site and a known freshwater source was 2.4 km. People generally camp within 100 m of freshwater (e.g. Gould 1968), but only 13 of Bird’s 53 inland sites were this close to known water sources. It is possible, then, that the bulk of the sites represent occupation when surface water was abundant (during winter), as people dispersed into smaller groups at this time; this may explain the predominance of minor sites. Sites closer to water sources may have been used between October and May, with the timing and length of occupation depending on the reliability of individual water sources. Four major sites were within 100 m of known freshwater sources. These may represent people clustering there at times when other water was scarce, but this idea cannot be evaluated using Bird’s (1985) data. Southern coast–Mount Ridley Smith (1993, 2011) conducted a regional analysis along a transect of the southern coast between Esperance and Israelite Bay, stretching inland as far as Mount Ridley and Mount Ragged (Figure 4.2). The area incorporated three environmental zones: offshore islands, coast and the interior. She surveyed approximately 10% of her study area and recorded 300 sites and analysed the data from 217 of them, of which 215 yielded artefacts (Smith 1993:223; Smith 2011). Her systematic survey methodology and deliberate focus on surveying all landscape types equally meant that site distribution could be used as a proxy for settlement patterns (Smith 1993:211). Across all three zones, sites were most common on granite outcrops, where artefact densities were correspondingly higher (Smith 1993:228). She related the popularity of granite outcrops to their water catchment properties, the availability of shelter, diversity and predictability of plant foods, and use of high-points as 68 lookouts to monitor weather patterns as well as the movement of people and game (Smith 1978, 1993:237). Throughout her study area, most sites were 500–800 m from water sources (known or potential) and, unsurprisingly, artefact scatters were generally found within 100 m of gnammas (Smith 2011). It is important to note that while Smith (1993:229) assumed all salt lakes would have access to soaks of drinkable water, this is not always the case, as Roe (1836, 1852) found. Of the 193 artefact scatters she recorded, 86% (n=167) had < 350 artefacts, including debris; only four scatters, all on the coast, yielded > 1000 artefacts (Smith 2011). Based on the consistently limited spatial extent of her sites, their regular distribution, and low artefact counts, Smith (1993:287–288) argued that her study area was occupied by small, highly mobile groups, and that their mobility was likely a response to resource scarcity and unpredictability. The few sites with more artefacts may represent areas where people congregated, possibly with people from other cultural groups. The inland zone sample comprised 50 artefact sites and two quarries, an average of one site per 6.25 km2 (Smith 1993:270–272). Nearly 60% of sites were on/around granite exposures, but salt lakes, rivers/swamps and sandplains were also often associated with archaeological material (Table 4.2). Most sites were adjacent to thick vegetation, in which people could take shelter if required (Smith 1993:281). No detailed artefact data are available for the interior sites, but most (n=46, 88%) yielded < 350 artefacts and 21 had < 100 artefacts, echoing the small-site pattern evident across her entire study area. Larger assemblages were restricted to granite and salt lakes, with the former yielding the only sites with 501–1000 artefacts (Table 4.2). Overall, sites on granite accounted for 75% of artefacts discarded in the interior, followed by salt lakes (15%) and watercourses (7%) (Smith 1993:271–272). One interior site, Mount Ridley, also has rock art, including hand stencils, bird tracks, and complex linear designs; most of the nearly 1000 artefacts were discarded well away from the rock art. Smith (1993:276) said her informants indicated that Mount Ridley was a meeting place, where people from Esperance gathered with their neighbours from the north. Mitchell’s (2016) analysis supported this idea – he identified silcrete sourced from Ngadju country, to the north, and evidence of intrasite spatial organisation that Conkey (1980) considered an indicator of aggregation. 69 Mitchell (2016:172) also argued that evidence of both Place and Individual Provisioning indicated repeated use of Mount Ridley over time combined with longrange travel, both of which would be expected at an aggregation site. Table 4.2 Archaeological data from Smith’s (1993:271) interior zone, showing the number of sites in different topographic units, and the number of sites falling within different assemblage size categories. Topography Number of artefacts TOTAL 0–30 31–99 100–350 351–500 501–1000 Granite 4 4 18 2 2 30 Salt lake 2 2 1 2 0 7 Rivers/swamp 2 2 2 0 0 6 Sandplain 4 1 2 0 0 7 Dune/gravel 1 0 0 0 0 1 Chert ridge 0 1 0 0 0 1 TOTAL 13 10 23 4 2 52 4.3 ETHNOHISTORICAL BACKGROUND In 1826 the British founded Albany, and the Swan River colony (now Perth) followed in 1829 (Stannage 1981). He was not the first to head inland in search of agricultural land, but Surveyor General J. S. Roe (1852) appears to be the first European who ventured near the study area – he passed through Karlgarin (Figure 4.2) and just south of Hyden during an extensive trip southeast of Perth in 1848–1849. In the 1860s, pastoral stations were established at Narembeen and Kellerberrin (Leake 1950). Holland's Track was cut in 1893, and runs from Broomehill to Coolgardie, passing through King Rocks approximately 30 km north-east of Hyden. Europeans were in the study area soon after: the words ‘M. Cronin 1894’ were painted inside Mulka’s Cave, but were subsequently removed (Randolph 1973). While the surrounding area was farmed from the 1920s, the township of Hyden was not gazetted until 1932 (Landgate 2015). Ethnographic and historical documents may provide a wealth of information but, as McBryde (1979) and Trigger (1982) noted, these sources can be problematic due to biases held by observers and/or authors, as well as their methods of collecting data. Several early documents mention the Indigenous inhabitants of southwestern Australia, but generally focus on information useful to future settlers/explorers, such as Aboriginal names for specific water sources, or ‘exotic’ subjects like initiation 70 ceremonies, burial rites, kinship systems, spiritual beliefs, and tribal warfare. These do not present a balanced picture of Aboriginal life (Green 2011; McBryde 1979). Furthermore, sources make no allowance for the impact of introduced disease on traditional life (Briscoe 2003; Campbell 2002; Cleland 1928; Dowling 1997; Green 1981) or the disruption of traditional occupation patterns by colonial settlement and agriculture (Host and Owen 2009; Neville 1948; Radcliffe-Brown 1930). The sources must therefore be treated with caution. More recent ethnographic information is also available, predominantly from cultural heritage surveys. Under Section 18 of the Western Australian Aboriginal Heritage Act (1972), any development must be preceded by a cultural heritage survey, involving Aboriginal people with a demonstrated connection to the area. O’Connor et al. (1989:54) argued that appropriate spokespeople for specific areas should meet at least one of the following criteria: 1. Long term (normally childhood) association with one or more members of the ‘pivotal generation of culture transmitters’ – individuals who had, as children themselves, been exposed to people who passed on their traditional knowledge; 2. Demonstrated knowledge of an area’s natural resources, including food plant and animal locations, hunting grounds, camping areas and water sources; 3. Politically active individuals, who have worked hard to amass traditional cultural knowledge relevant to a specific area. The information disclosed by such people can also be biased and problematic, as it represents only that portion of traditional knowledge that has survived the cultural dislocation that followed European settlement. Nevertheless, it is a welcome supplement to the fragmentary documentary record. 4.3.1 Aboriginal Social Boundaries Much of southwestern Australia is covered by the South West Native Title Settlement (Figure 4.3); the Traditional Owners of this area all identify as Noongar, but subgroups also exist. Whereas the current native title boundary is rigid for legal and administrative reasons, boundaries in the past were likely more fluid, both spatially 71 and temporally. It is generally accepted that, before European arrival, the southwest was inhabited by a group of people with similar languages and customs. They shared a common initiation rite, scarification, and did not practice circumcision or sub-incision, distinguishing them from the eastern groups who did (Bates 1938; Curr 1886, 1887; Tindale 1974; White 1985). The boundary that separated circumcising from non-circumcising groups was arguably not a static border but rather a westward moving frontier before its impetus was arrested by European settlement (Gibbs and Veth 2002). In all but Curr’s (1886, 1887) system, the study area lies within the noncircumcising group (Figure 4.4), but local group names and territories vary. The occupants of the study area have been identified as Kikkar (Davidson 1938), EastOf-Perth tribe (Hammond 1933 – though his geographical boundaries are vague), and Njakinjaki (Tindale 1974). Hyden appears to be on the boundary for Curr’s (1886, 1887) Kokor group, but he noted no adjacent groups. While the inhabitants of the study area were evidently Noongar, proximity to the (moving) circumcision boundary means that desert influences were likely active. Social gatherings and visits were common both within and between tribes (Hammond 1933), and groups were permitted into other people’s lands provided correct protocols were observed (Myers 1982). Mechanisms for cultural exchange were certainly in place, so each tribe/group should not be viewed as isolated entities. Figure 4.3 Area covered by South West Native Title Settlement, showing subdivisions (Source: National Native Title Tribunal). ILUA = Indigenous Land Use Agreement. 72 A B C D Figure 4.4 Aboriginal social boundaries in southwestern Australia according to various sources (redrawn from original maps by A.M. Rossi). A: Curr (1886,1887) separated groups into the noncircumcising Western Division (green) and the circumcising Central Division (beige). Group names are given in italics, where known. Red crosses indicate locations where circumcision was practised, but recorded locations are often vague and their placement on the map may be somewhat arbitrary. E.g. the cross within Kokor territory is noted in text as ‘east of York’. Original image provided courtesy of Fisher Rare and Special Collections, University of Sydney. B: Bates (White 1985) placed the division for her non-circumcising group (the Bibbulmum) further inland. Her eastern boundary locations are shown in blue. At the time of her research, Karlgarin was the most inland town established in the vicinity of the study area, so was the only real point of reference. The western boundary for the circumcising group (Karatjibbin – specific locations shown in red) is considerably further east. C: Davidson (1938). D: Tindale (1974). Groups east of his circumcision/subincision boundary are not shown. 4.3.2 Ethnohistory of the Study Area Ethnohistorical information for the study area is limited, as historical records of inland travels barely mention Aboriginal people. For example, while Roe (1836) encountered Aboriginal people around Narembeen and Mt Hampton (Figure 4.2), he gave few details of the encounter. Most accounts are dominated by the search for water, and in what locations and quantities it was eventually encountered – 73 unsurprising since several exploration parties nearly came to disastrous ends for want of water (e.g. Roe 1852). When Landor and Lefroy (1843) were travelling between Dumbleyung and Lake Grace, their guides refused to venture further east (closer to the study area), ostensibly due to lack of food and water. This perceived harshness of the landscape may explain why early settlers believed that Aboriginal people rarely, if at all, visited the study area, and merely passed through on their way to better lands (Meeking 1979; Mouritz 1986). However, Landor’s and Lefroy’s (1843) experience may actually reflect their guides’ unwillingness to travel beyond their own lands to an area where they may be less familiar with the location of food and water sources, creating significant risk (Hiscock 1994). Indeed, Landor and Lefroy turned northwards and shortly found the land far more hospitable, indicating the variability of inland environs. Hammond (1933:21, 23) stated that people from the ‘eastern regions’ dealt with resource scarcity and variability by splitting into small groups and moving around frequently; similar strategies are predicted under Optimal Foraging Theory (see Chapter 2.2). During his inland journey, Roe (1836) only encountered Aboriginal people travelling in small groups of up to ten individuals or noted the remains of a single hut; around present-day Narembeen he identified the remains of five huts clustered around a large water hole. The more limited visibility associated with small, highly mobile groups may also have contributed to early settlers’ belief that Indigenous people did not systematically occupy the study area. There is some evidence that ceremonial and social events may have occurred in the study area, including at Wave Rock: We call Wave Rock ‘Kutter Kich’. This place is very important to us. This is on the boundary where a lot of people used to meet for ceremonies. Our old people walked the trail from the nanuk (neck) of the river at Guildford to Kutter Kich for thousands and thousands of years. (Winmar 1996:28). This suggests that the study area may have functioned as a place where occupants could meet with their western neighbours, and perhaps also with desert groups. A member of the Kalaako group, the Njaki Njaki’s closest neighbour from across the circumcision boundary, indicated that his people had ceremonial and trade relationships with their western neighbours, but that this contact was infrequent 74 (Macintyre et al. 1992). He did not state in whose territory these meetings occurred, but similar inter-group aggregations have been identified in the archaeological record (Mitchell 2016). If these meetings also included groups from nearer to the coast, the study area would have provided a convenient intermediate meeting place between the west coast and the inland desert areas. The study area also apparently contained sites of mythological significance, which may have been taboo areas for women and children; these mythological sites were on or near granite outcrops (Macintyre et al. 1992). One prominent informant noted that Mulka’s Cave at The Humps was likely a significant ceremonial or mythological site (Macintyre et al. 1992); another individual noted that when he set up camp near Mulka’s Cave, elders told him to move further away as he should not stop there (Goode 2011). These sites may have been used in conjunction with, or separately from, the social gatherings noted above. 4.4 CONCLUSION While ethnohistorical sources are limited, they indicate that the study area people were part of the larger southwestern group, the Noongar, who marked initiation by scarification, not circumcision. People living in the study area had contact with their neighbours from as far away as the west coast, as well as those from the circumcising desert groups (Macintrye et al. 1992; Winmar 1996). Gatherings with their neighbours possibly occurred at Wave Rock, and other granite outcrops in the area may have had ceremonial or mythological significance; neither claim can be addressed with the current suite of archaeological data. Due to the harsh environment of the study area, its occupants were probably highly mobile and travelled in small groups, moving across the landscape according to local resource availability. This suggestion is supported by the archaeological data from the study area and its immediate surrounds, as well as other parts of inland southwestern Australia that is dominated by small sites focussed on and around granite outcrops or salt lakes. While no sites in the area have been dated radiometrically, Quartermaine (2000) found flaked bottle glass, indicating that Aboriginal people were using the area after Europeans arrived in Western Australia, contrary to the beliefs of early settlers (Meeking 1979; Mouritz 1986). 75 CHAPTER 5 CONSTRUCTING THE RESOURCE MODEL 5.1 INTRODUCTION This chapter details the methods used to construct the resource model for the study area, beginning with the means of expressing spatial variability in resource distribution by creating of Landscape Divisions defined by vegetation and landform characteristics. The process for compiling floral and faunal lists is then outlined, including the identification of food plants and animals, and the means of determining their seasonal availability and distribution across the landscape. The procedure for quantifying the longevity of rainwater preserved within soil profiles and rock structures is then discussed, allowing for the highly variable rainfall patterns evident in the climatic data. Finally, the methods of constructing the occupation model are explained, focussing on the ranking system used to classify plant foods. 5.2 LANDSCAPE DIVISIONS The study area is small, but there is no reason to assume that food and water resources would be uniformly distributed across its landscape. It was therefore necessary to devise some means of representing variation at a manageable level. This is commonly achieved by dividing a landscape into different units; this approach is employed in archaeological research (e.g. Bird 1985; Smith 1993) and agricultural land management (e.g. Murphy-White 2007; Murphy-White and Hauck 1999). The precise divisions formulated depend on the size of the study area, as well as the aims of specific studies. For this research, the study area has been divided into distinct landscape types – referred to hereafter as Landscape Divisions – based on landform and vegetation characteristics recorded by Beard (1972, 1979, 1980). His data were selected since they represent the pre-European distribution of native vegetation, and extensive spatial data are easily accessible via the NRInfo web application (DPIRD 2016b). Six Landscape Divisions were recognised: Granite, Heath, Mallee, Saline, Thicket and Woodland – these names are capitalised to distinguish them from generic vegetation communities or landform characteristics. Most incorporate a single vegetation community, but the Saline division comprises 76 several that occur on naturally saline soil, as well as salt-lakes themselves (Table 5.1). Mosaic units were classified based on the dominant vegetation form or, where neither was dominant (e.g. scrub-heath/thicket), the more diverse assemblage. The distribution and character of each Landscape Division are shown in Figures 5.1–5.2. Vegetation communities are dictated by soil, topography and hydrology, so the Landscape Division unit should capture and express variation in all pertinent resources including food plants, vertebrate fauna, and water; sites within the same Landscape Division are therefore presumed to share identical resource bases. This assumption of homogeneity is an oversimplification, albeit commonly employed (e.g. Smith 1993:203), and it is important to note that variation will occur at a much smaller scale than can be depicted in Figure 5.2. Indeed, most vegetation types grade into one another so while firm boundaries are delineated, they are somewhat artificial. Nevertheless, each Landscape Division should accurately characterise the dominant vegetation and landform of a given area, which facilitates analysis at a broad, landscape level (determining how the entire study area could have been used) as well as permitting site-specific analysis of resource abundance. The latter is especially significant, as it allows all archaeological sites in the study area – even those for which scant data are available (e.g. those merely listed on the Aboriginal Heritage register) – to be better contextualised by an understanding of the local resource base and the impact it has on occupation patterns. Table 5.1 Landscape Divisions and their associated vegetation communities (see descriptions in Table 3.2 and distribution in Figure 3.14) and landform. Vegetation communities are defined following Beard (1972, 1975, 1976, 1979, 1980). Landscape Division Vegetation community / landform Woodland Woodland Mallee Mallee Heath Scrub-Heath Thicket Broombush Thicket Granite Granite outcrops Saline Melaleuca Thicket Succulent steppe Tea-tree scrub Salt lakes 77 A B Figure removed due to copyright restrictions C D E F G H Figure 5.1 Vegetation and landform characteristic of each Landscape Division – all photographs and species information from Beard (1990:119, 121, 123, 131–132, 134, 136). A–B: Saline, dominated by short, salt-tolerant shrubs (Atriplex spp., Tecticornia spp.). Note the taller Eucalyptus sp. trees on slightly higher ground in B. C: Eucalyptus spp. Woodland, note the generally open nature of the canopy and the relatively sparse understorey. D: Mallee, note the multi-stemmed Eucalypts and the thicker understorey than in C. E–F: Heath, dominated by a diverse range of low shrubs including Grevillea spp. and Verticordia spp. Note the Mallee in the background of E, indicating the mosaic nature of Landscape Divisions. G: Thicket of Acacia, Allocasuarina and Melaleuca species – note the closely spaced shrubs. H: Granite. Note the thicker vegetation at the base of the slope, due to water runoff. The rock is mostly bare except where pockets of soil support shrub and small tree growth. 78 Figure 5.2 Distribution of Landscape Divisions within the study area – note that the Saline Landscape Division comprises both saline valleys and salt lakes. Spatial data extracted from NRInfo (DPIRD 2016b). Examples of each Landscape Division are shown in Figure 5.1. 79 5.3 QUANTIFYING PLANT AND ANIMAL FOODS The resource base undoubtedly affects when and how humans can occupy a given area. It is well-known that Aboriginal people generally had a highly mobile lifestyle, and frequently moved around in response to shifts in resource abundance (e.g. Clarke 2009). It is therefore important not to simply analyse a list of foods available in a given area, but to consider their seasonal availability and varying distribution across the landscape. This is not a new concept; various studies have outlined the seasonal food availability in southern Western Australia. Meagher (1974) compiled the first comprehensive list of food sources – both plant and animal – exploited by Aboriginal people in southwestern Australia by conducting an extensive review of historical sources. Aside from identifying the genera or species described in various sources, she also noted the month or season in which the document was written, or the observation made, as an indication of the season during which a particular food was exploited. Others have attempted similar seasonal analyses for southern Western Australia (e.g. Bird 1985; Smith 1993; Webb 2000), but Meagher's (1974) study remains the most detailed. In contrast, few have considered the spatial variation resource distribution, either neglecting this factor entirely, or considering it in a more superficial manner. Bird (1985) simply noted those food types that would be restricted to one or more of her environmental zones – coast, coastal plain, intermediate and inland salt lake – without providing a comprehensive list of all available foods. She noted the preferred habitat of some plant food species (Bird 1985: Appendix A – Table 6), but this information was not integrated into her analysis. Similarly, while Smith (1993) specified the vegetation communities near each of her sites – and assumed that each represented a distinct resource zone – she did not quantify the resource beyond identifying those communities in which some targeted food species occurred (Smith 1993:101). While the aim of this study is not unique – to quantify the spatio-temporal availability of plant and animal foods – the approach differs from those described above in its focus and scale. The intention here was not only to identify the edible plant and animal species present in the study area, but to produce a finer grained analysis by defining the distribution of every exploited species across each of the six Landscape Divisions, considering the seasonal availability of all associated food items. Therefore, this research sought to quantify the full extent of spatial variation, rather 80 than merely assuming its presence. Similarly, previous approaches focussed on records of the season in which different food products were used by Aboriginal people, while the method used here relied on organism-specific information to determine when each plant or animal food was available. While there is some overlap between these measures, the period of use does not necessarily represent the entire window over which a food is available. Indeed, the period of use or nonuse may vary between areas, depending on the relative abundance or dearth of other food items available at a particular time. For the same reason, taboo foods were not considered, and indeed there are no relevant data available for the study area. The methodology employed, outlined in more detail below, was optimised to produce high-resolution data illustrating spatio-temporal variation in food resource availability, since this would undoubtedly have placed constraints on human occupation of the study area. 5.3.1 Compiling Floral and Faunal Lists Before food sources could be analysed, a list of the plant and animal species found within each Landscape Division was compiled; invertebrate fauna were omitted as data were lacking. Much of the study area has been cleared for agriculture, but the various surveys documenting the remnant native flora and vertebrate fauna in uncleared areas provided a good source of data. A total of 33 sites (often very large nature reserves) provided usable data, and most incorporated several vegetation communities. Beard’s (1972, 1979, 1980) larger scale studies were also of value since they described individual formations within the study area. The relevant Landscape Division was identified by comparing vegetation and landscape descriptions to those in Tables 3.2 and 5.1. Figure 5.2 was not used as a classification tool since it merely indicates the dominant vegetation and landscape type and so does not represent variation at a smaller scale. Formations in recently burnt or otherwise disturbed areas were excluded from analysis. Community descriptions were then used to compile species lists for each Landscape Division; floral data were abundant, so vegetation lists were compiled by marking each species against the Landscape Division/s in which it had been observed. Faunal data were more limited, and many reports provided general faunal lists rather than noting which animals were encountered in each vegetation community. 81 Therefore, as well as noting in which Landscape Divisions an animal was observed, they were also matched against records of animals’ habitat preference or targeted food sources (e.g. some birds are attracted to flowers of certain plants). This reduced bias towards those Landscape Divisions overrepresented in more detailed faunal surveys (e.g. McKenzie et al. 1973) and allowed for faunal mobility, as animals may have been observed in just one of their characteristic areas (e.g. Halse et al. 2004). The lists were checked against published and online sources (see list in Appendices A–C) to reflect taxonomic changes and to remove introduced species or native animals observed outside their natural range. Genus-level identifications (e.g. Macropus sp; Eucalyptus spp.) were removed unless no exact species were identified in a particular Landscape Division, or the named species did not represent the diversity of species within a Division. 5.3.2 Identifying and Classifying Food Species To identify potential food species, the floral and faunal lists were compared to published accounts of plants and animals eaten by Noongar people (e.g. Bindon 1996; Bird and Beeck 1988; Meagher 1974 – see complete list in Appendix B). These lists are often based on records of explorers and early settlers, particularly Grey's (1841) diaries, but most only include species that had been named, or those identified by their descriptions, meaning it is almost impossible to compile a list of all plant species utilised (Clarke 2003). This does not affect the faunal assemblage, as Hassell (1975:10) stated that Aboriginal people did not discriminate between animal species and that 'everything was for food'. A list of animals available in an area, then, probably closely resembles those that were eaten. The faunal assemblage was divided into four categories: amphibians, birds, mammals and reptiles, with a subcategory for bird and reptile eggs. Eggs were presumed to occur within the same Landscape Division as the animals themselves. The echidna (Tachyglossus aculeatus) also lays eggs, but they are retained in a pouch, so cannot be encountered independent of the animal; these eggs were not listed separately. For plants, however, there were other factors to consider – primarily whether the plant is edible, inedible or, in certain cases, toxic. Given Clarke's (2003) comment that plant food lists are often incomplete, plants that are known or likely to be edible, such as members of the Orchidaceae family (Pate and Dixon 1982:222), have been 82 included even where there are no recorded instances of them being eaten. Considering Aboriginal people could render some toxic items edible, such as Macrozamia kernels (Beaton 1982; Beck 1992), it is reasonable to assume that any edible plant would have been eaten. Plant species are listed where they have been identified in the literature and grouped by genus where the identification was at genus or family level. Species-level data were prioritised over more general information. Genus-level data (e.g. Acacia gum is edible) was only applied to species that were not individually identified in food lists, under the assumption that if an identified species also provided that food, it would be noted in the literature. Plant foods were divided into seven categories, defined below: o Flowers: flower structure and/or nectar; o Fruit: fleshy fruits and the kernels/seeds; o Gum: resinous substance produced by trees under stress, generally through heat and lack of moisture (Nussinovitch 2010:8) and exuded through injury points; o Leaves/roots/stems: main body of the plant, including stems, leaves and/or fibrous roots (that transport water and nutrients rather than storing them); o Lerps/manna: secretions produced by insects sucking sap containing organic nutrients, produced by photosynthesis (Atkins and Ruan 2010– 2016); o Seeds: all seeds from trees, shrubs and grasses except large kernels contained within fleshy fruits; o Storage organs: Stem or root structures (tubers, bulbs, corms, etc.) used to store nutrients for prolonged periods, usually below the ground surface. Some species have perennial storage organs, while others consume the organ to reshoot following a summer dormancy, and spend the growing season refilling and replenishing it (Brown et al. 2008:7; Pate and Dixon 1982). This affects the period over which certain storage organs would be available. 5.3.3 Determining Seasonal Availability The seasonal availability of plant foods and animals is dictated by climate, or the biology, life cycle and habits of an organism (Table 5.2). Seasonal availability, (rather 83 than season of use), was defined using published sources in the first instance, prioritising data relating to the study area or southwestern Australia over more generalised information. Terrestrial animals were present year-round, while migratory birds were seasonal; eggs were available during the relevant bird and reptile breeding seasons. Plant food availability was more difficult to determine, as while the plant itself may be present year-round, the edible product may not. Availability was determined from species-level information and, when that was lacking, data were sourced from the same genus. Occasionally, generalisations had to be made, relating the timing of a particular item’s availability to the flowering period, as the latter data were abundant (e.g. Paczkowska and Chapman 2000; Western Australian Herbarium 1998-). The factors governing the availability of all food items are shown in Table 5.2. Availability was analysed in monthly intervals. Foods were available if present for most of a given month. To avoid misrepresentative inflation, two caveats were imposed. First, only the main period of availability was noted. For example, plants may have a secondary flowering period that occurs after heavy summer rains, and animals may have a similarly irregular breeding cycle under unusual climatic conditions. These periods were omitted in favour of the regular annual cycle. Second, some plant foods may be available in small quantities throughout the year, but in significantly greater abundance during a particular season, especially when availability is dictated by climatic variables. In those cases, the period of availability has been defined as the period of increased production. This may underrepresent the diversity of food sources at certain times of year, but the small amounts available were unlikely to influence occupation patterns in the same way that a seasonal bounty would. Thus, the period of availability may define the only period during which a food was available, or the period when abundance greatly exceeded that at any other time of year. Periods of heightened value (e.g. tubers with high water content in summer, macropods in better condition through winter) have been identified as they may affect the timing of resource exploitation. 84 Table 5.2 Factors governing seasonal availability for plant and animal foods, and generalisations used to determine periods of availability. Where no governing factors were present the items would be available year-round. Generalisations were employed where data were lacking. For details, see Appendices B–C. Category Plants Animals Availability Flowers o As per individual genus/species. Fruit o As per individual genus/species and/or: • Unripe: two months after flowering; • Ripe: three months after flowering; • Naturally dried (e.g. on tree or ground): one month after fruiting period ends. Gum o Periods of high temperature and limited moisture. Leaves/roots/stems o As per individual genus/species. Lerps/manna o Periods of increased photosynthesis (higher temperature, more sunlight hours). Seeds o As per individual genus/species and/or: • Unripe seeds (tree/large woody shrubs): one month before ripe seeds or two months after flowering; • Ripe seeds (small shrubs, herbs and grasses): one month after flowering. Storage organs o Perennial organs: year-round; o Annual organs: dormant period plus a month either side, when near maximum size. Amphibians o Year-round. Birds o Non-migratory: year-round; o Migratory: as per individual migration patterns. Bird eggs o Non-migratory: breeding season; o Migratory: breeding season, breeding location preferences and timing of presence in study area. Mammals o Year-round. Reptiles o Year-round. Reptile eggs o Breeding season. 5.3.4 Benefits and Limitations of the Approach The floral and faunal lists (Appendices A–C) were integral to evaluating the spatiotemporal availability of plant and animal foods, but their limitations must be acknowledged. Beard (1972, 1979) attempted to quantify the pre-European distribution of vegetation communities – data that formed the backbone for the Landscape Divisions used herein – while all floral and faunal surveys recorded the modern distribution of plant and animal species. This approach therefore applied 85 modern species data to pre-European spatial data, but it is well-known that European farming and introduced species have been detrimental to the native flora and fauna (Abbott 2002, 2008; Benson 1991; Bickford and Gell 2005; Dickman 1996; Witt et al. 2006; Woinarski et al. 2015 – see also Chapter 3.6). Some effects can be ameliorated: obviously alien species, or those outside their natural range, were omitted from the lists here, but some may remain. Similarly, extinct species were not represented; even if the relevant species could be identified, data were too scarce to assign each to the relevant Landscape Division/s. These omissions will largely affect the faunal assemblage, as floral species are more likely to endure (see Chapter 3.5.1 and 3.6). Finally, while all reasonable measures were taken to ensure that floral and faunal species could be confidently assigned to a Landscape Division, there is no reason to assume that all the species listed will be represented in every iteration of a particular Landscape Division. The lists are merely indicative of species that may be encountered in a particular area. In many cases, animals should not be considered restricted to any particular Landscape Division – they may merely be more commonly found in certain areas. However, mobile resources are less likely to impact site patterning (Bird 1985:131), so any introduced bias should not have serious implications. Despite the limitations noted above, this approach has several important benefits: it makes use of the abundant data created as a by-product of agriculture and maximises its application. While a number of locations contributed floral and faunal data, they were extended by considering each vegetation community separately, capitalising on the fact that vegetation is heavily mosaicked at a smaller scale; each site therefore yielded data that could be applied to multiple Landscape Divisions. This increased sample size meant that local data were sufficient and should address an issue identified by Gibson et al. (2004), whereby Beard's structural vegetation units do not correlate with species composition at a larger scale. It also permitted the compilation of more comprehensive floral and faunal lists for each Landscape Division, facilitating a higher resolution analysis that quantified the full extent of spatial variation in food resource distribution, in contrast to previous studies (e.g. Bird 1985; Smith 1993). This finer grained analysis allows a more thorough treatment of factors that may affect or dictate occupation patterns. 86 5.4 QUANTIFYING WATER AVAILABILITY Water availability is paramount to human occupation of any area – it cannot be transported as readily as food. As noted in Chapter 3.3, there are no freshwater lakes or rivers in the study area, and the regional groundwater is saline–hypersaline. Small pockets of less saline groundwater can be found around granite outcrops, but these are at depths > 20 m, so wells must be drilled to exploit them (Bettenay and Hingston 1963; Davidson 1977); these sources were inaccessible to the Indigenous occupants of the study area. Some plants yield water but are mostly used while travelling or during extreme drought (Clarke 2011: 80–81); these are not considered further, as they are unlikely to impact occupation patterns. The main source of potable water, therefore, was preserved rainfall, herein separated into what falls or runs onto soil, and that captured by rock structures. This distinction is important, as water will be lost at different rates according to the storage mechanism. This research seeks to quantify the longevity of water in various soil profiles and rock structures, under different rainfall conditions, as the spatio-temporal availability of these sources probably patterned human occupation. 5.4.1 Representing Rainfall Variability All potable water sources in the study area are forms of preserved rainfall, so their availability cannot be evaluated without considering the inter-annual rainfall variability noted previously (see Chapter 3.4.2). Rainfall variation was generalised as low, average or high, to characterise winter rainfall and annual totals from the Hyden meteorological station (BoM ID 010568), for 1929–2017. Category ranges were calculated assuming that values within one standard deviation of the mean represented a fairly average winter (97–177 mm) or annual (256–424 mm) total; values above or below those ranges were classified as high or low, respectively. Rain falling outside the winter period was considered by quantifying isolated falls of 15 mm and above; they can increase the annual total considerably. Falls were classified in 10 mm increments, from 15 mm to ≥ 105 mm. Cumulative falls were not considered as they could not be generalised in the same manner. Indeed, larger cumulative falls often incorporated one or more isolated falls, so were captured elsewhere. Six years were selected, primarily to represent variation in winter rainfall total – as rain is most predictable and reliable during winter – while also taking care 87 to represent some variation in annual rainfall total and the number of isolated falls (Table 5.3). More recent years were prioritised for data quality and completeness, because the Hyden station does not record the post-clearing drop in winter rainfall that occurred in other parts in the Wheatbelt (e.g. Andrich and Imberger 2013). The Hyden station does not record evaporation data and historical evaporation data are not stored in the same way as rainfall data. Therefore, average daily evaporation values were derived from the monthly totals for Kondinin, 60 km west of Hyden (Luke et al. 1987). Table 5.3 Years chosen to represent rainfall variability, showing annual and winter rainfall totals (mm) and classification, and the number of isolated falls > 15 mm, occurring outside winter (from Hyden station, BoM ID 010568). Year Winter rainfall Annual rainfall Total Classification Total Classification Isolated falls ≥ 15 mm 1992 217.1 High 514.1 High 5 1998 178 High 324.6 Average 1 2009 150.1 Average 295.9 Average 2 2010 69.5 Low 142.1 Low 0 2011 126.4 Average 472.4 High 2 2012 84.2 Low 280 Average 2 5.4.2 Soil Water Soil water is defined here as that stored temporarily within the soil profile, rather than that which has entered the groundwater system. While soaks and wells are often mentioned in the archaeological literature, they are generally near-surface expressions of fresh groundwater, or subsurface remnants of seasonal watercourses fed by runoff in sparsely vegetated areas (Brodie et al. 2012; Gould 1968, 1971a, 1984). There are scant references to soil water in the historical literature, but Roe (1836, 1852) recounts several, often unsuccessful, attempts at digging for water in inland southwestern Australia. Nevertheless, the archaeological literature has never attempted to quantify soil water availability, which probably reflects its lower importance due to higher/more reliable rainfall and/or the presence of other significant freshwater sources, such as lakes, rivers, or potable groundwater. The 88 lack of such resources in the study area heightens the importance of soil water, so in order to accurately reconstruct spatio-temporal occupation patterns, the longevity of this source must be quantified. Modelling soil water content The water-holding properties of soils vary due to the interplay of several factors, including soil texture, profile depth, the presence/absence of permeability barriers (e.g. less permeable subsoils, as in duplex soils), and salt content (Hillel 2004; Houser 2003; Moore 2001; Schoknecht and Pathan 2013). These variables influence permeability and hydraulic conductivity – how quickly soils transport water from the ground surface into the profile, and the rate at which this water subsequently moves through the profile – as well as the amount of water a profile can store and the depth at which it is stored. Therefore, modelling soil water generally requires complex mathematical or computer models (see review by Ranatunga et al. 2008), as well as the expertise to correctly apply them. The Soil Water Tool (DPIRD 2016a) differs because, while it is underlain by extensive mathematical theory devised by Ritchie (1972), it was designed to be easily accessible. Using a few simple inputs and local rainfall data, it models the daily Plant Available Water (PAW – Figure 5.3); this tool was used to quantify the availability of soil water in the study area. Figure 5.3 The relative water content of various soil types, showing available and unavailable water (Moore 2001:82). Available water can be extracted by plants, referred to here as Plant Available Water (PAW). Note that a soil may be holding water, even if none is available for plant use (unavailable water); coarser soils (those with larger particle sizes) have lower water-storage capacity than finer soils but require less input before plants are able to extract water. 89 Table 5.4 outlines the specific inputs used in the Soil Water Tool, the most important of which is the crop coefficient, which defines the rate of crop evapotranspiration relative to evaporative potential, as dictated by climatic conditions (Allen et al. 1998). This coefficient varies according to a plant’s developmental stage, but herein a single coefficient was used throughout to simulate mature vegetation. Accurate coefficients are readily available for crops (e.g. Allen et al. 1998) but not for other vegetation, partly due to the greater number of variables such as plant health, density and water availability (Pereira et al. 2015). The crop coefficient results from the interplay of various climate- and organism-specific variables, but a major contributor is the Leaf Area Index (LAI), that quantifies the proportion of ground surface shaded by foliage. LAI is more influential than overall plant size, as small crops with greater LAI values (e.g. lettuce, cabbage) have higher crop coefficients than larger tree-crops (e.g. citrus), where ground-coverage is not as dense (Allen et al. 1998 – Table 12). LAI is generally lower for natural vegetation, especially in arid or semi-arid areas (Allen et al. 1998). Therefore, a conservative coefficient of 0.5 was selected, indicating that the vegetation lost water at around half the rate of atmospheric demand. This coefficient was applied to all the soil profiles (and the Landscape Divisions they represent) since plant size is not the defining factor in rate of water loss. Data from the Soil Water Tool cannot be exported and is only output as a graph. However, daily values are displayed when the mouse pointer is hovered over the appropriate part of the chart (Figure 5.4). These data had to be manually entered into an Excel spreadsheet for analysis; data-entry accuracy was verified by spotchecks and visual comparison of Excel charts and the original output. Analytical methods were simple – quantifying the longevity of soil-water under various rainfall conditions, using each modelled soil profile as a direct representation of a particular Landscape Division. The quantity of Plant Available Water (PAW) was assumed to broadly represented the amount that could be removed from the soil by digging. For the Granite Landscape Division, however, there are significant differences between the modelled sand profile and real-world conditions. In the model, soils receive input only from rain falling directly onto the ground surface. This represents the real-world scenario in most Landscape Divisions, but Granite areas receive additional input from runoff – when water is shed from outcrops onto the surrounding soils. This extra input heightens the importance of accurately determining profile depth, since deeper 90 profiles have greater water storage capacity. The modelled sand profile is 2.5 m deep (Oliver and Robertson 2009), but soils around Granite outcrops are generally only 0.8–1.5 m deep (DPIRD 2016b; Schoknecht & Pathan 2013). Differences likely exist between other model profiles and the Landscape Divisions they represent, but under normal conditions storage capacity invariably exceeds input, rendering them immaterial. Table 5.4 Soil Water Tool (DPIRD 2016a) data fields and inputs used. A soil profile was selected to represent each Landscape Division (provided in parentheses after profile) based on generalised soil characteristics (Beard 1979, 1990; DPIRD 2016b; Murphy-White 2007; Schoknecht and Pathan 2013). Model years were chosen to represent a variety of rainfall quantities and distributions, as per Chapter 5.4.1. The 1992 and 1998 model periods were abridged due to unusually high Plant Available Water (PAW) balances occurring at germination, resulting from model mechanics whereby plants begin to draw on a store of water accumulated under preceding fallow conditions. This does not accurately represent land under native vegetation, as plants would be drawing water year-round. The April break-season rule was voided to ensure that germination occurs during winter, so model periods were more directly comparable. A single crop coefficient was used throughout, to simulate mature native vegetation. Field Description Input Weather station Selected from a list of BoM and DPIRD facilities in the Wheatbelt. o Hyden BoM Station (010568) Soil type Selected from a list of ten idealised soil profiles defined by Oliver and Robertson (2009). o o o o o o Sand (Granite) Sandy Earth (Heath) Deep Sandy Duplex (Mallee) Shallow Loamy Duplex (Saline) Gravel (Thicket) Loamy Earth (Woodland) Date range Maximum 12-month period must be selected if cropped simulations are required. o o o o o o 20/3/1992 – 31/12/1992 20/4/1998 – 31/12/1998 1/1/2009 – 31/12/2009 1/1/2010 – 31/12/2010 1/1/2011 – 31/12/2011 1/1/2012 – 31/12/2012 Break season date Quantity of rainfall required to prompt germination, specified as a value over three days after 25 April and/or after 5 June. o 5 mm of rain over three days after 5 June Crop coefficient Rate of evapotranspiration relative to atmospheric demand. Separate values can be provided for each of four recognised growth stages: initialisation, development, mid-stage, late-stage. o 0.5 (throughout) 91 Figure 5.4 Example output from the Soil Water Tool (DPIRD 2016a), showing soil water (PAW) under fallow (black line) and cropped conditions (green line), plus daily rainfall (blue bar). Fallow and cropped values will be in-sync until germination date ('break'); thereafter, modelled crops draw on stored water accumulated under preceding fallow conditions. An increase in the PAW balance is registered only when daily input (rainfall) exceeds daily loss (through evaporation/evapotranspiration). Daily data values (box) are displayed when the mouse is hovered over the relevant point on the chart output (black filled circle). Special methods had to be devised to model PAW in Granite soils due to the additional input and shallower profile depth, as noted above. The precise amount of additional input varies due to the individual morphology of each outcrop, as well as rainfall characteristics (Fernie 1930; Laing and Hauck 1997). For consistency, it was assumed that Granite soils received double the input that other areas did, comprising direct input (rainfall) plus the same amount again through runoff. On average, sands can store 90 mm of water per vertical metre of deposit. A Granite profile 1.15 m in depth (the mid-point between the 0.8–1.5 m depth range for Granite soils) could hold approximately 100 mm of water. To calculate the daily PAW balance of Granite soils, it was necessary to determine the starting balance and the subsequent rate of water loss. These measures are outlined below: 92 o Starting balance: the quantity of water held in a Granite profile at germination, or after the profile re-wets. This was calculated by adding the PAW balance of the sand profile – which incorporates the daily rainfall minus the quantity of water required to recharge the profile and the daily loss – to the rainfall amount again, to simulate additional input through runoff; o Daily losses: the amount of water lost from a Granite profile on a daily basis, through evapotranspiration. Daily losses were calculated for each day that the sand profile held water (since the losses will be identical for Granite profiles) across all modelled years and used to calculate an average daily loss value for each month. Sand profiles never held water in October or November, so these values had to be interpolated based on those from other months. Beginning with the Granite starting balance, subsequent daily PAW balances were determined by doubling daily rainfall and applying the observed daily loss (from the sand profile) or, where that value was unavailable, the relevant average daily loss value. PAW balances in the Granite profile were not permitted to exceed 100 mm, to reflect the limited profile depth. Due to the methods employed, a starting balance could only be calculated when the sand profile responded to a rainfall event, but Granite would likely respond when sand did not due to the impact of runoff. These responses could not be modelled, but could be estimated based on the rainfall amount, drying time, and the quantity of water required to recharge a profile devoid of water (see Chapter 7.2.2). Benefits and limitations of the approach It is important to acknowledge the limitations of this approach, both those inherent in the Soil Water Tool, and those arising from the application of model results to address the research aims. The Soil Water Tool does not allow for certain barriers to water infiltration (water-repellence), sodicity (salt content), or other characteristics that may impact water-storage capacity. However, these variables are minor compared to those encompassed within the model, so the omissions are unlikely to alter the results appreciably. The main limitations concern the interpretation of the 93 results, in terms of how applicable specific values are to the real-world scenarios this research aimed to replicate. For example, it was assumed that any water that can be extracted by plants (PAW) would also be accessible to humans by digging, but plants may be more efficient at removing water from the soil; there are no data permitting this idea to be evaluated. The crop coefficient applied (0.5) may also introduce errors – at best, it is an accurate indication of the longevity of water under native vegetation and at worst, merely a measure of how long various soils provide water to a crop transpiring at half the rate of atmospheric demand. In either case, the patterns are broadly representative (i.e. the relative longevity of PAW in different soils in different years) but specific values such as dates, quantities and duration should be accepted with caution, particularly for Granite areas, where model data was not used in its original form. Finally, there is no reason to expect soil and vegetation characteristics to be consistent throughout a single Landscape Division. Pereira et al. (2015) noted that natural vegetation coefficients rarely represent even a minority of scenarios, due to the variables noted above. Nevertheless, the assumption of intra-Division homogeneity is inherent to the use of Landscape Divisions as analytical units. Where considerable intra-Division variation was conceivable, this was factored into interpretations. The benefits of using the Soil Water Tool considerably outweigh the limitations of the approach used here. The Tool is underlain by extensive theory and mathematical models, is tailored to Western Australian soil and rainfall conditions, and provides predictions and calculations that would otherwise be inaccessible to those unfamiliar with soil physics. It permits a far more robust analysis of the spatio-temporal availability of soil water than would otherwise be possible by analysing basic soil characteristics. The data permit the establishment of firm timeframes for the longevity of soil water in different Landscape Divisions, even if some specifics should be treated with caution. 5.4.3 Rock Structures Herein, rock structures are defined as depressions that form naturally through the erosion of basement rock and permit the temporary storage of rainwater – these are assumed to be present primarily within the Granite Landscape Division. Two main forms are recognised: pans are wide, shallow depressions, generally < 100 mm 94 deep; and gnammas, narrow cavities that can reach several metres in depth, especially if they intersect natural fractures (Campbell 1997). While other forms are recognised in the geological literature (e.g. Campbell 1997), these types are often defined based on their mode of formation, while the focus here is on physical characteristics that would affect the rate of water-loss. Rock structures (particularly gnammas) were often discussed in the scientific literature, but more often with reference to their overall capacity (e.g. Gunn 2006; Rossi 2010; Timms 2013; Webb and Rossi 2008) without regard for how they receive and lose water. In low rainfall zones such as the study area, capacity alone can be an inaccurate means of characterising rock structures since, like soils, their capacity may greatly exceed input. It is well-known that Aboriginal people used the water stored in rock structures (e.g. Bayly 1999, 2015; Roe 1836, 1852) – therefore, an attempt was made to quantify the volume of water captured by gnammas and pans under various input conditions, and to determine the longevity of the water supplies. However, claims that Aboriginal people intentionally enlarged depressions (e.g. Bayly 1999; Lullfitz et al. 2017) are not addressed, since these do not affect the rate of water capture or loss in the generalised structures considered here. Modelling the longevity of water in rock structures Rock structures are closed systems, as they have no interaction with soil or the groundwater table. Therefore, quantifying input and water-loss is simple using rainfall and evaporation data, but the latter cannot be used in its original form. While there is some inconsistency and inaccuracy associated with the Evaporation Pan instrument (Lowe et al. 2009), the main concern is how accurately the Pan Evaporation rate represents evaporation in natural bodies of water. Evaporation Pans are raised metal constructions, with sides that are exposed to direct sunlight – they cannot store heat in the way that natural bodies of water can (Dingman 1994; Tanny et al. 2008). While there are many complex methods of estimating evaporation relative to the Pan Evaporation rate (e.g. Deardorff 1977, 1978; McMahon et al. 2013), the most common is to formulate a Pan Coefficient, a proportion of the Pan Evaporation rate that can be applied to various water bodies. Coefficients are not necessarily consistent year-round, and a unique value may be required for each month or 95 season due to the thermal properties of larger bodies of water (Lowe et al. 2009). Lakes and reservoirs tend to have higher water temperatures than Evaporation Pans in winter, and lower in summer, due to this insulating effect (Winter 1981). Globally, there has been considerable research into developing coefficients for reservoirs and lakes (e.g. Tanny et al. 2008), to quantify water-loss in agricultural or urban settings. Australian research is comparatively limited, and an average pan coefficient of 0.79 is generally applied (Lowe et al. 2009). This coefficient cannot be applied to rock structures, but it forms a valuable baseline measure from which to develop coefficients for pans and gnammas. Domingues-Villar et al. (2008) found that shallow rock structures (pans) lost water at a rate similar to Pan Evaporation, but a precise rate was not specified. Evaporation pans have exposed sides, and metal transfers more heat to stored water than rock will – therefore, a coefficient of 0.9 was applied to rock pans herein, as water would be lost far more quickly than from reservoirs. Unlike reservoirs, gnammas store water further below the ground surface and have a small surface area, providing more protection from the wind and sun, two of the main drivers of evaporation (Shaw 2014:64–65). Hillel (2004:421) noted that areas sheltered from wind had evaporation rates around 20% lower than unsheltered areas. To reflect this, a coefficient of 0.4 was applied to gnammas. In reality, evaporation would likely vary with depth, with the coefficient being closer to the rock pan measure when water was stored close to the surface (i.e. when the gnamma was nearly full) but lower where water was stored at the base of a very deep gnamma. Nevertheless, a coefficient of 0.4 should represent most scenarios. Seasonal coefficient variation was not considered here, since small bodies of water will not store heat in the same way that larger ones do (Dingman 1994). Rock structure dimensions vary greatly (Timms 2013), but morphological generalisations were required in order to apply the methods used in this research. A maximum water-storage capacity of 100 mm was imposed on pans due to their shallow depth, but no limit was placed on gnamma depth. Structures were assumed to be straight-sided, so dimensions were identical at the top and bottom of the structure. A sub-category was created to represent structures receiving additional input: runoff gnammas and runoff pans. These subcategories were necessary since 96 not all structures will receive runoff. As with Granite soils (see Chapter 5.4.2), runoff structures were assumed to receive twice the daily rainfall value, while regular pans and gnammas received direct input only. Water-loss was assumed to occur only through evaporation – animal use was discounted as it cannot easily be quantified. Similarly, water-loss or conservation by human actions (drinking, using lids to reduce evaporation) was not considered, as the aim was to quantify the longevity of water under natural conditions; consumption and conservation are considered elsewhere (see Chapters 8–12). Using the coefficients derived above, in conjunction with the relevant rainfall and evaporation data, it was possible to determine the daily water-balance of each of the four rock structure types (gnammas, runoff gnammas, pans, runoff pans) across the six modelled years, using an Excel spreadsheet. Evaluating longevity was simple, but the use of a spreadsheet permitted more flexibility and allowed conditions to be altered to better replicate the desired scenarios. For example, the longevity of water was often affected by large falls that occurred outside winter. These falls could be voided to investigate the longevity of water accumulating via winter rains. Similarly, isolated falls of various sizes could be simulated when structures were dry, and the longevity of water evaluated. The results of these analyses are provided in Chapter 7. Limitations The main limitations with this method concern generalisations, and thus how representative the results may be. Rock structures are undoubtedly more common in the Granite Landscape Division, but they may occur elsewhere on exposures too small to be visible at the scale represented in Figure 5.2. Similarly, not all Granite areas will exhibit all types of rock structures – pans are ubiquitous, but gnammas are less common. These issues cannot be resolved at a Landscape Division level, but the presence/absence of different rock structures will be considered when evaluating the resource base of specific archaeological sites and localities. Similarly, the generalised dimensions and runoff characteristics of rock structures will not represent all scenarios. Some structures may receive runoff from a larger area or may be more protected from evaporation by vegetation or overhanging rock – these structures would then receive and lose water at different rates. Moreover, an outcrop will shed varying proportions of individual rainfall events, based on the size and 97 timing of the fall (Laing and Hauck 1997). Nevertheless, generalisations permit broad trends to be evaluated, and the baseline data can be amended to better represent the specific morphological and runoff characteristics of rock structures at particular sites (see Chapters 9–12). Finally, the coefficients used may introduce error, as with the crop coefficient used when quantifying soil water (see Chapter 5.4.2). While precise quantities and longevity estimates may therefore be flawed, the patterns should be representative of the relative longevity of water in various structures, under various rainfall conditions. 5.5 FORMULATING THE OCCUPATION MODEL An accurate occupation model cannot solely rely on the diversity of food resources available in each Landscape Division, but must also consider the nutritional content and procurement costs (time taken to collect and process particular resource using traditional technology and methods) associated with different resource types. These data are used to calculate return rates, generally expressed as the number of calories that can be obtained per handling hour (Table 5.5). By ranking foods by return rates, it is possible to determine which may exert more control over occupation patterns, since higher ranked items should be pursued first, and patches yielding them should be prioritised over lower ranked patches (Macarthur and Pianka 1966; see Chapter 2.2). For plant foods, the required data were fairly easily accessible. Published return rates were used where available, i.e. where the genus/species used in the calculation was representative of, or similar to, those occurring in the study area (Table 5.5). Otherwise, return rates were calculated manually, using published handling times for plant food categories in conjunction with representative nutritional data, either for a genus/species abundant in the study area or a published average value. Handling times were unavailable for several plant food categories, so values were inferred by determining whether one kilogram of a certain food would take longer to collect/process than another for which handling time is known. In reality, handling time and return rates vary according to the individual undertaking the task (Cane 1987; O'Connell and Hawkes 1981) and the species being collected, but this variation could not be incorporated into the model. 98 Table 5.5 Ranked plant food categories, based on return rates (cal./hr) determined by handling time (hr/kg) and nutritional value (cal./kg). Handling time and nutritional content are given only where return rates were calculated manually. Asterisks indicate inferred handling times. Moderate High Rank Resource category Handling Nutrition Return (hr/kg) (cal./kg) (cal./hr) Gum 0.25* 2370 9480 o Nutrition – Brand-Miller and Holt 1998 (average for gum). Lerps/ manna 0.5* 2870 5740 o Nutrition – Brand-Miller et al. 2014:234 (Psylla sp. scale lerps, manna). Storage organs 0.5 1290 2580 o Handling time – O'Connell et al. 1983 (average of rates provided). o Nutrition – Brand-Miller and Holt 1998 (based on averages for bulbs and tubers). Fruit 0.5 950 1900 o Handling time – O'Connell and Hawkes 1981; O'Connell et al. 1983. o Nutritional – Brand-Miller and Holt 1998 (average for fruit). 1462 o Bird et al. 2009 (Grevillea eriostachya). 1013 o Nutrition – Brand-Miller and Holt 1998 (based on averages for leaves and stalks/buds). 538 o O'Connell and Hawkes 1981 (Acacia spp.). Flowers Low Reference Leaves/ roots /stems Seeds 0.75* 760 Handling time is even more complex for animal foods due to varying methodologies – some researchers cite post-encounter returns, while others incorporate tracking time along with success rates, which vary seasonally (see Bird et al. 2009, 2012). Published return rates (using a consistent definition of handling time) were not available for all faunal categories, and could not be calculated manually; furthermore, prey size is not an accurate indicator of rank (Bird et al. 2009). Therefore, the distribution of animal types was incorporated into the model but in a less formal manner. Indeed, fixed resources (plants) are more likely to dictate occupation patterns than mobile prey (Bird 1985:131, 140). Based on these data, optimal foraging locations were identified in Landscape Divisions yielding the greatest diversity of higher ranked resources at a given time, but allowance was made for instances where diversity was not representative of abundance, for example, if a dominant genus or species provided a particular food product. While fauna could not be ranked, their distribution must be considered in order to accurately characterise the total resource base accessible at a particular time and place. Faunal distribution was often used to identify optimal patches where 99 plant foods could not, i.e. where two Landscape Divisions had similar suites of plant food but different faunal assemblages. Animal foods were assumed to be more influential in seasons when plant foods were less abundant and seasonal abundances (e.g. waterbirds arriving in winter) were assumed to exert a stronger control than fauna that were available year-round. Residential sites were presumed to occur within/near optimal foraging locations, in areas where potable water could be found. 5.6 CONCLUSION This chapter has demonstrated how the Landscape Divisions, based on vegetation and landform characteristics, can represent spatial variation at a manageable level. Plant and animal foods were identified within larger floral and faunal lists compiled for each Landscape Division, and the temporal availability of each food item quantified using published species- or genus-level data and a series of generalisations, where required. Freshwater drainage is lacking, and groundwater is saline, so water availability is strongly linked to rainfall, which varies considerably between years. Inter-annual rainfall variability was represented by selecting several model years, based on the quantity of winter rain and the distribution of other falls. These model years were used to evaluate the longevity of water in several different soil profiles – each representing a distinct Landscape Division – using the Soil Water Tool (DPIRD 2016a). These model years were also used in a specially devised method to evaluate longevity of water in rock structures. These methods provide a useful framework to quantify resource availability across all Landscape Divisions, under varying climatic conditions. The availability of low, moderate and high-ranked plant foods indicates optimal foraging locations that, when combined with the presence of potable water, can indicate which parts of the landscape were used at different times of the year. 100 CHAPTER 6 PLANT AND ANIMAL FOODS 6.1 INTRODUCTION This chapter focuses on quantifying the spatio-temporal availability of plant and animal resources that were identified following the methods outlined in Chapter 5. Plants and animals are considered separately, as different factors govern their distribution over space and time. The seasonal availability of each plant food category is considered, as well as the diversity of species in each Landscape Division, and how species composition affects the period over which each particular product was available. These data are then summarised to indicate when plant foods are most abundant, and where particular types are most commonly found. Animal foods are evaluated similarly, both the animals themselves and their eggs, where relevant, while also considering when certain species would be in peak condition and when others would be more easily encountered and captured. The temporal distribution of animal foods is then evaluated, before a final discussion outlining how faunal diversity varies across the study area. 6.2 PLANT FOODS A total of 467 plant species found in the study area furnished 674 different edible plant food items (Appendix B.1). Their frequency is shown in Table 6.1, along with a brief description of their recorded uses for each category. These data indicate that a diverse range of plant food products were present in the study area, a selection of which are pictured in Figure 6.1. They are unevenly distributed both seasonally and spatially, so the availability of each must be considered in the context of the relevant season and Landscape Division. 6.2.1 Gum A total of 101 species produce edible gum, comprising 69 Acacia, 30 Hakea, as well as Eucalyptus occidentalis and Pittosporum phillyreoides – the Acacias are some of the most prolific gum-producers world-wide (Cribb and Cribb 1975:184). Gum is only available during summer (Table 6.2), as it is produced when trees undergo stress, 101 normally due to heat and/or drought (Nussinovitch 2010:8). Gum is the highest ranked plant food due to its high caloric content combined with low handling time, as it can be exuded in great quantity and requires no processing before consumption (Figure 6.1b). The greatest diversity of gum-bearing species occurs in Mallee (n=60), followed by Heath (n=52), Woodland (n=38), Thicket (n=30) and Granite (n=27); only eight species are present in Saline areas. Acacia dominates in all Landscape Divisions, but Hakea is also a strong contributor in Mallee and Heath areas. Table 6.1 Total number of plant food products found in the study area. The number of plant species providing a particular food item is listed, so the overall total indicates the total number of plant food items in the study area rather than the number of food-bearing species. A species may yield more than one edible product. Plant uses are based on Bindon (1996), Low (1991), Meagher (1974) and Webb (2000); not all uses correspond to each food item and/or species. Food items are ranked from highest to lowest, as per Table 5.5. Plant food item No. species o Eaten in natural state; o Blended with other ingredients to make drinks; o Gum made into balls or cakes for future use. 101 Lerps/manna o Eaten in natural state; o Blended with other ingredients to make drinks. 46 Storage organs o Eaten raw or roasted; o Species with high water content eaten to quench thirst. 68 Fruit o Flesh eaten when fruits ripe; o Unripe fruits roasted and eaten; o Flesh eaten or stored once dried (intentionally or naturally); o Large seed or kernel inside fruit occasionally eaten, raw or roasted. 50 Flowers o Occasionally eaten whole or with other parts of plant; o Nectar sucked out of flowers; o Flowers soaked to make sweet drink. 252 Leaves/roots/stems o o o o Leaves/stems eaten raw, roasted or steamed; Tender leaf base sole edible part for some species; Roots cooked; Root bark pounded into flour and mixed with other ingredients. 38 Seeds o Ripe seeds eaten raw; o Ripe seeds ground into flour, mixed with water and formed into damper or cake that was then baked in ashes; o Certain unripe seeds roasted to render them edible. 119 High Gum Moderate Low Recorded uses TOTAL 674 102 Table 6.2 Spatio-temporal availability of all plant food categories within the study area, showing the number of species providing specific products during particular months. Foods are ordered by rank, from highest to lowest. HEATH GRANITE Summer Winter Spring Dec Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Gum 27 27 27 0 0 0 0 0 0 0 0 0 Lerps/manna 3 3 3 0 0 0 0 0 0 0 0 0 Storage organs 49 49 49 49 10 10 10 10 10 10 10 49 Fruit 14 14 9 8 5 6 9 10 12 12 12 14 Flowers 37 17 17 3 5 5 11 16 33 61 61 49 Leaves/roots/stems 14 14 14 15 15 15 15 15 15 15 15 15 Seeds 33 32 14 13 13 14 14 14 29 35 35 35 TOTAL 177 156 133 88 48 50 59 65 99 133 133 162 Gum 52 52 52 0 0 0 0 0 0 0 0 0 Lerps/manna 15 15 15 0 0 0 0 0 0 0 0 0 Storage organs 14 14 14 14 6 6 6 6 6 6 6 14 Fruit 22 27 16 15 8 8 16 16 17 17 17 18 Flowers 111 57 57 15 19 22 36 47 86 156 156 131 7 7 7 8 8 8 8 8 8 8 8 8 Leaves/roots/stems MALLEE Autumn Seeds 60 60 32 32 32 33 33 33 58 61 61 61 TOTAL 281 232 193 84 73 77 99 110 175 248 248 232 Gum 60 60 60 0 0 0 0 0 0 0 0 0 Lerps/manna 32 32 32 0 0 0 0 0 0 0 0 0 Storage organs 9 9 9 9 5 5 5 5 5 5 5 9 Fruit 17 20 11 11 7 7 15 15 16 16 16 17 Flowers 117 76 76 11 14 15 29 37 69 153 152 130 Leaves/roots/stems 9 9 9 10 10 10 10 10 10 10 10 10 Seeds 67 67 26 26 26 27 27 27 65 68 68 68 TOTAL 311 273 223 67 62 64 86 94 165 252 251 234 103 Table 6.2 Continued WOODLAND THICKET SALINE Summer Autumn Winter Spring Dec Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Gum 8 8 8 0 0 0 0 0 0 0 0 0 Lerps/manna 17 17 17 0 0 0 0 0 0 0 0 0 Storage organs 11 11 11 11 6 6 6 6 6 6 6 11 Fruit 11 10 7 3 1 3 4 5 7 8 9 10 Flowers 49 41 36 0 0 0 3 6 16 54 53 51 Leaves/roots/stems 29 29 29 29 29 29 29 29 29 29 29 29 Seeds 12 11 4 4 4 5 6 6 14 14 14 13 TOTAL 137 127 112 47 40 43 48 52 72 111 111 114 Gum 30 30 30 0 0 0 0 0 0 0 0 0 Lerps/manna 16 16 16 0 0 0 0 0 0 0 0 0 Storage organs 7 7 7 7 3 3 3 3 3 3 3 7 Fruit 13 16 12 11 4 4 8 9 9 9 9 10 Flowers 84 49 49 13 15 18 30 39 61 112 112 97 Leaves/roots/stems 6 6 6 7 7 7 7 7 7 7 7 7 Seeds 34 34 16 16 16 16 16 16 32 34 34 34 TOTAL 190 158 136 54 45 48 64 74 112 165 165 155 Gum 38 38 38 0 0 0 0 0 0 0 0 0 Lerps/manna 25 25 25 0 0 0 0 0 0 0 0 0 Storage organs 15 15 15 15 3 3 3 3 3 3 3 15 Fruit 15 15 11 10 6 7 9 10 13 14 14 16 Flowers 68 48 48 4 5 5 12 19 35 84 84 75 Leaves/roots/stems 11 11 11 12 12 12 12 12 12 12 12 12 Seeds 40 40 14 13 14 15 15 15 40 42 42 42 TOTAL 212 183 153 54 40 42 51 59 103 155 155 160 104 B D C A I Figure removed due to copyright restrictions G F E K M H J L N Figure 6.1 Examples of plant foods found in the study area – photographs and information sources as cited. A: Acacia acuminata, 3–5 m tall, bearing edible seeds and gum (Bindon 1996:251). B: Edible gum exuded from Acacia spp. (Low 1991:152). C: Typical Acacia seed pod (Sweedman and Merritt 2006:176). D: Woody fruits (2–4 cm long) containing edible seeds of Allocasuarina humilis (Erickson et al. 1979:21). E–F: Eucalyptus redunca and its flowers; tree to 6 m tall (Bindon 1996:131). G–H: Lerps on Eucalyptus spp. leaves (Low 1991:153). I: Chamelaucium megalopetalum, edible flowers 10–15 mm wide (Erickson et al. 1979:88). J: Enchylaena tomentosa fruits, approximately 5 mm wide (Low 1991:167). K: Thysanotus patersonii tubers, generally 10–30 mm long and 5 mm wide (Bindon 1996:251). L: Carpobrotus virescens, fruits and leaves (up to 65 mm long) are both eaten (Daw et al. 1997:4–5). M: Atriplex vescaria shrub (≤ 1 m high) which bears edible leaves (Bindon 1996:46). N: Edible fruits of Santalum acuminatum, tree grows to 5 m, fruit 20–40 mm wide (Daw et al. 1997:52– 53). 105 6.2.2 Lerps/Manna Lerp refers to tiny deposits on Eucalyptus leaves, comprising starchy and sugary excretions of sap-sucking psyllids (Figure 6.1g–h). Other insects cause trees to exude sugary liquids that harden into flakes, called manna (Low 1991:153). Both substances occur only on Eucalyptus and are produced as a result of insects feeding on sap, specifically phloem sap, which carries organic nutrients within a liquid substrate. Phloem sap is produced in leaves and shoots as a result of photosynthesis and transported to non-photosynthetic parts of the plant (Atkins and Ruan 2010–2016). Photosynthesis requires sunlight, and while it occurs year-round in evergreen trees, rates are significantly increased where temperatures are higher and there are more sunlight hours per day (Bell and Williams 1997); both of these measures peak in summer. Therefore, the availability of lerps/manna is restricted to the summer months (Table 6.2) following the methodology employed herein (see Chapter 5.3.3). According to Low (1991:153), lerps and manna were important foods to Aboriginal people, and could be gathered in considerable quantity. These foods are ranked highly due to their high nutritional value and low handling time (Table 5.5). A total of 46 Eucalyptus species in the study area produce lerps/manna – the remaining three species may also have provided these foods, but no data were available. The greatest diversity of lerps/manna-bearing species is found in Mallee (n=32) and Woodland (n=25) areas. Heath, Thicket and Saline areas all preserve 15–17 species, while only three are present in Granite. 6.2.3 Storage Organs A total of 68 plants supplied edible storage organs, dominated by Caladenia (n=15), Drosera (n=9), Platysace (n=7) and Thelymitra (n=6). Storage organs provide the best returns of all moderately ranked plant foods, as their good nutritional content is combined with short handling time (Table 5.5). A two-phase availability period is evident as some species have perennial storage organs, while others exhaust their organ to reshoot following a period of summer dormancy (Figure 6.2). Their organs are then replenished over the subsequent growing season, but this replacement effectively renders them unavailable for that period. 106 60 Storage organs Granite Mallee Thicket Heath Saline Woodland 50 no. species 40 30 20 10 0 Dec Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Jun Jul Aug Sep Oct Nov Jun Jul Aug Sep Oct Nov Jun Jul Aug Sep Oct Nov 30 Fruit 25 no. species 20 15 10 5 0 Dec Jan Feb Mar Apr May 180 Flowers 150 no. species 120 90 60 30 0 Dec Jan Feb Mar Apr May 90 Seeds 75 no. species 60 45 30 15 0 Dec Jan Feb Mar Apr May Figure 6.2 Plant food categories demonstrating strong temporal variation in availability, showing monthly species diversity within each Landscape Division. The same colour key applies to each chart. 107 During the first phase, April–October, only the 18 species with perennial storage organs are available. Diversity is greatest in Granite areas (n=10), followed by Saline and Heath (n=6 each), Mallee (n=5), Woodland and Thicket (n=3 each). From November–March, an additional 50 species are available – those that spent the previous months refilling their storage organs in preparation for the impending dormant period. In this second period, Granite areas still yield the most diverse assemblage (n=49), and no other Landscape Division preserves < 25% of the species found in the study area. However, plants with perennial storage organs now form but a small proportion of the overall assemblage so, elsewhere, the spatial pattern of diversity shifts. Woodland areas now preserve the second most diverse assemblage (n=15), closely followed by Heath (n=14), while Saline, Mallee and Thicket all yield less diverse assemblages (n=7–11). Five species have high watercontent during summer; 3–4 of these can be found in most Landscape Divisions, but only two are present in Woodland and Saline areas (Appendix B). 6.2.4 Fruit Edible fruits are provided by 50 species representing 18 genera, the most dominant being Leucopogon (n=11), Persoonia (n=8) and Cassytha (n=6). Most fruits are small, but the quandong (Santalum acuminatum) can measure 20–40 mm in diameter (Figure 6.1n). A total of 18 species, including Leucopogon spp. belong to the Eriaceae family that produce very small fruits less < 10 mm in diameter, with hard stones surrounded by sweet pulp (Low 1991:128). Fruit rank moderately among plant foods, since their small size increases handling time. Overall diversity is greatest in Heath areas (n=32), followed by Mallee (n=26), Woodland (n=21), Granite and Thicket (n=19 each). Saline areas supply the fewest species, just 14. This pattern holds true for most of the year, but some seasonal variation is evident. Fruits diversity peaks during December–January, when 44 of the 50 fruit species are available, 10–27 of which are found in each Landscape Division (Figure 6.2). The assemblage is most diverse in Heath areas, unsurprising since Eriaceae, 16 of which produce fruit during summer, are common in heath formations. Diversity decreases in February and March, as Leucopogon no longer produces fruit – at this time, Thicket areas yield the second most diverse assemblage after Heath. Fruit availability reaches its minimum in April, when only 12 species are producing fruit, 108 and no more than eight can be found in a single Landscape Division. The most diverse assemblages are present in Heath, followed by Mallee, Woodland, Granite, Thicket and Saline areas. This pattern persists thereafter, but fruits become more abundant from June, when Leucopogon are again available. At this time, diversity values are more clustered – Heath and Mallee yield similarly-sized assemblages (15–16 species), as do Woodland, Granite and Thicket (8–9 species). This clustering becomes less pronounced by late winter and values are more evenly spread throughout spring, as the availability of fruit increases. The increase is particularly pronounced in Granite and Woodland areas, whose late spring diversity values matched or exceeded those of summer, when diversity peaked in all other Landscape Divisions. 6.2.5 Flowers Flowers comprise the most abundant form of plant food, with 252 species representing 31 genera. The assemblage is heavily dominated by Eucalyptus (n=47), Melaleuca (n=47), Hakea (n=30), Grevillea (n=25) and Verticordia (n=22); Grevillea, Melaleuca and Verticordia yield no edible products aside from their blooms. Flowers were occasionally eaten whole, or with other parts of the plant, but most were exploited purely for their sweet nectar. Flowers are a moderately ranked resource because most species only produce a small amount of nectar in each flower, increasing handling time and limiting returns (Table 5.5). The most diverse assemblages are found in Heath (n=158) and Mallee (n=155) areas, followed by Thicket (n=114), Woodland (n=86), Granite (n=62) and Saline (n=54). Temporal variation is pronounced (Figure 6.2) – many species flower in spring and summer, but several produce flowers year-round, or at other times of the year. The distinct flowering window exhibited by each genus or species, combined with the varying representation of genera create a unique pattern of seasonal flower availability within each Landscape Division. A total of 189 plants flowered at some point during summer, including every species from four of the dominant genera; only Hakea were not flowering. Unsurprisingly, then, summer diversity was greatest in those areas with the highest number of total species, namely Mallee and Heath, followed by Thicket, Woodland, Saline and Granite. In all cases, diversity was greatest in December, when 37–117 species 109 flowered in each Landscape Division. Diversity then decreases in January, when Verticordia and most Grevillea are no longer flowering, and reaches its minimum in March, at which point Eucalyptus and Melaleuca are no longer in bloom. Just 26 species produced flowers at some point during autumn, including ten Banksia that flower throughout the year. There is a slight increase in diversity as autumn progresses, but diversity is greatest where the Banksia genus is well-represented, specifically Heath (n=15–22), Thicket (n=13–18) and Mallee (n=11–15); Woodland and Granite provided no more than five species per month, and no plants in Saline areas flowered in autumn. From June–August diversity increased, as Grevillea and Hakea begin to flower at some point during winter. This increase is especially pronounced in Heath, Mallee and Thicket areas (61–86 species), where those genera are well-represented. In contrast, other Landscape Divisions yielded more reduced assemblages, between 16 and 35 species. By September, overall diversity reaches its peak, as several dominant genera begin to flower, including Eucalyptus (n=46) and all Melaleuca (n=47) and Verticordia (n=22). Diversity is greatest in Heath and Mallee areas, where over 150 species are flowering, and particularly limited in Granite and Saline, where fewer than 65 are. Species diversity declines slightly throughout spring, as Hakea (n=30) do not flower after October; the decline is most pronounced in areas where that genus is well-represented, but the inter-Division diversity rankings remain largely unchanged. 6.2.6 Leaves/Roots/Stems Leaves/roots/stems is the least diverse plant food category in the study area – just 38 species are present, including six Atriplex and five Frankenia. Despite the limited diversity, leaves/roots/stems are available throughout the year (Table 6.2). Only one species – Thysanotus patersonii – exhibits seasonal variation due to summer dormancy, at which point the above-ground component of the plant dies off (Daw et al. 1997). Leaves/roots/stems are a low-ranked resource as their low nutritional content means that large quantities must be collected to provide sufficient calories, increasing handling time (Table 5.5). This resource is most diverse in Saline areas, which provided a total of 29 species, 16 of which were not present in any other Landscape Division. Diversity and exclusivity are pronounced here, because many 110 plants within the leaves/roots/stems category are adapted to survive in salt-affected areas. Less than half the study area assemblage was present elsewhere – just 15 species were found in Granite areas, 12 in Woodland, 10 in Mallee, eight in Heath and seven in Thicket areas. 6.2.7 Seeds After flowers, seeds are the most diverse plant food category found in the study area. A total of 119 species from 11 genera produce edible seeds. The assemblage is heavily dominated by Acacia (n=70; Figure 6.1c) and Hakea (n=30). While seeds have good caloric value, particularly Acacia, the nutritional benefits are offset by the increased handling time required for collection and processing (Table 5.5). The seed-production periods of the dominant genera – one of which lasts for six months while the other occurs year-round – dictate the temporal variation in this food group and creates two distinct six-month availability periods (Figure 6.2). During the first period, February–July, seeds are comparatively rare. The bulk of the assemblage comprises Hakea (n=30) and Allocasuarina (n=9); their seeds are available year-round as they are encased in woody fruits that persist on the plant for a year or more (Wildlife Society of Western Australia 2013). Just three species have shorter seed-bearing windows centred around autumn and/or winter. Overall diversity was therefore skewed towards areas where the Hakea and Allocasuarina were well-represented, primarily Heath (n=33), followed by Mallee (n=27), Thicket (n=16), Woodland (n=15), Granite (n=15) and Saline (n=6). Seeds are abundant from August–January, when up to 116 different species produce seeds, including most of the Acacia genus. The peak occurs from September to November, where each Landscape Division yields 14–68 different types of seeds. The more heterogenous assemblage alters the spatial pattern of diversity, with the greatest assemblage found in Mallee (n=68) and Heath (n=61) areas, followed by Woodland (n=42), Granite (n=35), Thicket (n=34) and Saline (n=14) areas. 6.2.8 Summary Overall, plant foods are most diverse in Mallee and Heath areas (360 and 341 items, respectively) and scarcest in the Saline Landscape Division, which yields < 150 plant foods (Figure 6.3). However, there is strong temporal variation in plant food 111 availability (Figure 6.2), so rather than focussing on overall diversity it is important to consider what items are available in different seasons, how these items are ranked, and where the greatest diversity and/or quantity can be found. Plant foods are most abundant in summer, peaking at 585 different items in December. More importantly, all seven plant food categories are available, including the highly ranked gum and lerps/manna that are unavailable for the rest of the year (Figure 6.4). Lerps/manna is derived from Eucalyptus, so would be most abundant in Woodland and Mallee areas where that genus dominates. Both of these Divisions also preserve a diverse range of gum-bearing plants, but these are restricted to the understorey, so would be smaller and less common. Despite the limited species diversity (n=30), Thicket areas would yield the greatest quantity of gum, since Acacia is one of the three dominant genera and a prolific gum-producer (Table 3.2; Cribb and Cribb 1975:184). A more varied assemblage is present in Heath areas, but the quantity of gum would be reduced due to a more diverse floral assemblage combined with smaller plant sizes. Moderate-ranked summer foods comprise storage organs, fruits and flowers. The latter two categories are abundant in Mallee and Heath, while Granite provides three times as many plants with storage organs – which have the highest returns of the moderately ranked plant foods (Table 5.5) – than all other Landscape Divisions. Five species also yield high water content in summer, and Mallee, Heath and Granite areas each yield three or four of these (Appendix B). 400 350 Flowers Fruit Gum Leaves/roots/stems Lerp/manna Seeds Storage organs 300 no. items 250 200 150 100 50 0 Granite Heath Mallee Saline Thicket Woodland Figure 6.3 Total number of different plant foods within each Landscape Division. 112 All plant foods 700 600 Flowers Fruit Gum Leaves/roots/stems Lerp/manna Seeds Storage organs 500 no. items 400 300 200 100 0 Dec Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Plant foods by rank 700 Low Moderate 600 High no. items 500 400 300 200 100 0 Dec Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Figure 6.4 Total number of plant food items (top) and rank (bottom) available in each month across the entire study area. By autumn, plant foods are much scarcer – highly ranked items are no longer available and most other categories exhibit a marked decline in species diversity. From April, moderately ranked foods are restricted to plants that retain their subsurface storage organs year-round (n=18) as well as 23 species with edible flowers, and 13 with fruit. As in summer, these foods are most diverse in Heath, Mallee and Granite areas, though the latter mostly yields storage organs, which are also fairly diverse in Saline areas. At this time, both low-ranked items (leaves/roots/stems and seeds) comprise one-third of the available plant foods; most are available year-round, rather than coming into season in autumn. Leaves/roots/stems are most abundant and diverse in Saline areas, while a range of seed-producers can be found in Heath and Mallee. Thicket would provide a more homogenous assemblage, but seeds could be collected in great quantity as one of the three dominant genera (Allocasuarina) retains seeds year-round. As winter approaches plant foods are still comparatively scarce, but as the season progresses more fruits and flowers become available. The latter are abundant in Thicket areas 113 and the former in Heath and Mallee, but Woodland and Granite also provide a good range of fruits. By spring, moderately ranked foods are present in good supply, but the increase is primarily due to the flower category, which is well-represented in Heath, Mallee and Thicket areas. Lower-ranked seeds are also increase in abundance in those same Landscape Divisions. By November, storage organs are again widespread and diverse as many species have replenished their subsurface energy stores over the growing season, in preparation for summer dormancy. Storage organs would once again be an important food source, concentrated in Granite areas. 6.3 ANIMAL FOODS A total of 142 vertebrate fauna species are present in the study area, comprising 89 birds, 28 reptiles, 20 mammals and five amphibians (Appendix C). As well as the meat from these animals, eggs were available from 85 bird and 23 reptile species. Each of these animal types, and the resultant food products, are discussed separately below, in order of abundance. 6.3.1 Birds and their Eggs A total of 89 bird species are present, representing 15 orders (Appendix C); 85 of these birds breed in the study area. Passeriformes (perching birds) dominate the assemblage, but waterbirds (n=9) and cockatoos/parrots (n=8) are fairly common; emu and malleefowl are also present (Figure 6.5). Most species are non-migratory (n=68), and inhabit the study area for the entire year, while the remainder (n=21) are rarely present for more than six months of the year. It is these migratory patterns that create the seasonal variation in the avian assemblage, but temporal variation in each Landscape Division is more subdued (Table 6.3; Figure 6.6). Woodland and Mallee always yield the most diverse assemblages, while Heath, Granite and Thicket are closely grouped, yielding around 15–20 fewer species than Woodland. The least diverse assemblages are generally found in Saline areas, but waterbirds (n=9) are restricted to this Division – the three sedentary species are likely confined to the more permanent saline waters in the eastern portion of the study area. Emu (Dromaius novaehollandiae) are ubiquitous, while malleefowl (Leipoa ocellata) are restricted to Mallee, Heath and Thicket areas. 114 A B C F G Figure removed due to copyright restrictions D H E I J Figure 6.5 Examples of birds found in the study area. Measurements refer to the body length (from crown to tip of the tail) unless noted otherwise. Common names follow species name. Photographs and measurements from Nevill (2014:52–55, 78–81, 90–91, 106–111). A–C: Passeriformes (perching birds): Phylidonyris niger, white-cheeked honeyeater, 18 cm (A); Pomatostomus superciliosus, whitebrowed babbler, 21 cm (B); Petroica goodenovii, red-capped robin, 12 cm (C). D–G: Waterbirds: Himantopus himantopus, black-winged stilt, 37 cm (D); Tadorna tadornoides, Australian shelduck, 56–72 cm (E); Charadrius ruficapillus, red-capped plover, 15 cm (F); Vanellus tricolor, banded lapwing, 27 cm (G). H: Dromaius novaehollandiae, emu, 160–190 cm tall. I: Leipoa ocellata, malleefowl, 55–60 cm. J: Barnardius zonarius, Australian ringneck, 36 cm. 115 Table 6.3 Spatio-temporal availability of all animal food categories within the study area, showing the number of species providing a certain product during each month, across all Landscape Divisions. MALLEE HEATH GRANITE Summer Autumn Winter Spring Dec Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Amphibian 4 4 4 4 4 4 4 4 4 4 4 4 Bird 48 47 46 46 47 47 48 49 50 50 49 49 Bird eggs 16 4 0 2 2 1 4 15 39 46 43 28 Mammal 9 9 9 9 9 9 9 9 9 9 9 9 Reptile 15 15 15 15 15 15 15 15 15 15 15 15 Reptile eggs 11 0 0 0 0 0 0 0 0 11 11 11 TOTAL 103 79 74 76 77 76 80 92 117 135 131 116 Amphibian 1 1 1 1 1 1 1 1 1 1 1 1 Bird 50 49 48 47 49 49 50 52 53 52 52 51 Bird eggs 18 5 0 2 2 1 4 15 40 48 46 29 Mammal 11 11 11 11 11 11 11 11 11 11 11 11 Reptile 21 21 21 21 21 21 21 21 21 21 21 21 Reptile eggs 16 0 0 0 0 0 0 0 0 16 16 16 TOTAL 117 87 81 82 84 83 87 100 126 149 147 129 Amphibian 0 0 0 0 0 0 0 0 0 0 0 0 Bird 61 60 59 58 60 62 63 65 67 65 65 62 Bird eggs 19 6 0 2 1 1 5 20 53 62 57 36 Mammal 14 14 14 14 14 14 14 14 14 14 14 14 Reptile 23 23 23 23 23 23 23 23 23 23 23 23 Reptile eggs 19 0 0 0 0 0 0 0 0 19 19 19 TOTAL 136 103 96 97 98 100 105 122 157 183 178 154 116 Table 6.3 Continued Summer WOODLAND THICKET SALINE Dec Jan Autumn Feb Mar Apr Winter May Jun Jul Spring Aug Sep Oct Nov Amphibian 0 0 0 0 0 0 0 0 0 0 0 0 Bird 40 39 39 38 39 40 45 46 50 49 47 44 Bird eggs 12 3 0 1 2 2 6 18 44 43 39 26 Mammal 2 2 2 2 2 2 2 2 2 2 2 2 Reptile 4 4 4 4 4 4 4 4 4 4 4 4 Reptile eggs 3 0 0 0 0 0 0 0 0 3 3 3 TOTAL 61 48 45 45 47 48 57 70 100 101 95 79 Amphibian 1 1 1 1 1 1 1 1 1 1 1 1 Bird 45 44 44 44 46 47 48 49 49 49 48 46 Bird eggs 15 4 0 2 1 1 3 15 39 47 43 23 Mammal 7 7 7 7 7 7 7 7 7 7 7 7 Reptile 8 8 8 8 8 8 8 8 8 8 8 8 Reptile eggs 6 0 0 0 0 0 0 0 0 6 6 6 TOTAL 82 64 60 62 63 64 67 80 104 118 113 91 Amphibian 0 0 0 0 0 0 0 0 0 0 0 0 Bird 66 65 63 61 62 64 65 67 69 68 69 67 Bird eggs 21 5 0 2 2 2 5 18 54 65 60 40 Mammal 17 17 17 17 17 17 17 17 17 17 17 17 Reptile 23 23 23 23 23 23 23 23 23 23 23 23 Reptile eggs 19 0 0 0 0 0 0 0 0 19 19 19 TOTAL 146 110 103 103 104 106 110 125 163 192 188 166 117 Overall, seasonal change in avian diversity is limited, with minimum and maximum values for a single Landscape Division varying by 4–12 species. From December– March all assemblages decline, as eight migratory species leave the area. The decrease is most pronounced in Woodland areas, where seven of these species are found. Diversity then slowly increases throughout autumn and winter, as migratory birds arrive, and peaks in August, when 49–69 species are present in each Landscape Division. The increase is particularly evident in Saline areas, due to the arrival of migratory waterbirds that arrive when ephemeral salt lakes fill with winter rain. As a result, Saline, Granite and Thicket areas now yield similarly diverse assemblages (49–50 species). Diversity decreases through spring, as nine migratory birds leave and only three arrive. This decline is most pronounced in Saline areas, due to the departure of migratory waterbirds, so the gap in diversity between this and other Landscape Divisions again begins to widen. 90 Birds 80 Granite Heath Mallee Saline Thicket Woodland 70 60 no. species 50 40 30 20 10 0 Dec Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Jun Jul Aug Sep Oct Nov 80 Bird eggs 70 60 no. species 50 40 30 20 10 0 Dec Jan Feb Mar Apr May Figure 6.6 Monthly availability of birds (top) and birds’ eggs (bottom) in each Landscape Division. The same colour key applies to each chart. 118 Avian egg-laying is seasonal (Figure 6.6) – of the 89 species of birds that visit or inhabit the study area, < 30% produce eggs during summer (n=22). Most of these species lay in December and, unsurprisingly, the pattern of diversity parallels that of birds themselves, and is highest in Woodland and lowest in Saline areas. Diversity declines through January when only six species, including the malleefowl, are laying. By February, no eggs are available, and numbers remain low through autumn, with no more than two species laying eggs in any Landscape Division at one time. Significantly, emu produce their extremely large eggs in May, and could be found throughout the study area. By June, eggs are still scarce, with the most diverse assemblage found in Saline areas, where migratory waterbirds breed. Diversity increases through winter, as nonmigratory birds also begin laying, and a separation starts to form between Woodland and Mallee areas, where diversity increases at a greater rate, and the remaining Landscape Divisions where fewer bird species are found. In Saline areas, egg diversity peaks in August (n=44) and declines thereafter. Elsewhere, the maximum occurs in September when Woodland and Mallee yield 62–65 species, while Heath, Granite and Thicket preserve 46–48. Diversity then decreases sharply through October and November, and the gaps between Landscape Divisions begin to narrow once again. 6.3.2 Reptiles and their Eggs The reptile assemblage comprises 25 lizards and three snakes. Most of these produce eggs except four lizards (all skinks) and one snake that give birth to live young. The lizard group includes 10 skinks (family Scincidae), nine geckos (suborder Gekkota) and five dragons (Figure 6.7). Reptiles are most diverse in Woodland and Mallee areas, which both harbour 23 species, including one – a species of gecko – not found elsewhere; these Landscape Divisions provide all snake and most dragon and gecko species. Slightly fewer reptile species are present in Heath areas (n=21), and the dragon and gecko assemblages are less diverse than in Woodland and Mallee areas. Just 15 reptiles are present in Granite areas, including two dragons, and a sand monitor. Reptiles were comparatively rare in Thicket (n=8) and Saline areas (n=4). 119 B D A E C Figure removed due to copyright restrictions H F G K I J L Figure 6.7 Example of native non-avian fauna present in the study area. Common names are provided after species name, measurements cited are maximum total length (TL – including tail) or body length (BL – excluding tail) Photographs and measurements from Nevill (2014:124–129, 136– 137, 142–143, 151, 156–157, 162–163, 184–185). A–D: Reptiles: Ctenophorus cristatus, bicycle dragon, TL 370 mm = (A); Liopholis multiscutata, bull skink, TL 250 mm (B); Underwoodisaurus milii, barking gecko, TL 165 mm (C); Varanus gouldii, sand monitor, TL 1.6 m (D). E–K: Mammals: Sminthopsis crassicaudata (family Dasyuridae), fat-tailed dunnart, BL 75 mm (E); Dasyurus geoffroii (family Dasyuridae), chuditch, BL 310–360 mm (F); Macropus fuliginosus, western grey kangaroo, BL 0.95–2.2 m (G); Chalinolobus gouldii, Gould’s wattled bat, BL 65–75 mm (H); Isoodon obesulus, southern brown bandicoot, BL 340 mm (I); Trichosurus vulpecula, common brushtail possum, BL 380 mm (J); Tachyglossus aculeatus, short-beaked echidna, BL 400 mm (K). L: Myobatrachus gouldii, turtle frog, BL 50 mm. 120 Reptile species are present year-round (Table 6.3), but would be more accessible in spring and summer as they are considerably more active, especially when seeking mates. In winter, they are far less active, so would be especially difficult to catch unless their burrows could be located; as a result, opportunities to hunt them would be limited. Eggs are only available for four months of the year, September– December. The greatest variety occur in Woodland and Mallee areas, where 19 different types of egg were available, followed by Heath (n=16) and Granite (n=11); reptile eggs are rare in Thicket (n=6) and Saline (n=3) areas, due to their limited reptile assemblages. 6.3.3 Mammals Twenty mammal species were found in the study area, comprising six Dasyurids (belonging to the family Dasyuridae – carnivorous marsupials, all small and mouselike in the study area except for a chuditch), three bats, three macropods, three possums, two rodents, as well as a single echidna, numbat and bandicoot species (Figure 6.7). A total of 17 species inhabit Woodland areas, including four that do not occur elsewhere (a numbat, possum, bat and Dasyurid). Mallee areas housed 14 mammal species, while Heath had only 11, lacking all bats and many of the Dasyurids found in Woodland and Mallee areas. Mammals were even rarer in the remainder of the study area – just nine species were recorded at Granite areas and seven at Thicket. Only two mammal species – one macropod and the echidna – were present in Saline areas. Temporal variation is very subdued, as all mammal species are present year-round (Table 6.3). However, two macropod species (Macropus fuliginosus, Osphranter robustus) are in peak condition from June to September due to abundant winter pasture; their condition deteriorates from October as water became scarcer. It is possible that other herbivores would exhibit similar seasonal variation in bodily condition. 6.3.4 Amphibians Just five species of amphibian are found in the study area: two frogs, two toadlets and one froglet (Figure 6.7). Amphibians are only found in three Landscape Divisions – Granite, Heath and Thicket. They are most diverse in Granite areas, which yield two toadlets, a frog and a froglet. The turtle frog (Myobatrachus gouldii) is confined to Heath and Thicket. This frog can live in areas without standing water since eggs 121 are laid in the soil and the young frogs develop therein (Cogger 2014:172; Cronin 2014:38). All species are present year-round (Table 6.3), but Meagher (1974) noted that people often preferred to eat females as they may be holding eggs at various times of the year. The breeding period, when eggs may be present, generally lasts for 3–4 months and is generally concentrated in autumn or winter. In Granite areas, eggs may be found in three different species from March–June; just one species breeds until September. The turtle frog has a different breeding cycle – it calls for mates around July, and breeds several months later, so the young frogs are ready to emerge with the first autumn rains (Cronin 2014:38; Western Australian Museum n.d.). Hence, the females may contain eggs from October to December. 6.3.5 Summary While the suite of plant foods varies seasonally, there is little temporal variation in the faunal assemblage (Table 6.3, Figure 6.8). Seasonal changes relate primarily to bodily condition, ease of capture or egg production rather than the presence/absence of particular species. In spring and summer, reptiles are most active, most easily caught, and are also producing eggs. In winter, macropods are in peak condition after feeding on abundant winter pasture. Migratory waterbirds also arrive in the area and many avian species are laying eggs. Reptiles would be difficult to catch, however, as they typically retreat to nests or burrows to await warmer weather. Female amphibians may contain eggs anytime from autumn to spring, depending on the species. Variation is more pronounced at the spatial level, reflecting the habitat preferences of different animal species, so it is necessary to consider the overall faunal diversity represented at each Landscape Division. Woodland areas exhibit the most diverse faunal assemblage, providing the largest number of animal foods (n=200) and the maximum diversity in five of the six food categories (Figure 6.9). Nearly all reptile and mammal species recorded in the study area can be found in Woodland locations, although amphibians are notably absent. Mallee areas supply slightly fewer animal food products overall, but reptiles (and their eggs) are as diverse as in Woodland areas, and a good range of birds and mammals are also present. Heath areas yield 155 animal foods, and all six categories are present. Bird and bird egg diversity is lower than in Woodland and Mallee, and just over half of the mammals present in the study area can be found in 122 Heath (n=11) – larger mammals would be less abundant due to the dense vegetation (Macintyre et al. 1992). Slightly fewer reptiles and their eggs are present but, unlike Woodland and Mallee areas, Heath harbours a single amphibian, the turtle frog, which may hold eggs in winter (Appendix C). 250 200 Amphibian Bird Bird eggs Mammal Reptile Reptile eggs no. items 150 100 50 0 Dec Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Figure 6.8 Total number of animal food items available in each month across the entire study area. 250 200 Amphibian Bird Bird eggs Mammal Reptile Reptile eggs Granite Heath no. items 150 100 50 0 Mallee Saline Thicket Woodland Figure 6.9 Number of different animal food items within each Landscape Division. Note that Heath and Thicket both preserve a single amphibian species, not visible at the scale of this figure. The faunal assemblage is less diverse in Granite areas, but this Landscape Division exhibits the greatest range of amphibians, of which four species are present. Females of three species may hold eggs in autumn, and another species in winter. Of the reptiles, birds and mammals found within the study area, around half are represented in the Granite Landscape Division, but both species of macropod are 123 present. Thicket areas have an even lower faunal diversity but, as in Granite and Heath areas, all six animal food categories are represented. Bird and bird egg diversity is similar to Granite, but mammals are slightly scarcer and larger ones would be rare, as in Heath, due to closely-spaced vegetation. The reptile assemblage is heavily reduced, with less than 30% of study area reptiles found in Thicket areas. Saline areas consistently exhibit the least diverse fauna. While 108 animal food items are present, 99 of these are birds or bird eggs – only four reptile and two mammal species are present. The avian assemblage is almost as diverse as Heath, Granite and Thicket areas, although the Saline Landscape Division represents the only area where waterbirds may flock in winter, when salt lakes are holding water. 6.4 CONCLUSION Plant foods are clearly abundant throughout the study area, but availability varies over space and time. They are abundant in spring and summer, when all or most categories are well represented, while autumn and winter are characterised by a reduced suite of plant foods. Heath and Mallee areas generally supply the greatest diversity of plant foods across all ranks, while the fewest species are associated with Saline environments, although the relative diversity evident in each Landscape Division can vary seasonally depending on the species composition – specifically the proportion of plants providing their products year-round compared to those with more restricted periods of availability. Temporal variation is less pronounced among the fauna, as most animals are available year-round, but some migratory birds are present for shorter periods. Bodily condition and ease of capture varies seasonally for terrestrial animals, as does the availability of eggs, concentrated in spring and early summer for reptiles, and August–October for birds. Woodland and Mallee areas preserve the most diverse fauna, while Saline areas yield few non-avian species. 124 CHAPTER 7 WATER AVAILABILITY 7.1 INTRODUCTION This chapter seeks to quantify the longevity of water sources available to human occupants of the study area. Due to its unique hydrological conditions, the only form of potable water is rainwater, preserved in either the soil profile or rock structures. Soil water availability is evaluated using the Soil Water Tool (DPIRD 2016a), and rock structures using methods devised specifically for this research, as outlined in Chapter 5. The availability of water within soil profiles and rock structures is quantified throughout six modelled years that represent periods of low, average and high winter rainfall, and also incorporate variously sized isolated falls (occurring outside winter); the spatio-temporal variation in each of these resources is then characterised. Finally, the soil water and rock structure data are synthesised to define the availability of potable water within the study area and its Landscape Divisions under different rainfall conditions, while considering the frequency at which various conditions occur. 7.2 SOIL WATER To quantify the availability of soil water in the study area, it is first necessary to analyse longevity within each Landscape Division – as represented by a unique idealised soil profile – under low (< 97 mm), average (97–177 mm) and high (> 177 mm) winter rainfall conditions. Each rainfall scenario is discussed separately, below. 7.2.1 Low Winter Rainfall – 2010, 2012 The small amount of rain received in winter 2010 (69.5 mm) meant that soil water was a short-lived resource and only available in limited quantities (Figure 7.1). Five of the six Landscape Divisions never had more than 11 mm available at any one time. Even Granite areas, despite receiving additional input by runoff, exceeded 40 mm of Plant Available Water (PAW) on just one occasion. Following germination (the point at which modelled crops begin growing and drawing on soil water – see Table 5.4) on 15 June 2010, Mallee soils held water for just a single day, while Saline and 125 Thicket areas preserved PAW for four or five days, respectively, and Heath and Woodland for ten days; Granite soils dried after 19 days. This relatively short waterholding period resulted from the rainfall distribution – while 14 mm fell on 15–16 June, just 2.7 mm was received over the remainder of the month. Most soils re-wetted as a result of 27 mm of rainfall between 9 and 12 July, but Mallee soils remained dry. Saline areas held water for three days, while Thicket, Heath and Woodland areas retained water for 6–10 days, before all dried permanently in mid-July. Subsequent falls were insufficient for any of these soils to re-wet, so none held water for more than three weeks over the entire year. Granite soils dried on 4 August, after holding water for 26 consecutive days, and 45 days in total. Granite soils may have responded to 16.2 mm that fell over five days in mid-August, but this cannot be evaluated with the current methodology. If PAW did accumulate, it would have been fairly short-lived, since just 0.8 mm of rain fell over the next two weeks. In 2012, most profiles held water more consistently than they had in 2010. Nevertheless, in five of the six Landscape Divisions water was present only for a few weeks in June. This is hardly surprising, considering that half of the winter rainfall total fell during 7–30 June. As in 2010, Mallee soils only held water for one day, as the preceding dry period meant that much larger falls would be required to recharge these profiles. Saline and Thicket soils held water intermittently until 24 and 25 June, respectively – Saline areas temporarily dried twice for a total of six days, while Thicket soils dried for just two days. Heath and Woodland areas held PAW consistently for three weeks, before drying on 28 June. Aside from Granite, only Heath and Woodland soils held more than 10 mm of PAW at any one time. None of the dry profiles responded to two 11 mm falls in August, and a number of larger falls later in the year (26.6 mm on 1–2 November 2012, 60.6 mm on 24–29 November 2012) were preceded by too lengthy a dry period for these soils to again accumulate PAW. Granite soils held water consistently for 75 days, from 7 June to 20 August 2012, nearly two months after Heath and Woodland areas dried. The early November falls may have elicited a weak response, but PAW would have been very short-lived. However, those in late November (totalling 60.6 mm of rain over five days) would have resulted in PAW being present for around two and a half weeks, when the recharge amount and average daily loss are considered (see also Chapter 7.2.2 and 7.2.4). 126 45 40 1… 1… 22-Jun-10 2… 2… 2… 29-Jun-10 3… 3… 6-Jul-10 6… 9… 13-Jul-10 1… 1… 1… 20-Jul-10 2… 2… 27-Jul-10 2… 3… 3-Aug-10 2… 5… 8… 10-Aug-10 1… 1… 17-Aug-10 1… 2… 24-Aug-10 2… 2… 2… 31-Aug-10 1… 4… 7-Sep-10 7… 1… 2010 14-Sep-10 1… 1… 1… 21-Sep-10 2… 2… 28-Sep-10 2… 1… 5-Oct-10 4… 7… 1… 12-Oct-10 1… 1… 19-Oct-10 1… 2… 26-Oct-10 2… 2… 3… 2-Nov-10 3… 6… 9-Nov-10 9… 1… 16-Nov-10 1… 1… 2… 23-Nov-10 2… 2… 30-Nov-10 3… Granite Woodland Saline Heath 6… Mallee 3… 7-Dec-10 Thicket 27-Dec-12 35 20-Dec-12 30 13-Dec-12 25 6-Dec-12 20 29-Nov-12 15 22-Nov-12 10 15-Nov-12 5 8-Nov-12 0 1-Nov-12 20 25-Oct-12 PAW (mm) 15 18-Oct-12 10 11-Oct-12 5 4-Oct-12 0 27-Sep-12 15-Jun-10 2012 20-Sep-12 60 13-Sep-12 50 6-Sep-12 40 30-Aug-12 30 23-Aug-12 20 16-Aug-12 7… 1… 1… 1… 1… 2… 2… 2… 1… 4… 7… 1… 1… 1… 1… 2… 2… 2… 3… 3… 6… 9… 1… 1… 1… 2… 2… 2… 3… 2… 5… 8… 1… 1… 1… 2… 2… 2… 2… 2… 5… 8… 1… 1… 1… 2… 2… 2… 2… 1… 4… 7… 1… 1… 1… 1… 2… 2… 2… 1… 4… 7… 1… 1… 1… 1… 2… 2… 2… 3… 10 9-Aug-12 0 2-Aug-12 40 26-Jul-12 35 19-Jul-12 rain (mm) PAW (mm) 30 12-Jul-12 25 5-Jul-12 20 28-Jun-12 15 21-Jun-12 5 14-Jun-12 10 0 127 Figure 7.1 Daily moisture balance of soils in each Landscape Division (mm of plant available water – PAW – coloured lines) and rainfall (black bars) for years with low winter rainfall. Note that the same colour key applies to both. rain (mm) 7-Jun-12 9… 1… 14-Dec-10 1… 1… 21-Dec-10 2… 2… 28-Dec-10 2… 3… 7.2.2 Average Winter Rainfall – 2009, 2011 In 2009, Heath, Thicket and Woodland held water for similar periods, while the remaining Landscape Divisions had different patterns of soil moisture presence (Figure 7.2). The overall quantity of PAW was greater than in the low rainfall years, but only Granite held > 35 mm at any one time. Heath, Thicket and Woodland areas all accumulated PAW on 20 June 2009 and held it for 43–44 days, drying on 2–3 August, prompted by sub-2mm falls after 22 July. These soils then remained dry for nearly two weeks, before re-wetting following 12.8 mm of rain on 15 August; responses were likely aided by 11.8 mm that fell between 6–14 August. Heath and Thicket areas finally dried on 24 August, while Woodland areas held water for one more day – these areas yielded PAW for a total of 52 to 54 days. Saline and Mallee areas held PAW for far shorter periods. Saline soils accumulated water on 20 June but did not re-wet after drying on 2 August, despite reasonable falls in the first half of August. While 25.2 mm of rain fell during the first eight days of the model period, Mallee soils did not accumulate water until 30 June 2009 after a 12.2 mm rainfall event. This water was retained for just two days before the profile dried, and subsequently re-wetted after 21.2 mm of rain fell over two days in early July. Mallee soils were devoid of water by 28 July, having yielded PAW for just 20 days. Despite some moderately sized falls later in the year (14 mm on 11–12 September, 12.6 mm on 27 October 2009, 18.2 mm on 13 November), none of these Landscape Divisions accumulated PAW; higher evaporation rates meant that input did not sufficiently recharge profiles. In contrast, Granite soils held water continuously for 121 days, from 20 June until they dried on 19 October 2009. The profile may have briefly re-wetted following the 27 October fall, but water would have been short-lived due to the increased evaporation rate characteristic of the warmer months. Subsequent rainfall events (18.2 mm on 13 November, 11 mm on 19 November) would have been insufficient to recharge a completely dry Granite profile (i.e. one that was not just devoid of PAW but had rather approached or reached its minimum volumetric water content), but may have elicited a weak response due to the short dry period that preceded the falls. 128 PAW (mm) 1… 1… 2… 2… 2… 3… 2… 5… 8… 1… 1… 2… 2… 2… 2… Mallee Granite Woodland Saline Heath 9 -0 2… ec 2… 9 -0 2… 28-Dec-11 ec 2… 9 -0 1… 21-Dec-11 ec 1… Thicket 8… 1… 14-Dec-11 9 -0 ec 5… 7-Dec-11 9 -0 ov 30-Nov-11 9 -0 ov 2… 23-Nov-11 9 -0 ov 1… 16-Nov-11 9 -0 ov 9-Nov-11 09 ct- 2-Nov-11 09 ct- 26-Oct-11 09 ct- 19-Oct-11 09 ct- 1… 120 9… 12-Oct-11 09 ct- 6… 2009 3… 9 -0 ep 3… 5-Oct-11 100 2… 9 -0 ep 2… 28-Sep-11 2011 2… 9 -0 ep 1… 1… 21-Sep-11 80 1… 9 -0 ep 9… 14-Sep-11 60 6… 9 -0 ug 3… 7-Sep-11 40 3… 9 -0 ug 2… 2… 31-Aug-11 20 2… 9 -0 ug 1… 24-Aug-11 9 -0 ug 1… 9 -0 ug 1… 17-Aug-11 9 l-0 Ju 1… 9 l-0 Ju 4… 7… 10-Aug-11 9 l-0 Ju 1… 9 l-0 Ju 2… 3-Aug-11 9 -0 un 2… 9 -0 un 2… 27-Jul-11 0 2… 20 1… 1… 20-Jul-11 15 1… 10 8… 13-Jul-11 20-Jun-09 22-Jun-09 24-Jun-09 26-Jun-09 28-Jun-09 30-Jun-09 2-Jul-09 4-Jul-09 6-Jul-09 8-Jul-09 10-Jul-09 12-Jul-09 14-Jul-09 16-Jul-09 18-Jul-09 20-Jul-09 22-Jul-09 24-Jul-09 26-Jul-09 28-Jul-09 30-Jul-09 1-Aug-09 3-Aug-09 5-Aug-09 7-Aug-09 9-Aug-09 11-Aug-09 13-Aug-09 15-Aug-09 17-Aug-09 19-Aug-09 21-Aug-09 23-Aug-09 25-Aug-09 27-Aug-09 29-Aug-09 31-Aug-09 2-Sep-09 4-Sep-09 6-Sep-09 8-Sep-09 10-Sep-09 12-Sep-09 14-Sep-09 16-Sep-09 18-Sep-09 20-Sep-09 22-Sep-09 24-Sep-09 26-Sep-09 28-Sep-09 30-Sep-09 2-Oct-09 4-Oct-09 6-Oct-09 8-Oct-09 10-Oct-09 12-Oct-09 14-Oct-09 16-Oct-09 18-Oct-09 20-Oct-09 22-Oct-09 24-Oct-09 26-Oct-09 28-Oct-09 30-Oct-09 1-Nov-09 3-Nov-09 5-Nov-09 5 5… 0 140 120 100 2… 2… 6-Jul-11 PAW (mm) rain (mm) 80 60 40 20 0 0 29-Jun-11 110 100 90 80 70 60 50 40 30 20 10 129 Figure 7.2 Daily moisture balance of soils in each Landscape Division (mm of plant available water – PAW – coloured lines) and rainfall (black bars) for years with average winter rainfall. Note that the same colour key applies to both. rain (mm) As in 2009, the 2011 model period was characterised by intermittent soil water availability. Saline, Thicket, Woodland and Heath areas had all received PAW on 29 June 2011 but dried out between 2 and 5 July because the 9.5 mm of rain received in the first three days was followed by just 0.7 mm over the subsequent nine-day period. These soils re-wetted after 14.5 mm of rain fell on 11–12 July and held PAW until 17–19 August. Only Woodland soils held water consistently during this second period – all other soils dried temporarily at least once, for up to four days at a time. As in 2009, Mallee soils did not accumulate PAW when all other soils did, but remained dry until almost a month later, when cumulative falls (58.8 mm between 29 June and 28 July) had sufficiently recharged the profile. Mallee soils subsequently dried on 2 August, after holding water for four days. None of these profiles accumulated PAW as a result of several > 10 mm falls that occurred from September to November. Granite soils held water consistently from 20 June to 27 November but, as a result of the rainfall distribution noted above, the profile very nearly dried out in early July. In contrast to all other Landscape Divisions, Granite soils held water for longer after the 2011 winter than they did after high winter rainfall in 1992 and 1998. This is due to the 2011 post-winter rainfall distribution, whereby the average winter total was followed by a few moderately sized falls in later months (12.6 mm on 19 September, 13 mm on 23 October, 12.6 mm on 4 November). The Granite profile was still holding water when these falls occurred and was thus able to respond strongly to them. These falls not only arrested PAW decline, but actually increased balances sufficiently to prolong the longevity of water. These falls would have elicited a much weaker response had they occurred more than a few days after Granite soils had dried, as the profile would need to recharge before PAW would accumulate. Only Granite and Thicket soils, the two coarsest profiles, responded when 105 mm of rain fell on 7 December 2011. The 5.4 mm received the previous day would have been largely ineffective due to higher atmospheric demand at this time of year. It is well-known that coarser profiles respond more readily to rainfall events as lower quantities of water are required for recharge (Figure 5.3; Moore 2001:86), but it is noteworthy that even such a large rainfall event was insufficient to recharge the remaining soils. Thicket soils had been dry for 112 days, yet still accumulated 41.6 130 mm of PAW as a result. The remaining 63.4 mm represents the quantity of water required to recharge the profile, plus the daily loss. The Thicket profile stored this water for just 13 days but would have dried slightly sooner were it not for the 11.6 mm fall on 13 December that temporarily slowed the rate of loss. The sand profile (from which Granite soils were modelled, after allowing for runoff – see Chapter 5.4.2) accumulated 4 mm of PAW as a result of the same fall, meaning 101 mm was consumed through recharge and daily loss. For the shallower Granite profile, just 46 mm would be required to recharge a completely dry profile – this could be achieved with 23 mm of rain. The Granite profile had only been devoid of PAW for nine days before the 105 mm fall, so even less water would be required for recharge. In any case, the Granite profile could not accept all the water available to it, due to the limited storage capacity of 100 mm. Granite soils stored this water until the end of the model period (31 December 2011) and, based on the rate of loss, would have held it for perhaps another eight days into 2012, presuming no further rainfall occurred. This isolated fall, as well as the subsequent 11.6 mm fall noted above, therefore resulted in the presence of PAW for a total of 33 days in Granite soils. 7.2.3 High Winter Rainfall – 1992, 1998 Apart from Granite, all of the Landscape Divisions demonstrated similar patterns of water-retention in 1992. They held water, rarely > 20 mm at any one time, from 13 June until late September/early October (Figure 7.3). All of the five Division’s soils temporarily dried at least three times, but Mallee soils dried on five separate occasions and for longer periods than the others. The three major drying events were each prompted by six- to twelve-day periods where rain was scarce, and daily falls were < 2 mm; PAW balances were not large enough to insulate profiles against drying under these circumstances. These profiles would likely have remained dry if not for 20 mm of rain that fell on 20 September, followed by 7.1 mm over the next four days. This brought the September rainfall total to 53.4 mm, while the mean is just 28.1 mm (BoM data). This extra post-winter input permitted all soils to re-wet for another five to eight days, before drying between 25 and 28 September. Thereafter, Mallee, Heath and Saline soils remained dry, having held water for a total of 75, 92 and 94 days respectively. Thicket and Woodland soils briefly re-wet after 10 mm of rain fell over two days in early October, and finally dried on 8 October 1992; they held water for 99 and 102 days respectively. Subsequent falls were too small (< 4 131 mm) and well-spaced to permit any of these profiles to accumulate PAW. In contrast, Granite soils held water consistently from 13 June to 3 November, due to the additional input they received from runoff. Their larger PAW balance, which reached the 100 mm storage capacity on several occasions, allowed them to retain PAW through the rainless periods that prompted all other profiles to dry. Nevertheless, even with extra input from runoff, the light falls later in 1992 were insufficient to recharge the Granite profile. In 1998, the same five Landscape Divisions were again closely grouped, as they were in 1992. However, in contrast to the earlier period, soils retained water continuously from 5 June until their final drying date. This probably resulted from rainfall distribution – a total of 60.4 mm of rain fell on 5–11 June, and all soils stored > 40 mm of PAW. These larger balances meant that all soils retained water even though only 16.4 mm of rain fell from 12 to 30 June 1998. Balances then increased due to strong rainfall in the first half of July (39.6 mm from 1–15 July) before gradually declining due to limited falls thereafter. All profiles dried within a week of each other: Saline and Heath on 29 July, and Thicket, Woodland and Mallee – in that order – between 2 and 4 August. Just 12 mm of rain fell over the first 15 days of August, so soils remained dry, rendering the later falls (22.2 mm on 28 and 29 August) insufficient to recharge the dry profiles. The well-spaced nature of late winter rainfall events meant these soils held water for a much shorter period than they had in 1992, despite the winter rainfall totals varying by < 40 mm. As in all other modelled periods, Granite areas held water for the longest. In this case PAW was available for 136 days, from 5 June to 18 October, over ten weeks after all other areas had dried. Subsequent rainfall events thereafter were too small (generally < 3 mm) and widely spaced to permit Granite soils to again accumulate PAW. 132 120 2… 2… 27-Jun-92 2… 1… 4-Jul-92 4… 7… 11-Jul-92 1… 1… 1… 18-Jul-92 1… 2… 25-Jul-92 2… 2… 1-Aug-92 3… 3… 6… 8-Aug-92 9… 1… 15-Aug-92 1… 1… 22-Aug-92 2… 2… 2… 29-Aug-92 3… 2… 5-Sep-92 5… 8… 1… 1992 12-Sep-92 1… 1… 19-Sep-92 2… 2… 26-Sep-92 2… 2… 3-Oct-92 2… 5… 8… 10-Oct-92 1… 1… 17-Oct-92 1… 2… 24-Oct-92 2… 2… 2… 31-Oct-92 1… 4… 7-Nov-92 7… 1… 14-Nov-92 1… 1… 1… 21-Nov-92 2… 2… 28-Nov-92 2… 1… 1… Heath 7… 12-Dec-92 Saline 4… Granite 5-Dec-92 Woodland 25-Dec-98 1… Mallee 18-Dec-98 1… 1… 20-Jun-92 Thicket 11-Dec-98 100 4-Dec-98 80 27-Nov-98 60 20-Nov-98 40 13-Nov-98 20 6-Nov-98 0 30-Oct-98 30 23-Oct-98 25 16-Oct-98 20 9-Oct-98 15 2-Oct-98 5 25-Sep-98 10 18-Sep-98 13-Jun-92 1998 11-Sep-98 0 4-Sep-98 120 28-Aug-98 100 21-Aug-98 PAW (mm) rain (mm) 80 14-Aug-98 60 7-Aug-98 40 31-Jul-98 5… 8… 1… 1… 1… 2… 2… 2… 2… 2… 5… 8… 1… 1… 1… 2… 2… 2… 2… 1… 4… 7… 1… 1… 1… 1… 2… 2… 2… 3… 3… 6… 9… 1… 1… 1… 2… 2… 2… 3… 3… 6… 9… 1… 1… 1… 2… 2… 2… 3… 2… 5… 8… 1… 1… 1… 2… 2… 2… 2… 2… 5… 8… 1… 1… 1… 2… 2… 2… 2… 20 24-Jul-98 0 17-Jul-98 30 10-Jul-98 25 3-Jul-98 20 26-Jun-98 15 19-Jun-98 5 5-Jun-98 12-Jun-98 10 0 133 Figure 7.3 Daily moisture balance of soils in each Landscape Division (mm of plant available water – PAW – coloured lines) and rainfall (black bars) for years with high winter rainfall. Note that the same colour key applies to both. Title rain (mm) 1… 1… 19-Dec-92 1… 2… 26-Dec-92 2… 2… 3… 7.2.4 Summary Figures 7.1–7.3 indicate that, for most of the study area, soil water was short-lived, and its presence centred on winter. Initial winter falls recharged soil profiles and provided the impetus for PAW to accumulate, while the size and distribution of subsequent falls dictated the quantity, longevity and consistency of soil water. There is a simple relationship between the quantity of winter rainfall and the longevity of soil water – most soils stored water for longer periods after high rainfall winters than after low rainfall winters (Table 7.1). Nevertheless, the timing was still influential. For example, in 1998 (a high winter rainfall year) soils were mostly dry by early August because over half the monthly rainfall total fell over a few days in late August, when soils had been dry for too long to respond. Similarly, most soils dried at some point during the winter of 1992 despite this year having the highest winter rainfall total (n=217.1 mm). Falls were light and widely spaced, so PAW balances were too low to prevent profiles drying during short rainless periods. Truly isolated falls, those that occur when profiles have been dry for long periods of time, were largely ineffective, since input must be sufficient to recharge the profile before water can be made available to plants (Figure 5.3). Granite and Thicket areas required 46 and 63.4 mm of water to recharge their respective soil profiles. For Granite soils, recharge could be achieved with a fall of around 25 mm, due to the extra input received from runoff. Recharge values are unavailable for other Landscape Divisions, but must exceed 100 mm due to the lack of response in December 2011. Table 7.1 Longevity of PAW (days) in modelled years, across all six Landscape Divisions. Landscape Division Granite Heath Woodland Mallee Saline Thicket Low winter rainfall Average winter rainfall High winter rainfall 2010 2012 2009 2011 1992 1998 45 17 74 21 121 52 177 41 144 92 136 54 20 1 21 1 54 20 45 4 102 75 59 60 7 10 11 16 43 52 32 50 94 99 54 58 134 Overall, Mallee areas had the poorest soil water potential. They took the longest to accumulate PAW, were generally the first to lose it, and were most likely to dry out during short rainless periods. In periods of low or average winter rainfall, soil water was present for just a single day or a few weeks (Table 7.1). Water was more plentiful after high winter rainfall, when Mallee soils held water for up to 11 weeks, but the PAW balance still rarely exceeded 20 mm at any one time (Figure 7.1–7.3). Saline soils generally held water for longer than Mallee profiles, but also frequently dried out during short rainless periods when superior soils retained water. Like Mallee, these soils were also less likely to re-wet once dried. Saline areas may yield soil water for 1–13 weeks, depending on the quantity of winter rainfall. In some parts of the Saline Landscape Division, sodic soils/subsoils may render this stored water unpotable – indeed when Roe (1836:299) dug for water near a salt lake he encountered only brackish, undrinkable water. Thicket retained water for slightly longer than Saline soils, preserving PAW for up to 14 weeks. Heath and Woodland areas out-performed Thicket soils in lower rainfall periods, holding water for an additional week or more, but all three areas performed similarly under higher rainfall conditions. However, Thicket soils could accumulate PAW from large isolated falls > 65 mm. The smaller amount of water required to recharge coarsely textured Thicket profiles allowed these soils to respond, even after lengthy dry periods, when the finer textured Heath and Woodland soils cannot. Granite areas represented the best prospects for soil moisture under all rainfall conditions, primarily due to increased input through runoff. This increased the consistency, quantity and longevity of soil moisture. All other soil profiles rarely held > 10–40 mm PAW, with the upper limit occurring during high rainfall winters. In contrast, Granite soils stored 40–50 mm of PAW during low winter rainfall years (2010, 2012), and achieved their upper storage limit (100 mm) several times during all other modelled years. Even when winter rainfall was limited, Granite soils held water for at least 6.5 weeks; under average or high rainfall conditions these soils could hold water for up to 22 weeks. Their limited profile depth and coarse texture, combined with extra input from runoff, means that Granite soils accumulated PAW after isolated falls > 25 mm. They benefit more from smaller rainfall events that are spread over a longer period, as occurred in 2011 (Figure 7.2) than closely-spaced 135 heavier falls, as the profile quickly reaches its water-storage capacity under higher input conditions. 7.3 ROCK STRUCTURES Four types of rock structure are considered here: gnammas, pans, runoff gnammas and runoff pans. Each exhibits unique characteristics of water reception and retention. The availability of water in each structure is described for periods of low, average and high winter rainfall, using the same six model periods as above. 7.3.1 Low Winter Rainfall – 2010, 2012 Rainfall was infrequent for the first four months of 2010, so all rock structures only held small quantities of water (< 20 mm) for short periods, no more than nine days, centred around specific March and April rainfall events (Figure 7.4). Pans only responded to the heaviest falls due to the lack of runoff combined with fairly high evaporation rates. By mid-May, falls were greater and evaporation rates lower, so structures held water for 6–16 days. In contrast, runoff gnammas received enough input to retain water consistently through to winter and well beyond. Winter rains commenced when 14 mm fell over two days in mid-June but falls thereafter were fairly irregular. The season was characterised by long periods where falls are absent, or very light (< 1 mm), punctuated by a few heavier events. Pans held water inconsistently throughout winter for no more than 19 days at a time. Gnammas and runoff pans accumulated large enough balances from early to mid-July falls (30.5 mm over five days) to retain water throughout subsequent dry periods, when just 3.3 mm of rain fell over four weeks. Runoff pans dried on 29 August and gnammas on 4 September, after continuously holding water for 51 and 57 days, respectively. Pans, runoff pans and gnammas all remained dry throughout the remainder of September and October, as falls were light and widely spaced; only 5.9 mm of rain fell during that two-month period. In contrast, runoff gnammas accumulated a large enough water balance to retain water throughout much of this period. They dried on 22 October after holding water continuously for more than five months. All structures briefly re-wet in mid-November when 8.4 mm of rain fell, but none held water for more than five days. Only runoff gnammas responded to two lighter falls in 136 2… 22-Jan-10 2… 29-Jan-12 2… 29-Jan-10 2… 5-Feb-12 5… 5-Feb-10 5-… 12-Feb-12 1… 12-Feb-10 1… 19-Feb-12 1… 19-Feb-10 1… 26-Feb-12 2… 26-Feb-10 2… 4-Mar-12 4… 5-Mar-10 5-… 11-Mar-12 1… 12-Mar-10 1… 18-Mar-12 1… 19-Mar-10 1… 25-Mar-12 2… 26-Mar-10 2… 1-Apr-12 1… 2-Apr-10 2-… 9-Apr-10 9-… 1… 16-Apr-10 1… 22-Apr-12 2… 23-Apr-10 2… 29-Apr-12 2… 30-Apr-10 3… 6-May-12 6… 7-May-10 7-… 13-May-12 1… 14-May-10 1… 20-May-12 2… 21-May-10 2… 27-May-12 2… 28-May-10 2… 3-Jun-12 3… 4-Jun-10 4-… 10-Jun-12 1… 11-Jun-10 1… 17-Jun-12 1… 18-Jun-10 1… 24-Jun-12 2… 1-Jul-12 1… 8-Jul-12 8… 15-Jul-12 1… 22-Jul-12 2… 29-Jul-12 2… 5-Aug-12 5… 12-Aug-12 1… 19-Aug-12 1… 26-Aug-12 2… 2-Sep-12 2… 9-Sep-12 16-Sep-12 23-Sep-12 30-Sep-12 7-Oct-12 14-Oct-12 21-Oct-12 28-Oct-12 4-Nov-12 11-Nov-12 18-Nov-12 9-Jul-10 9-… 16-Jul-10 1… 23-Jul-10 2… 30-Jul-10 3… 6-Aug-10 6-… 13-Aug-10 1… 20-Aug-10 2… 27-Aug-10 2… 3-Sep-10 3-… 10-Sep-10 1… 17-Sep-10 1… 1… 2… 24-Sep-10 2… 1-Oct-10 1-… 8-Oct-10 8-… 3… 7… 1… 15-Oct-10 2… 22-Oct-10 2… 29-Oct-10 4… 5-Nov-10 1… 12-Nov-10 1… 19-Nov-10 26-Nov-10 2… 3-Dec-10 9… 10-Dec-10 1… 16-Dec-12 17-Dec-10 2… 23-Dec-12 24-Dec-10 3… 30-Dec-12 31-Dec-10 2… 2… 5-… 1… 1… 2… 3-… 1… 1… 2… 3… Gnamma 2… 1… Pan 9-Dec-12 2-… Runoff gnamma 2-Dec-12 2… 2-Jul-10 Runoff pan 25-Nov-12 9… 25-Jun-10 2010 8… 2012 8-Apr-12 15-Apr-12 December (3.2 mm on 21 December, 2.6 mm on 25 December), retaining water for a 1… 22-Jan-12 single day in each case, due to higher atmospheric demand. 8-… 15-Jan-10 80 8-Jan-10 1… 60 8… 40 8-Jan-12 15-Jan-12 20 0 20 15 10 5 0 1-… 100 depth of water (mm) 1-Jan-10 120 rain (mm) 180 160 140 120 100 80 60 40 1-Jan-12 1… 20 0 40 35 30 25 20 15 5 10 0 137 Figure 7.4 Daily water balance (mm) in rock structures (coloured lines) and rainfall (black bars) for years with low winter rainfall. Note that the same colour key applies to both. depth of water (mm) rain (mm) As in 2010, the beginning of 2012 was characterised by limited water availability – each structure held water for short periods centred on rainfall events large enough to offset daily loss through evaporation. Runoff gnammas held water for up to nine days at a time, while all other structures dried no more than three days after rain. Pans only responded to larger falls, and held water for just one day after 10.2 mm of rain fell in mid-March. Pans, runoff pans and gnammas held water for 7–16 days in May, while runoff gnammas retained water consistently until winter. Winter rainfall distribution differed from 2010, however. A total of 26.8 mm of rain fell over a week in early June 2012, but subsequent falls were more frequent than in 2010, and the intervening dry periods shorter. As a result, most structures consistently held water throughout winter (Figure 7.4). In contrast, pans dried after holding water for 23 consecutive days in June, as the water balance (< 15 mm) was insufficient to survive the late June and early July dry period. For the remainder of winter, pans held water for less than six days at a time, while runoff pans and gnammas retained water consistently until 31 August and 13 September, respectively. All three structures then re-wetted several times in September, for one or two days at a time. Subsequent larger (29.4 mm 31 October–2 November; 60.6 mm 25–29 November) allowed them to hold water for 4–11 days in the first instance, and 7–18 days after the later fall. Due to the considerable balance accumulated during winter (> 110 mm), runoff gnammas were still holding water when the October and November falls occurred. This increased the water balance sufficiently that they still held a small quantity of water on 31 December, 241 days after they began to fill; the remaining 22 mm would have been lost over the first five days of January 2013. Were it not for the larger isolated falls noted above, which are well above the November mean (19.2 mm – BoM data), runoff gnammas would probably have dried in early to mid-November. Due to the higher annual rainfall total, created by heavier and more frequent rain outside the winter period, all structures held water for considerably longer than they had in 2010. 7.3.2 Average Winter Rainfall – 2009, 2011 In early 2009, a few moderate to large falls helped all rock structures accumulate water at least once. All structures except pans held water for up to three days in midJanuary after an 8 mm fall (Figure 7.5). On 28 January, 21 mm of rain fell, followed 138 by 10.4 mm on 1 February. Pans and runoff pans dried out between these two events, and only held water for one to three days at a time. In contrast, gnammas and runoff gnammas did not dry out between these closely spaced falls. Nevertheless, they still only held water for fairly short periods, just six and 14 days, respectively, highlighting the impact of greater atmospheric demand in the summer months. Most rock structures held water intermittently, for up to 15 days at a time, in May–early June, but runoff gnammas accumulated enough to remain wet until winter rains commenced. From 19 June, rainfall events became more frequent and a few large falls (> 12 mm) in late June and early July, combined with subsequent smaller falls, allowed all structures to build up large stores of water. This permitted them to retain water throughout winter, even when rainfall was scarce. Pans finally dried on 1 September, but regained water three times in September–October, for no more than four days at a time, after daily or cumulative falls of 8–15 mm. Runoff pans dried on 10 October, and briefly re-wetted in late October, while gnammas retained water consistently until 4 November. All rock structures briefly accumulated water as a result of two larger November falls (18.2 mm on 13 November, 11 mm on 19 November), and held this water for up to five days. Subsequent falls were too light to exceed the daily loss through evaporation, so pans, runoff pans and gnammas all remained dry for the rest of the year. Runoff gnammas, however, were still holding water on 31 December, a total of 224 consecutive days. This water would have been lost by 11 January the following year, presuming no further input occurred. Unlike 2012, the November 2009 rains were only slightly above average (31.4 mm). Therefore, the increased water longevity in 2009 was created not by isolated falls, but due to higher winter rainfall and the large store of water (> 265 mm) this created in runoff gnammas. During first four months of 2011 water was present intermittently, for up to eight days at a time in runoff gnammas, while other structures never yielded water for more than four consecutive days. In the earliest months, pans only responded to falls above 10 mm (15.6 mm on 6 January, 11.4 mm on 30 January), and held water for a single day. They responded to smaller falls thereafter, when evaporation rates were lower, but held this water for equally short periods. By mid-May, runoff gnammas had accumulated enough water to prevent them from drying during short rainless periods. All other structures still held water intermittently throughout June, due to low rainfall 139 (total 18.6 mm). From late June, gnammas and runoff pans held water consistently, while pans filled intermittently until mid-July. Thereafter, rainfall events became heavier and more frequent – 70.9 mm fell between 11 and 31 July. This input allowed pans to hold water for 51 consecutive days, before they dried on 31 August. Subsequent wet periods were short (up to six days) and centred on falls of 5–15 mm. Runoff pans did not dry until 13 October, after holding water for 110 consecutive days. A total of 35.5 mm of rain fell on 18–23 October, which increased the monthly total to more than twice the average (49.1 mm, mean 21.6 mm). This permitted pans and runoff pans to accumulate water for up to six and 18 days at a time, respectively, before they both dried in early November. These late October falls also increased the water balance in gnammas and runoff gnammas, which were still holding winter water. Gnammas finally dried on 17 November, after holding water for 145 consecutive days. At this time, runoff gnammas were still holding > 250 mm of water. On 7 December 2011, 105 mm of rain was received – this is the single largest individual fall ever recorded by the Hyden meteorological station. This input allowed water to accumulate in all previously dry rock structures and increased the already considerable store of water in runoff gnammas. Nonetheless, pans dried after 12 days, despite another > 20 mm of rain that fell over the following week. Runoff pans dried after 15 days because their limited depth meant they could not accept additional input generated by runoff. Due to these falls and another 7.2 mm that fell on 31 December, three of the four rock structures held water on the final day of the model period. Runoff pans would have held this water for a day, while gnammas would have preserved their water supply until about 6 January 2012. Runoff gnammas held > 350 mm on 31 December, so would have yielded water for the first few months of 2012. Had the 7 December fall not occurred, these structures would probably have held water until early to mid-January. 140 depth of water (mm) 1… 22-Jan-11 2… 22-Jan-09 2… 29-Jan-11 2… 29-Jan-09 2… 5-Feb-11 5… 5-Feb-09 5… 12-Feb-11 1… 12-Feb-09 1… 19-Feb-11 1… 19-Feb-09 1… 26-Feb-11 2… 26-Feb-09 2… 5-Mar-11 5… 5-Mar-09 5… 12-Mar-11 1… 12-Mar-09 1… 19-Mar-11 1… 19-Mar-09 1… 26-Mar-11 2… 26-Mar-09 2… 2-Apr-11 2… 2-Apr-09 2… 9-Apr-11 9… 9-Apr-09 9… 16-Apr-11 1… 16-Apr-09 1… 23-Apr-11 2… 23-Apr-09 2… 30-Apr-11 3… 30-Apr-09 3… 7-May-11 7… 7-May-09 7… 14-May-11 1… 14-May-09 1… 21-May-11 2… 21-May-09 2… 28-May-11 2… 28-May-09 2… 4-Jun-11 4… 4-Jun-09 4… 11-Jun-11 1… 11-Jun-09 1… 18-Jun-11 1… 18-Jun-09 1… 25-Jun-11 2… 25-Jun-09 2… 2-Jul-11 2… 2-Jul-09 2… 9-Jul-11 9… 9-Jul-09 9… 16-Jul-11 1… 16-Jul-09 1… 23-Jul-11 2… 23-Jul-09 2… 30-Jul-11 3… 30-Jul-09 3… 6-Aug-11 6… 6-Aug-09 6… 13-Aug-11 1… 13-Aug-09 1… 20-Aug-11 2… 20-Aug-09 2… 27-Aug-11 2… 27-Aug-09 2… 3-Sep-11 3… 3-Sep-09 3… 10-Sep-11 1… 10-Sep-09 17-Sep-11 1… 17-Sep-09 24-Sep-11 2… 24-Sep-09 1-Oct-11 1… 1-Oct-09 8-Oct-11 8… 8-Oct-09 15-Oct-11 1… 15-Oct-09 22-Oct-11 2… 29-Oct-11 2… 5-Nov-11 5… 12-Nov-11 1… 19-Nov-11 1… 24-Dec-11 31-Dec-11 1… 2… 3… 19-Nov-09 26-Nov-09 3-Dec-09 10-Dec-09 17-Dec-09 24-Dec-09 31-Dec-09 1… 8… 1… 2… 2… 5… 1… 1… 2… 3… 1… 1… 2… 3… Gnamma 17-Dec-11 1… 12-Nov-09 2… Pan 10-Dec-11 3… 5-Nov-09 1… Runoff gnamma 3-Dec-11 2… 29-Oct-09 1… Runoff pan 26-Nov-11 22-Oct-09 2009 15-Jan-09 300 1… 250 8… 15-Jan-11 200 1… 8-Jan-09 150 1-Jan-09 8… 50 0 25 20 15 5 10 0 1… 8-Jan-11 2011 1-Jan-11 100 rain (mm) 550 500 450 400 350 300 250 200 150 100 0 50 110 100 90 80 70 60 50 40 30 20 10 0 141 Figure 7.5 Daily water balance (mm) in rock structures (coloured lines) and rainfall (black bars) for years with average winter rainfall. Note that the same colour key applies to both. depth of water (mm) rain (mm) 7.3.3 High Winter Rainfall – 1992, 1998 In 1992, early water-holding events were rare – just one occurred, when 20.8 mm of rain fell over three days in February (Figure 7.6). Pans, runoff pans and gnammas held water for up to three days, while runoff gnammas retained water for 15 days, aided by subsequent 3–4 mm falls that slowed the rate of loss. These falls were insufficient to replenish supplies in the other structures. In mid-March, however, 103.7 mm of rain fell over eight days, and all structures accumulated water. These falls were sufficient for runoff gnammas to retain moisture until winter and beyond. Pans held water for 13 days but dried for one day between the initial 33 mm input (13–16 March) and the subsequent 70.7 mm (17–20 March), again highlighting the rapid rate of water-loss in this type of rock structure. Runoff pans and gnammas dried by early April and early May, respectively. Thereafter, dried structures rewetted for short periods centred on various rainfall events, holding water for up to 15 days at a time. Early June was dry, but water was more abundant after strong midJune rainfall totalling 66.7 mm. As a result, all structures held enough water to prevent them drying during July. August and September falls were well above average, 110.7 and 53.4 mm respectively – this permitted all rock structures to consistently hold water throughout these months. Pans dried on 2 October and runoff pans on 22 October, after holding water for 113 and 133 consecutive days, respectively. Runoff pans could not take full advantage of these falls due to their limited storage capacity, which was reached 22 times between July and September. Gnammas dried on 5 December, as the October and November falls barely slowed the rate of water loss; these falls were so widely spaced that no pans re-wetted. In contrast, runoff gnammas were holding almost 600 mm of water in mid-October. That permitted these structures to retain water until 31 December, a total of 294 consecutive days or over 80% of the year. The remaining > 355 mm would have been lost within the first few months of 1993, presuming average rainfall was received. 142 depth of water (mm) 22-Jan-92 2… 29-Jan-98 2… 29-Jan-92 2… 5-Feb-98 5… 5-Feb-92 5… 12-Feb-98 1… 12-Feb-92 1… 19-Feb-98 1… 19-Feb-92 1… 26-Feb-98 2… 26-Feb-92 2… 5-Mar-98 5… 4-Mar-92 4… 12-Mar-98 1… 11-Mar-92 1… 19-Mar-98 1… 18-Mar-92 1… 26-Mar-98 2… 25-Mar-92 2… 2-Apr-98 2… 1-Apr-92 1… 9-Apr-98 9… 8-Apr-92 8… 16-Apr-98 1… 15-Apr-92 1… 23-Apr-98 2… 22-Apr-92 2… 30-Apr-98 3… 29-Apr-92 2… 7-May-98 7… 6-May-92 6… 14-May-98 1… 13-May-92 1… 21-May-98 2… 20-May-92 2… 28-May-98 2… 27-May-92 2… 4-Jun-98 4… 3-Jun-92 3… 11-Jun-98 1… 10-Jun-92 1… 18-Jun-98 1… 17-Jun-92 1… 25-Jun-98 2… 24-Jun-92 2… 2-Jul-98 2… 1-Jul-92 1… 9-Jul-98 9… 8-Jul-92 8… 16-Jul-98 1… 15-Jul-92 1… 23-Jul-98 2… 22-Jul-92 2… 30-Jul-98 3… 29-Jul-92 2… 6-Aug-98 6… 5-Aug-92 5… 13-Aug-98 1… 12-Aug-92 1… 20-Aug-98 2… 19-Aug-92 1… 27-Aug-98 2… 26-Aug-92 2… 3-Sep-98 3… 2-Sep-92 2… 10-Sep-98 1… 9-Sep-92 9… 17-Sep-98 1… 16-Sep-92 1… 24-Sep-98 2… 23-Sep-92 2… 1-Oct-98 1… 30-Sep-92 3… 8-Oct-98 8… 7-Oct-92 7… 15-Oct-98 1… 14-Oct-92 2… 21-Oct-92 2… 28-Oct-92 5… 4-Nov-92 1… 11-Nov-92 1… 18-Nov-92 2… 25-Nov-92 3… 2-Dec-92 1… 9-Dec-92 1… 16-Dec-92 2… 23-Dec-92 3… 30-Dec-92 22-Oct-98 29-Oct-98 5-Nov-98 12-Nov-98 19-Nov-98 24-Dec-98 31-Dec-98 4… 1… 1… 2… 2… 9… 1… 2… 3… Gnamma 17-Dec-98 2… Pan 10-Dec-98 2… Runoff gnamma 3-Dec-98 1… Runoff pan 26-Nov-98 1992 2… 700 1… 22-Jan-98 600 15-Jan-92 500 8… 1… 400 8-Jan-92 15-Jan-98 300 1-Jan-92 8… 200 0 35 30 25 20 15 5 10 0 1… 8-Jan-98 1998 1-Jan-98 1… 100 rain (mm) 400 350 300 250 200 150 100 50 0 35 30 25 20 15 5 10 0 143 Figure 7.6 Daily water balance (mm) in rock structures (coloured lines) and rainfall (black bars) for years with high winter rainfall. Note that the same colour key applies to both. depth of water (mm) rain (mm) No rain fell in the first two months of 1998, so the first opportunity for any rock structure to accumulate water was when 4 mm of rain fell on 2 March, followed by 12.6 mm on 10 March. Water was only present for one to two days after the first fall, but for up to a week after the second. Water was available for longer periods in April as a result of 15.4 mm of rain that fell on 8–10 April, followed by 37 mm on 16–17 April. Pans held water for eight days, while runoff pans and gnammas both yielded water for around three weeks. The strong April rainfall allowed runoff gnammas to hold water consistently until winter, but was only present intermittently in other structures until 60.4 mm of rain fell on 5–11 June. They retained this water for at least three months, due to average July and August rainfall. Pans dried on 24 August after holding water for 80 consecutive days, but re-wetted for another 11 days soon after. The inconsistent presence of water resulted from the distribution of August rainfall, as over half of the monthly total fell on 27–31 August. Runoff pans dried on 9 October after holding water for 126 days. Again, longevity was impeded by the 100 mm storage capacity, which was reached on 13 separate occasions before they dried. Gnammas retained water for nearly a month longer, until 6 November, while runoff gnammas still held approximately 100 mm of water on 31 December. This would have been lost by 21 January 1999, presuming no additional input occurred. 7.3.4 Summary The relationship between annual rainfall total and overall longevity of water in rock structures is simple (Table 7.2). Rain is more effective in winter due to lower evaporative demand, but water can accumulate fairly easily at other times of year, as long as falls exceed the daily loss for a particular structure. Therefore, rainfall distribution is less significant than overall quantity. Rain is generally scarce in the beginning of each year, but structures may hold water for short periods centred on larger rainfall events; this water is generally very short-lived due to higher evaporation rates. Water is more consistently present from May–August when falls become heavier and more frequent and evaporation rates are lower. As winter rainfall increases, so too does the quantity, longevity and consistency of water in rock structures. When winter rainfall is low, most structures will still preserve water, albeit inconsistently, for most of the season. When winter rainfall is average or higher, structures hold water consistently throughout winter, and frequently a month or more beyond. 144 Post-winter longevity is dictated by the water-balance remaining at the end of winter, as well as the size and distribution of subsequent rainfall events. Structures frequently accumulate water in the later in the year, from isolated falls or short periods of sustained rainfall, although that water is rarely preserved for more than a week. Table 7.2 Winter and annual rainfall totals (mm) for each of the modelled years, and the longevity of water (days) in each rock structure. Years have been arranged by annual rainfall total (low to high). Note that runoff gnammas frequently held water from the end of one year and into the next – this is not reflected in the longevity figures. Year Annual rainfall Winter rainfall Gnammas Pans Runoff gnammas Runoff pans 2010 142.1 69.5 108 43 183 95 2012 280 84.2 155 60 257 139 2009 295.9 150.1 169 92 244 142 1998 324.6 178 202 106 277 168 2011 472.4 126.4 212 99 264 181 1992 514.1 217.1 258 141 311 190 Pans were consistently the poorest performing rock structures, as they held water for the shortest time and in the smallest quantity. They never reached their 100 mm storage capacity due to the lack of runoff combined with a high evaporation rate that nears atmospheric demand. Water-holding periods were closely centred on heavy or sustained rainfall events. Pans were most likely to dry out during short rainless periods, but could hold water consistently through winter, and up to a month beyond, if rainfall was average or better. When winter rainfall was low, water was present for as little as six weeks. Isolated falls were largely ineffective – even the largest (105 mm) resulted in water being present for less than two weeks. Runoff pans held water for longer periods, and more consistently than pans that did not receive runoff. Water was present throughout winter in all but the lowest rainfall year (2010) and runoff pans generally preserved winter water for a month longer than pans. Runoff was most valuable in low rainfall years, as storage capacity was quickly reached under average or high input conditions. This limited the effectiveness of large isolated falls, so the 105 mm fall was preserved for only 16 days. Runoff pans held water for almost 14 weeks in 2010 and for just over half of the year under higher rainfall conditions (1992), when the storage capacity was a stronger constraint. 145 Gnammas consistently held water for longer than pans and runoff pans, due to the lower evaporation rate. This improved performance was especially evident under higher rainfall conditions – when runoff pans were limited by their low storage capacity – which permitted gnammas to retain water for nearly 37 weeks in 1992, more than two months longer than runoff pans. Even under low rainfall conditions, gnammas held water for 15 to 22 weeks of the year. Runoff gnammas always yielded the greatest quantity of water, and held it for the longest period, due to the due to the additional input from runoff combined with the lower evaporation coefficient. Their large water balances prevented them from drying out during long rainless periods when other structures lost their water. As a result, in five of the six years modelled, runoff gnammas continuously held water from before winter until 31 December; they were generally dry for just two to four months a year. Runoff gnammas could also accumulate water from small isolated falls and preserve heavier falls for around two months. 7.4 SUMMARY OF WATER AVAILABILITY IN THE STUDY AREA Of the 88 years for which seasonal data are available (the Hyden station collected no weather information in 1947), 63 exhibited average winter rainfall. This equates to roughly 72% of the sample, or seven years in every ten (Figure 7.7a). Therefore, the water availability during those years represents the normal distribution of potable water in the study area, if unusually large pre- or post-winter falls are discounted (see Chapter 5.4.3). In average rainfall winters, water is available in most Landscape Divisions, but the timing varies (Figure 7.8). At the beginning of winter rain is sporadic, so water is only available in Granite areas with gnammas, runoff gnammas or runoff pans, which begin to accumulate water in mid-May. When rain becomes heavier and more frequent, generally in mid- to late June, soil water is available in most Landscape Divisions. Mallee profiles, however, accumulate water later in the year and hold it, intermittently, for a few weeks at best. Heath, Saline, Thicket and Woodland soils hold water fairly consistently until mid-August, but the generally low PAW balances mean these soils may temporarily dry out during prolonged periods of limited rainfall. By late August, water is only available in the Granite Landscape Division. Where present, pans will hold water until the end of winter, and runoff pans until mid-October. All Granite soils will hold PAW until late October. Beyond this, 146 water is only preserved at Granite locations housing gnammas and runoff gnammas. The former are dry by early November, while runoff gnammas can hold water until January–February the following year. 180 70 160 60 140 50 no. falls no. winters 120 40 100 80 30 60 20 40 10 20 0 0 low rainfall average rainfall high rainfall 15-24.9 25-34.9 35-44.9 45-54.9 55-64.9 65-74.9 75-84.9 85-94.9 95-104.9 ≥ 105 fall size (mm) Figure 7.7 Frequency of different rainfall scenarios in the study area from 1929–2017. A (left): number of low, average and high winter rainfall years. B (right): frequency of variously sized isolated falls. Based on Bureau of Meteorology data for the Hyden station (010568); note no data were recorded in 1947. High and low rainfall winters are far less frequent, accounting for just 11% and 17% of the sample, respectively, or around one or two years in every ten. Most soils and rock structures begin accumulating water at the same time as in average winter rainfall years, and dry out in the same order, but the longevity of their stored water is vastly different. In low rainfall winters, Mallee areas only retain water for a day, and Heath, Saline, Thicket and Woodland soils and profiles are dry by early July. Depending on rainfall distribution, the timing of water-holding periods may vary, but water will be consistently present unless rainfall events are interrupted by abnormally long drier periods, as in 2010 (see Chapter 7.2.1). Thereafter, water is again confined to Granite areas, but the longevity and consistency of its water retention is reduced. Pans hold water intermittently through most of winter, and even gnammas and runoff pans may dry temporarily if rainfall events are too widely spaced. Granite soils are dry by mid-August, and runoff gnammas before the end of spring. In high rainfall winters, Heath, Saline, Thicket and Woodland soils all retain soil water into early September, as does Mallee, in contrast to low and average winter rainfall years. Granite soils dry in late October and gnammas hold water until the end of November. Larger water balances permit runoff gnammas to retain water until early to midFebruary. Even under high rainfall conditions, then, Granite is the only Landscape Division that regularly preserves winter rains beyond early September (Figure 7.8). 147 Figure 7.8 Generalised water availability in soils and rock structures in each Landscape Division, in average (top), low (middle) and high (bottom) winter rainfall years. Note that Heath, Mallee, Saline, Thicket and Woodland areas only provide soil water. Solid lines indicate the generally consistent presence of water over the specified period, i.e. being dry for no more than a few days at a time. Dashed lines indicate when water may be available – its presence/absence depends on the timing of rainfall events. Unusually large pre- or post-winter falls have been voided (see Chapter 5.4.3), as they mask the normal longevity of water from winter rain. Water-holding periods for average (2009, 2011) and high (1992, 1996) winter rainfall years are based on the mean of both years. For low winter rainfall years (2010, 2012), only 2012 data was used as they were deemed more broadly representative of low rainfall conditions. In winter 2010 a few large falls were interrupted by three weeks of limited rainfall, causing soils to dry out when they normally would not have. Mallee soils are not pictured for low rainfall conditions as they held water for just a single day. 148 It is also important to consider isolated falls of moderate size (≥ 15 mm) that occur outside winter. A total of 266 such falls was recorded at Hyden between 1929 and 2017. Frequency declines as fall size increases; more than 60% of the sample comprises falls in the smallest category (15–24.9 mm), while falls of ≥ 75 mm made up just 2% of the sample (Figure 7.7b). In most cases, water would accumulate only in Granite areas, but the longevity of water in an individual Granite location would depend on fall size and the types of rock structures present. The distribution of pans and runoff pans is somewhat irrelevant, however, as Granite soils outperform these structures in all scenarios (Table 7.3). If evenly distributed, the lightest isolated falls would occur around twice per year. Soils would preserve water for a single day, gnammas for up to five days, but runoff gnammas may hold water for 11 days (Table 7.3). Falls of 25–34.9 mm occur every eighteen months, and water would then be present for slightly longer – up to six days in Granite soils, eight days for gnammas, and over two weeks in runoff gnammas. Larger falls, between 35 and 64.9 mm, occur more rarely, once every two and a half years. Again, water would only be present in Granite areas, for 6–20 days in soils and up to four weeks in rock structures. Pan 1–2 2–3 3–4 9–10 ≥ 10 Runoff gnamma 6–11 11–16 16–20 20–25 25–29 29–34 34–38 38–42 42–46 ≥ 46 Runoff pan 3–5 5–7 7–9 9–11* Soil 0–1 1–6 6–10 10–15 15–20 Soil 0 0 0 Thicket 10–12 12–14 14–17 17–19 19–21 21–24 4–5 0 5–6 11* 0 6–7 7–8 8–9 ≥ 105 mm 8–10 85–94.9 mm 35–44.9 mm 5–8 75–84.9 mm 25–34.9 mm 3–5 65–74.9 mm 15–24.9 mm Gnamma 55–64.9 mm Storage type Granite 45–54.9 mm Landscape Division 95–104.9 mm Table 7.3 Longevity of water (days) in various soils and rock structures, as a result of differently sized isolated falls. Durations are calculated assuming that falls occurred in December and were received by a dry structure or profile. Daily loss for soils is based on average December figures (4.4 mm for Granite, 4.8 mm for Thicket), and recharge amounts as per Chapter 7.2.2. Rainfall data sourced from Bureau of Meteorology, for Hyden (010568) for 1929–2017 (no data for 1947). Asterisks indicate that a structure or profile has reached its storage capacity, limiting longevity. ≥ 24 11* 11* 11* 11* 11* 20– 23* 23* 23* 23* 23* 1–3 3–5 5–7 7–9 ≥9 149 No falls of 65–74.9 mm have been recorded at Hyden and falls of ≥ 75 mm and above are very rare, occurring every 14.5 years. Water would be more prevalent around Granite areas, present for up to 23–24 days in soil and gnammas, and up to 6.5 weeks in runoff gnammas. In contrast to other isolated falls, however, Thicket soils would begin to accumulate PAW if ≥ 75 mm of rain was received. Nevertheless, this water would be short-lived, around 7–9 days for the heaviest falls. Even the largest isolated fall recorded (105 mm) was insufficient to recharge soils in Heath, Mallee, Saline and Woodland Landscape Divisions, so these areas would not accumulate water from any isolated falls unless their profiles had dried very recently. In contrast, the strong response of rock structures to isolated falls means that water would accumulate in Granite areas as a result of sub-15 mm falls. These falls are far more frequent, but the resulting water would be absent from soil profiles and pans, and short-lived in other rock structures. 7.5 CONCLUSION The availability of potable water in the study area is limited, over both space and time. Soil profiles are the only source of stored water for five of the six Landscape Divisions, but this water rarely lasts through winter, let alone beyond. Of these areas, only Thicket profiles will briefly accumulate water at other times as a result of rare isolated falls > 65 mm that occur every 14.5 years or so. In contrast, Granite areas regularly yield water outside the winter months. Winter rain is preserved for longer periods due to extra input through runoff, as well as the greater protection from evaporation that some rock structures offer. Granite soils can accumulate water after falls of ≥ 25 mm, and rock structures easily respond to even smaller isolated falls as, unlike soils, the length of the preceding dry period has no bearing on the amount of water accumulated. Nevertheless, water is generally scarce by December. Only runoff gnammas occasionally end years with a water balance large enough to persist more than a few weeks into the following year, either as a result of rare high rainfall winters – that occur around once ten years – or even rarer large isolated falls. 150 CHAPTER 8 OCCUPATION MODEL FOR THE STUDY AREA 8.1 INTRODUCTION This chapter outlines an occupation model for the study area, focussing on three main questions: whether occupation could be sustained year-round, what sort of seasonal movements would be undertaken, and if/when inter-group aggregation events could be held. These questions are addressed with reference to the distribution of food and water resources as quantified in Chapters 6 and 7, incorporating the findings from Forager Models discussed in Chapter 2. Then, the archaeological signature of the occupation model is described, focussing on the density and distribution of archaeological sites and other remains, and how this will vary over time and space – not only between different Landscape Divisions but also in different parts of the study area. Technological features are evaluated, describing some attributes that should be present in stone artefact assemblages originating from various sites and parts of the study area, drawing heavily on Kuhn's (1995) concept of technological provisioning systems. 8.2 VIABILITY OF YEAR-ROUND OCCUPATION Before seasonal routes can be evaluated, it is necessary to determine whether the study area could sustain year-round occupation under normal rainfall conditions, or if people could only remain in the area during certain seasons. Food is available throughout the year, but plant foods are less abundant in autumn and winter (Table 6.2). This situation is common elsewhere in southwestern Australia (e.g. Anderson 1984; Bird 1985), so the reduced quantity of food is unlikely to hinder occupation during those seasons. Instead, the availability of potable water was the limiting factor. The data discussed in Chapter 7 suggested that year-round occupation was impossible, as potable water could not be found in any Landscape Division between mid-January and mid-May (Figure 7.8). This probably oversimplifies water availability – particularly in gnammas – by not considering the purposeful preservation of water, the rate of human consumption, the impact of increased runoff and the specific volume of water preserved; each of these factors are evaluated below. 151 Individual gnammas are morphologically unique. Timms (2013) measured 80 gnammas in southwestern Australia, 15 of which are within the study area – on average, these measured 2.02 m2 in surface area, and 720 mm in depth. A gnamma with those dimensions would store approximately 2 litres of water for every mm of depth, assuming that it had vertical sides. These average values can be used to evaluate the specific quantity of water available in various types of gnammas throughout the year, while also considering factors influencing the longevity of water, such as the use of caps. It is well-known that Aboriginal people capped gnammas with slabs of rock or piles of branches to limit evaporation and keep the water clean (Bayly 1999, 2015; O'Connor and Prober 2010). Caps preserve the water level at the time they are applied; thereafter, human consumption is the only action reducing the water level. After winter rains ceased, gnammas that receive no runoff held on average 78 mm of water (Figure 7.5). Using the value derived above, this would represent around 156 litres of water. If caps were applied at this time, and each individual consumed 1.5 litres per day, a single gnamma would supply water for 104 people-days (10 people consuming water on a single day = 10 people days; Table 8.1). It is unlikely that capped gnammas could provide much more than a short-term or emergency source of water, possibly during longer distance logistical forays. In contrast, runoff gnammas achieve their maximum water balance in early spring, as they are able to benefit from smaller, post-winter falls. On average, they held a maximum of 273 mm of water, representing 546 litres and 364 people-days, if the structure was capped. Table 8.1 The volume of water held by various capped rock structures and the longevity of water supply (people days). People days = number of people x number of days, so ten people days may be one person consuming water for ten days, or ten people consuming water for a single day. Daily consumption was estimated at 1.5 L per person. Rock structure Capping date Volume of water (L) No. people days Gnamma End of winter 156 104 Runoff gnamma Early spring 546 364 High-input gnamma End of spring End of summer 1440 620 960 413 152 As outlined in Chapter 5.4.3, runoff gnammas were modelled fairly conservatively by assuming that they captured rain that fell directly within their margins and received the same amount again through runoff – effectively doubling the rainfall. Even though granite never sheds all the water that falls on its surface (Fernie 1930; Laing and Hauck 1997), a modelled runoff gnamma requires a total capture area just 2.5–3 times its size. This accurately represents a gnamma below a wider, pan-like depression that funnels water into the structure. For example, a gnamma 1 m in diameter would need an overlying pan of around 1.73 m in diameter to capture twice the amount of rainfall. However, considerably more runoff would be delivered into gnammas below larger pans or those located at the bases of slopes or fed by long intake channels; Bayly (1999) suggested that such channels were deliberately created by Aboriginal people for this very reason. In these cases, the main limiting factor is a gnamma’s depth, which limits its capacity to accept and store additional runoff. If a gnamma received direct rainfall plus four times that amount again (i.e. captured water from an area 8–10 times its size – a modestly sized slope or intake channel), it would easily hold 720 mm of water by late winter and retain it, with minor fluctuations in water balance, until the end of spring; thereafter, increased evaporation and limited input depletes the water balance. If capped at the end of spring, high-input gnammas would preserve 1440 litres of water, representing 960 people-days (Table 8.1). Even if these structures were left uncapped until the end of summer, they would preserve more water than a standard runoff gnamma that was capped at the optimal time (620 L compared to 546 L for runoff gnammas). Even when capped, gnammas and runoff gnammas permit comparatively short-term occupation. This could be extended if several gnammas were present at one location, or if they stored a greater volume of water per millimetre of depth. Nevertheless, year-round occupation probably relied on the use of high-input gnammas in conjunction with other sources during the rainless period. People generally used the most ephemeral water sources first, before moving to more permanent supplies (Anderson 1984; Bird 1985; Gould 1968, 1969b; Tonkinson 1978:29). It follows, then, that after exhausting surface and soil water reserves, gnammas would be used, followed by runoff gnammas. Only then would people have moved to the most reliable source – high-input gnammas – which would be the only viable source of water until the winter rains began. Particular gnammas could possibly be capped at 153 the optimal time by people foraging in the local area, rather than left uncovered (and losing water) until residential groups arrived on-site. Therefore, depending on the distribution, size, morphology and runoff characteristics of various gnammas, careful management of all water sources probably permitted year-round occupation of the study area. However, this occupation would be tethered to a subset of the Granite Landscape Division for much of the year. 8.3 SEASONAL ROUTE Seasonal routes are heavily influenced not only by the distribution of particular resources over space and time, but also their predictability. When food and water resources are predictable, seasonal routes are fairly regular, such as those identified in parts of southwestern Australia and those portions of the Western Desert where rainfall is seasonal, albeit limited (Anderson 1984; Bird 1985; Cane 1987; Veth 1987). In other parts of the Western Desert, rainfall can be very localised and highly variable both within and between years. There, regular seasonal routes were lacking, and people instead followed the rains – several years often passed before a specific area was revisited, when the rains chanced to lead them back in that direction (Gould 1969b, 1991; Tonkinson 1978:29). Resource distribution is fairly predictable in the study area, so a more regular seasonal route would be possible. This regular route is first described, defined as that which occurs under average rainfall conditions, before the impact of higher and lower winter rainfall on this annual pattern is considered. 8.3.1 Regular Seasonal Route Summer In summer, occupation was tethered to those Granite areas with several large runoff and/or high-input gnammas (Figure 8.1). It is unlikely that any single location would yield sufficient water to support sizeable groups of people, so there could be no large-scale congregation around reliable water such as occurred elsewhere in the southwest during the drier months (Anderson 1984; Bird 1985). Similarly, water would be relatively short-lived at most places with runoff gnammas, necessitating moves to high-input gnammas at some point during summer. As well as providing the most reliable water sources, Granite areas offered valuable protection from 154 bushfires – for this reason, fire-sensitive plants are often well-represented at granite outcrops (Hopper 2000; Hopper et al. 1997). Of the plant foods available during summer, gum and lerps/manna are highly ranked due to their high nutritional value combined with limited handling time, as neither requires processing before consumption (Table 5.5). Lerps/manna would be abundant in Eucalypt-dominated Mallee and Woodland areas, while gum would be prolific in Thicket due to the frequency of Acacia (Tables 6.2 and 8.2). Heath would also provide gum, though in more limited amounts than Thicket areas. Figure 8.1 Occupation model for the study area under average rainfall conditions, showing the primary water sources (blue), primary foraging (green) and residential areas (yellow) and the optimal portion of the study area (orange) that should be occupied at particular times of the year. For division location see Figure 8.2. 155 Figure 8.2 Location of the northern, central, southern parts of the study area and the optimal months to occupy each. Note that these divisions are not firm boundaries, but merely indicative of the broad areas referred to in the text and Figure 8.1. 156 Table 8.2 Temporal distribution of ranked plant food items (following Table 5.5), and unranked animal foods in each Landscape Division. WOODLAND THICKET SALINE MALLEE HEATH GRANITE Summer Autumn Winter Spring Dec Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov High 30 30 30 0 0 0 0 0 0 0 0 0 Moderate 100 80 75 60 20 21 30 36 55 83 83 112 Low 47 46 28 28 28 29 29 29 44 50 50 50 Unranked 103 79 74 76 77 76 80 92 117 135 131 116 High 67 67 67 0 0 0 0 0 0 0 0 0 Moderate 147 98 87 44 33 36 58 69 109 179 179 163 Low 67 67 39 40 40 41 41 41 66 69 69 69 Unranked 117 87 81 82 84 83 87 100 126 149 147 129 High 92 92 92 0 0 0 0 0 0 0 0 0 Moderate 143 105 96 31 26 27 49 57 90 174 173 156 Low 76 76 35 36 36 37 37 37 75 78 78 78 Unranked 136 103 96 97 98 100 105 122 157 183 178 154 High 25 25 25 0 0 0 0 0 0 0 0 0 Moderate 71 62 54 14 7 9 13 17 29 68 68 72 Low 41 40 33 33 33 34 35 35 43 43 43 42 Unranked 61 48 45 45 47 48 57 70 100 101 95 79 High 46 46 46 0 0 0 0 0 0 0 0 0 Moderate 104 72 68 31 22 25 41 51 73 124 124 114 Low 40 40 22 23 23 23 23 23 39 41 41 41 Unranked 82 64 60 62 63 64 67 80 104 118 113 91 High 63 63 63 0 0 0 0 0 0 0 0 0 Moderate 98 78 74 29 14 15 24 32 51 101 101 106 Low 51 51 25 36 26 27 27 27 52 54 54 54 Unranked 146 110 103 97 104 106 110 125 163 192 188 166 157 Following Macarthur and Pianka’s (1966) diet-breadth model, it is safe to assume that people would want access to both high-ranked plant food categories. However, other plant and animal foods would also be required since residential mobility may have been impeded by the infrequency of water sources. The most optimal summer camp location would be in Granite areas in the northern portion of the study area (Figure 8.2) – gum would be abundant in the extensive Thicket formations and lerps/manna in the intervening Mallee, which also provides a good range of lower ranked resources. Heath areas would still be accessible, and pockets of Woodland are common in the northern Mallee formations (Figure 3.14; Table 8.2). Mallee and Woodland areas both yield an abundance of fauna (Table 6.3), including reptiles that could be easily caught during summer. Less fauna is present in Granite, Heath and Thicket areas and larger game would be especially rare in dense vegetation (MacIntyre et al. 1992). Granite areas also provide a wealth of moderately ranked storage organs, including a considerable number that are only available for five months centred around summer – importantly, many have high water-content at this time (Table 6.2; Appendix B.2). As occupation span increased, higher ranked resources would become depleted, so lower ranked items would be pursued (Macarthur & Pianka 1966). Acacia seeds would be an ideal choice due to their high calorie, fat and protein content but, like most seeds, they require extensive processing so are ranked poorly compared to other plant foods (Brand-Miller and Holt 1998; O'Connell and Hawkes 1981; Table 5.5). Again, the northern portion of the study area provides the optimal solution, because Acacia seeds are abundant in Thicket areas – this resource could not be intensively exploited from further afield, since low-ranked items can only be profitably transported short distances (Orians and Pearson 1979). People may have stored some of these Acacia seeds and gum-cakes for autumn, when plant foods are scarcest, but this would only be profitable around high-input gnammas, where occupation may continue beyond summer. The northern area therefore provides a diverse diet through Mallee and Heath Divisions, while permitting easy access to numerous Acacia-derived resources found in Thickets, some of which would be especially significant as the stays lengthened; high-ranked gum and lerps/manna would also be abundant. In contrast, the southern 158 part of the study area would be suboptimal for human occupation at this time. Woodland and Mallee areas provide various types of game as well as lerps/manna, but Heath areas are smaller and Thicket is inaccessible, restricting the abundance and diversity of high-ranked resources and limiting the exploitability of low-ranked seeds that cannot be profitably transported. Autumn In autumn, the availability of potable water is still restricted to certain Granite areas – runoff gnammas were likely exhausted except where very large structures may have formed; occupation was most likely centred around high-input gnammas (Figure 8.1). High-ranked plant foods were unavailable after summer, heightening the importance of other resources. Moderate-ranked resources (fruit, flowers and storage organs) are most diverse in Heath, Mallee and Thicket areas, but flowers dominated Thickets (Tables 6.2 and 8.2). Granite areas provide an abundance of plants with storage organs, but only a fraction would be available after March. Low-ranked resources are readily available in all of these Landscape Divisions, and Thicket would again provide an abundance of seeds, now derived from Allocasuarina – another of the three dominant genera – that preserve their seeds year-round inside woody fruits. Animals would be a more important food source during autumn and the most diverse are found in Woodland and Mallee areas. Heath and Granite also provide reasonable faunal assemblages, and the latter supports three amphibian species that may contain eggs in autumn. Plant foods are scarce in autumn and high-ranked items are entirely absent. It would therefore be vital to have access to the greatest overall diversity of resources. The southern portion of the study area is again suboptimal. Thicket areas are inaccessible and Heath pockets are small and widely spaced. Saline areas are accessible, but yield the fewest moderate-ranked resources (although storage organs are relatively diverse), and fauna are scarce. The optimal solution can instead be found in the central study area (Figure 8.2), where Heath and Thicket meet, specifically at Granite areas surrounded by Heath and Mallee. Patches of Woodland occur within Mallee, and Thicket areas are also easily accessible (Figure 3.14). Depending on the distribution and capacity of high-input gnammas, it is possible that certain occupation sites could be used throughout summer and into 159 autumn. While the residential location may be fixed, there would likely be a shift in the foraging range, with more intensive use of Heath areas instead of Thicket (Figure 8.1). Longer stays would deplete resources, possibly necessitating an increased foraging range, likely incorporating long logistical forays to target plant or animal foods that could be profitably transported; capped gnammas or well-timed isolated falls may provide valuable water sources for such journeys. From mid-May, flexibility was increased somewhat – potable water was still restricted to Granite areas but may be intermittently present at outcrops with runoff pans and any gnammas. These water sources could be used as ‘stepping stones’ to allow people to begin moving away from summer and autumn camps towards winter foraging areas in preparation for rain. Winter At the beginning of winter, rain is generally infrequent and light, so reliable water is restricted to Granite areas (Figure 8.1). From late June, however, potable water can be found in all Landscape Divisions. This newly available water opens up parts of the landscape that were previously inaccessible, either through the lack of water or being too far from summer and autumn camps. Moderate-ranked plant foods are most diverse at Heath, Mallee and, to a lesser extent, Thicket areas (Table 8.2). Heath and Mallee have similar assemblages of fruits and storage organs, but flowers are more diverse in Heath; all supply a variety of low-ranked items. In contrast, Thickets yield few storage organs, fruits or low-ranked resources, but flowers are diverse. Mallee and Woodland again yield the most diverse faunal assemblages, but winter rains fill the ephemeral salt lakes, prompting the arrival of migratory waterbirds in Saline areas – a great number of these and other birds lay eggs from July. Their eggs are most diverse in Woodland, Mallee and Saline areas, and waterbirds are unsurprisingly restricted to salt lakes within the latter Landscape Division. Macropods, and potentially other herbivores, are at peak condition in winter and concentrated in the more open Woodland and Mallee areas (Appendix C). Reptiles would be difficult to catch after retreating to their winter burrows. Aside from Mallee, in winter there is a disconnect between the Landscape Divisions yielding the most diverse plant foods and those providing more abundant animal foods. The optimal location could therefore be the central study area where Mallee 160 (with some pockets of Woodland) is interspersed with Thicket and Heath, or the southern study area, where large expanses of Mallee occur along with Woodland, Saline and some Heath areas (Figure 8.2). The southern palaeovalley may have been preferable for several reasons – first, it provides access to comparatively large expanses of four of the five Landscape Divisions yielding diverse plant or animal assemblages, while the central area has only three. Secondly, the central area was occupied through autumn so would likely be suffering from some level of resource depletion and would benefit from people moving away to allow resources to regenerate. Finally, Aboriginal people are well-known for exploiting seasonal bounties, and characterise their seasons not only by weather conditions but by the plant and animal foods that become available at various times, and the activities associated with procuring them (e.g. Clarke 2009, 2011:54; Collard and Harben 2010; O'Conner and Prober 2010). The abundance of birds and their eggs, available primarily in Saline, Mallee and Woodland areas, would invariably be targeted. Similarly, while most moderate-ranked resources are rare in Saline areas, this Landscape Division yields the most diverse assemblage of plants with perennial storage organs, after Granite (Table 6.2). In winter, then, people would likely have spread out into the newly accessible southern part of the study area – particularly around the palaeodrainage network – making use of ephemeral surface and soil water. Occupation may have been more intensive in the eastern area, where larger salt lakes may attract a greater number of waterbirds. Water is only intermittently available in Mallee soils (Figures 7.2 and 7.8), so occupation sites should be primarily in Saline or Woodland areas, including where Woodland occurs within the Mallee Landscape Division. Granite areas may be too wet for comfortable habitation and Heath areas too dense – while access could be created via burning, heath vegetation was infrequently fired elsewhere in southwestern Australia (Prober et al. 2016). Considering the ephemerality of most water sources, residential mobility would be greater during winter. Frequent moves would also allow access to new resource patches, which may reduce reliance on low-ranked items such as seeds, which require considerable processing. Longer logistical forays should be unnecessary, since residential mobility is unimpeded and water is universally available, albeit in somewhat limited supply. The reduced resource base and ephemeral water would favour small groups, probably smaller 161 than those occupying the most reliable water sources in summer and autumn, so some level of social fragmentation likely occurred, as it did elsewhere in the southwest (Anderson 1984; Bird 1985). Spring From mid- to late winter, and certainly by spring, the choice of residential locations narrowed, as water is only available in the Granite Landscape Division (Figure 7.8). Plant and animal foods are mostly distributed as during autumn, but are becoming more abundant; reptiles are active and easily caught, and many are also laying eggs (Table 6.3). Plant foods are becoming somewhat more abundant, and moderateranked resources are again concentrated in Heath and Mallee areas (Table 8.2). While Saline areas still provide waterbirds and some eggs, Woodland and Mallee areas offer considerably more bird eggs, combined with far more diverse reptile assemblages. Other animal foods are also well-distributed in Woodland and Mallee areas, but the condition of large herbivores begins to deteriorate from October. Occupation was likely concentrated in the southern part of the study area, as in winter, but residential sites would have shifted towards Granite areas, where soil water was available. These reserves would be exhausted by mid- to late October, so groups would be forced to relocate to gnammas (Figure 8.1). Standard gnammas would only support short-term occupation, unless several very large structures were grouped together, but the presence of runoff or high-input gnammas would permit considerably longer stays. By November, runoff gnammas and high-input gnammas would be the main sources of water (Figures 7.8 and 8.1). At this time, most migratory waterbirds would have left, and ephemeral salt lakes would be dry; some waterbirds would still be laying, probably confined to the larger, permanent salt lakes in the eastern portion of the study area. Nevertheless, Saline areas would yield fewer eggs than most other Landscape Divisions in spring (Table 6.3). There would be fewer benefits associated with Saline valleys, so groups possibly began moving northwards towards the central study area, where extensive Mallee and Heath formations occur with pockets of Woodland. Heath and Mallee areas provide a diversity of moderate-ranked food resources, with Woodland areas yielding a similar diversity of fruit and storage organs but fewer flowers. An abundance of storage organs would be newly available 162 in Granite areas, specifically those that refilled their subsurface stores over the growing season in preparation for their summer dormancy. The central study area would have had nearly five months to recover from autumn occupation, so resources should again be fairly abundant. This shift placed groups in an ideal location before summer, when the northern part of the study area again represented the optimal settlement location. 8.3.2 Impact of Rainfall Variation While the quantity and distribution of rainfall influences the quality and abundance of plant and animal foods, this impact is difficult to measure and negligible compared to its effect on water availability. The latter should manifest itself primarily in the length of occupation at particular sites. Under average rainfall conditions, water is almost universally available in the study area for around two months, allowing maximum flexibility in settlement patterns; at all other times, occupation is tethered to the Granite Landscape Division (Figure 8.1). Under low rainfall conditions, however, there is perhaps one month during which people could spread out and utilise ephemeral water sources (Figure 7.8). Soil water, gnammas and runoff gnammas all hold water for shorter periods than under average rainfall conditions, so those areas could support correspondingly shorter occupation periods – only high-input gnammas would permit extended stays. Occupation would be centred around these both in the southern palaeovalley before the northward move in November, and in the central and northern study area from late spring through autumn. This would be especially problematic in the north and central study area as occupation is tethered to these sites for long periods even under average rainfall conditions. Any extension would heighten resource depletion, requiring more intensive use of lower ranked plant resources as well as longer logistical forays to target more distant resources. It is difficult to determine whether water supplies in these areas would last through to the following winter without site-specific data on the distribution and capacity of highinput gnammas. Movement into adjacent areas may be possible, but this would impose additional strain on their resource base, especially as the hydrological situation is unlikely to be improved nearby. Instead, it may have been necessary for some groups to remain in the palaeovalley in summer and autumn, around high- 163 input gnammas, and to forage more widely to offset the reduced diversity of plant foods and the effects of resource depletion from the winter and spring occupation. In contrast, high rainfall conditions offer increased flexibility as soil water supplies allow occupation of non-Granite areas for three months of the year. In contrast to average and low rainfall years, Mallee areas accumulate soil water fairly consistently when rains are strong, so people would be free to camp in the extensive Mallee formations at this time. From early spring, occupation would again be tethered to Granite areas, but all types of gnammas would permit longer occupation than under average rainfall conditions. However, these supplies may not have been heavily utilised, as Granite soils provided water until the end of October (Figure 7.8), and these permitted greater freedom of movement around the area. The increased water availability would not necessarily translate to longer occupation of the winter foraging grounds in the southern palaeovalley, but this may be optimal if higher rainfall encouraged more waterbirds to travel to the study area and to remain, and breed, for a longer period. In this case, gnammas would provide a viable water source after Granite soils had dried, until the northward movement had commenced. Once occupation shifted to the central and northern study area, runoff gnammas would provide water for considerably longer than under average rainfall conditions – occupation would therefore be tethered to high-input gnammas for a correspondingly shorter period. Importantly, this would reduce resource strain around them, permitting fewer logistical forays and less intensive use of low-ranked resources. 8.4 INTER-GROUP AGGREGATION The spatio-temporal availability of resources dictates regular seasonal movements but also influences the timing and location of inter-group aggregations for social, ritual or ceremonial purposes. As Gould (1969a:121) noted 'Aboriginal ceremonial life is not cut off from the practicalities of subsistence and daily living.' Aggregation patterns depend on local ecological conditions and people may aggregate due to resource abundance or, conversely, scarcity, such as when food is abundant but reliable water sources are rare (Conkey 1980; Stanner 1965). Some argued that food storage was an important component of larger aggregations (Hallam 1975:162; Gammage 2011:296; Pascoe 2013:106–107), while others claim that events were 164 instead timed to coincide with natural periods of abundance (Brown 1918; Clarke 2009; Gould 1969b, 1984, 1991; Hassell 1975). In either case, considerable quantities of food and water would be required. As noted in Chapter 4.2.2, Noongar people gathered into larger residential groups during summer, when food was abundant, in areas where reliable water was available, often on the coast (Anderson 1984; Bird 1985). While neither Anderson (1984) nor Bird (1985) specifically mention social or ceremonial gatherings, it is likely that if they occurred within their respective study areas, their timing would coincide with these regular periods of intra-group aggregation; large groups could not be supported at other times. Similarly, O'Connor and Prober (2010) noted that in the warmer months, abundant food and reliable water meant that permanent coastal residents of the Great Western Woodlands (east of the Wheatbelt) could be joined by people from further inland, so meetings and social gatherings could be held. Similar patterns existed in parts of the Western Desert where rainfall is predictable – people dispersed during the rainy period (in this case, summer) when food was less abundant, and congregated around more permanent water in the cooler, drier months when abundant food allowed ceremonies and meetings to occur (Cane 1987; Veth 1987). Elsewhere in the Western Desert, rainfall is much less predictable and seasonal rounds were less well-defined (Gould 1968, 1969b, 1991; Tonkinson 1978:29). Food could not be stockpiled for ceremonial events, as storage is only beneficial where there is a guarantee of returning to a particular location in the foreseeable future (Gould 1969b, 1991). Therefore, aggregations could only occur during the periods of natural abundance that occur after good rains, when plant foods, water and game were all simultaneously plentiful – consequently, these aggregations were often many years apart (Gould 1969b, 1984, 1991). Aggregations require either a synchronous abundance of food and water or, more commonly, abundant food combined with a water source capable of sustaining a large group of people for a short time. It is unlikely that food could be stored in sufficient quantity to permit winter gatherings, when water was more plentiful, so aggregation events in the study area would need to coincide with periods when food was naturally abundant. Summer would be ideal, due to the availability of highranked plant foods (gum and lerps/manna) and December may have been 165 particularly suitable as good quantities of bird and reptile eggs were still available (Table 6.3). However, the asynchronous distribution of critical resources means that this summer food abundance was accompanied by limited water availability. While larger isolated falls (≥ 35 mm) provide water that will withstand evaporation for a week or more (Table 7.3), it is unlikely these sources would support the rate of consumption associated with aggregations. The Hyden meteorological station has only recorded three ³ 35 mm falls in December since record keeping began in 1928. Therefore, aggregations could only occur by relying on winter rainfall stored in gnammas. Under average rainfall conditions, runoff or high-input gnammas would preserve sufficient water if capped at their peak; following high winter rainfall, all gnammas would probably be capable of supporting aggregations, especially if more than one was present at a particular site. Aggregating after high winter rainfall may be advantageous for several reasons: it places fewer restrictions on the location of aggregation sites, so they can be selected qualities other than the presence of a specific type of gnamma; water will be more widely available following heavy winter rains, which may facilitate travel to and from the study area; the increased longevity of stored water supplies mean that reserving a source for aggregation would not impact regular occupation by decreasing the amount of water available for consumption at other times; and plant and animal foods may be more abundant or of higher quality after good winter rainfall (e.g. macropods may stay at peak condition through October and beyond). High winter rainfall occurs roughly once per decade (BoM data; Chapter 7.4) so aggregation events may have been infrequent, and held many years apart, as they were in the Western Desert (Gould 1969b, 1984, 1991). It is also important to consider the parts of the study area that would best accommodate aggregations, since large groups could significantly deplete the resource base around a particular site, placing future occupation at risk. It seems reasonable to assume that aggregations were held in areas that were otherwise unused in summer, such as the southern palaeovalley (Figure 8.2). The central study area may have been within the summer foraging range, so would be less ideal. In the south, high-ranked lerps/manna would be abundant in the Mallee and Woodland areas and gum could be procured, albeit in smaller quantities, from most accessible Landscape Divisions. Moderate- and low-ranked plant resources, fauna and eggs 166 are fairly prevalent in Woodland and Mallee areas. Similarly, salt lakes may still be holding water after strong winter rainfall, so waterbirds may still be in the area. Granite provides an abundance of moderate-ranked storage organs throughout summer. If required, low-ranked resources such as Acacia seeds could be utilised. While these are most abundant in Thicket areas, they can be found in smaller quantities in all Landscape Divisions, so their proximity to the aggregation site should mean that transport costs are not prohibitive. Similarly, the timing and location of an aggregation is highly predictable as it is dictated by the quantity of winter rainfall and the presence of suitable rock structures and other site characteristics. Therefore, it may have been possible to stockpile foods at the site in advance during the regular palaeovalley occupation period. Gathering low-ranked items, such as Acacia seeds, could be embedded in regular foraging, reducing costs in the same manner as raw material procurement. 8.5 ARCHAEOLOGICAL AND TECHNOLOGICAL SIGNATURES As demonstrated above, residential mobility varies considerably when the north and central part of the study area are compared to the southern palaeovalley; archaeological and technological signatures should vary accordingly. While quartz is readily available at some Granite areas (see Chapter 3.2) it is not as ubiquitous so the widespread adoption of an informal technology could probably not be supported. As a result, evidence of Individual and Place Provisioning should vary according to the degree of residential mobility. Furthermore, temporal change may be identified in the archaeological record, since the length of occupation at particular sites is linked to the quantity of winter rainfall. While the previous discussion centred on individual years where winter rainfall is higher or lower than average, it is unlikely that such short-term shifts would be identifiable in the archaeological record. Therefore, changes will only be reflected across a larger time-scale, but the same patterns should be in place. For example, the archaeological signature expected during a high rainfall year would hold true over a longer period of increased rainfall. Analysis of changing occupation patterns may thereby provide valuable palaeoenvironmental data by indicating shifts in localised water availability as a proxy for effective precipitation (Rossi 2014b). 167 8.5.1 Northern and Central Study Area It was argued above that occupation was concentrated in the central and northern study area from late spring through autumn, but the scarcity of water limited residential mobility, tethering groups to runoff or high-input gnammas (Figure 8.1). This tethering meant that the timing and location of occupation was highly predictable, favouring a Place Provisioning strategy dominated by more expedient stone reduction technologies with limited recycling, retouch and maintenance of items (Kuhn 1995). Raw material should be predominantly quartz from the local area, as embedded procurement can occur during foraging, aided by the widespread distribution of the material. Longer term occupations may encourage some raw material conservation – this will be reflected by bipolar flaking techniques, which conserve stone while maintaining the benefits of a less formal technology (Hiscock 1996; Parry and Kelly 1987; Chapter 2.3.1). However, long occupations create a higher risk of resource depletion, so grindstones and/or grinding patches may indicate more intensive use of seeds – a low-ranked resource. Similarly, longer logistical forays were likely required to target distant food resources that could be profitably transported. In these cases, there should be some level of 'gearing up', a specialised form of Individual Provisioning (Binford 1979; Kuhn 1995). There may be some evidence, then, of higher quality raw materials along with retouch and more controlled flaking, but this debris will be numerically overwhelmed by evidence of Place Provisioning (Kuhn 1995:29). Higher quality raw materials would probably comprise finer textured quartz with fewer internal fracture planes, since few other types are available within the study area. Dolerite is heavy and coarse-grained, so more suited to ground artefacts (Glover 1984) and less likely to feature within a portable toolkit. Residential sites in the central and northern study area should have strong archaeological visibility for two reasons. First, a greater quantity of archaeological debris is generated by Place Provisioning when compared to Individual Provisioning strategies (Kuhn 1995:29). Furthermore, occupation is tethered to small areas near reliable water, as people rarely situate themselves more than 100m from water during drier months – the subsequent reoccupation of the same location results in the continued accumulation of archaeological material (Binford 1980; Gould 1968). As these sites will be palimpsests of several occupation events, they may preserve 168 evidence of changing occupation intensity over time, representing shifts in the quantity of water that could be procured. Under low rainfall conditions, runoff gnammas can support shorter occupations. In contrast, sites at high-input gnammas should demonstrate correspondingly longer stays, as no other viable water sources exist in the study area during the relevant seasons. The reverse is true under high rainfall conditions, where increased water availability at runoff gnammas permits shorter occupations at high-input gnammas. These shifts may be expressed as variations in the amount of debris deposited (discard rates), length of reduction sequences, and the incidence of bipolar reduction (Hiscock 1996). Longer occupations at high-input gnammas may also produce more evidence of ‘gearing up’, due to the increased importance of logistical forays. While residential sites should be located primarily around reliable water sources, there may be evidence of extractive sites or very short-term occupation elsewhere, reflecting logistical forays to procure higher ranked food resources. These sites will be more difficult to detect archaeologically, as the unpredictability of visitation would discourage Place Provisioning and limit the formation of stratified archaeological deposits. However, if local stone (most likely quartz) was available nearby, expedient tools may have been produced and discarded on-site. Bipolar flaking should be absent, due to short occupation phases, but there may be some evidence of toolkit maintenance, in the form of debris comprising high-quality raw materials. 8.5.2 Southern Study Area (Palaeovalley) The southern palaeovalley was the focus of occupation throughout winter and most of spring (Figure 8.1). During winter, water is more widely available, permitting a degree of flexibility in occupation patterns impossible at other times of the year. While the optimal Landscape Divisions can be predicted – primarily Saline and Woodland areas (including those occurring as pockets within Mallee) – there was little need to reoccupy the same precise location on subsequent occasions. Residential sites in Mallee areas lacking pockets of Woodland can only be the result of high winter rainfall, as intermittent soil water means they cannot be occupied at other times (see Chapter 8.3.2). Therefore, Individual Provisioning should have been be the dominant strategy, but if raw material was available nearby there could be some level of expedient artefact manufacture. Residential moves would be fairly 169 frequent due to the ephemerality of water resources, as well as the reduced suite of plant foods, so occupation debris will be scarcer than in the northern and central study area and should demonstrate little or no evidence of bipolar flaking. The reduced chance of reoccupation will also limit the accumulation of archaeological material in specific locations. After winter rains cease, only Granite areas can support occupation, primarily due to the presence of soil water. Nevertheless, precise residential locations are largely unpredictable, since people would be free to choose from any one of the numerous Granite zones occurring in the area. Similarly, campsites could be selected from innumerable locations at any Granite outcrop, since water is fairly ubiquitous. Hence, Individual Provisioning should dominate, likely supplemented by the local quartz that is accessible at many Granite outcrops. Archaeological visibility will be lower than in the northern and central study area, but higher than in Woodland and Saline areas, as Granite acts as a tether, albeit weaker than elsewhere in the study area. At gnammas, the occupation location is predictable, but duration is dictated by the number and size of rock structures present, the amount of runoff received, and the availability of water elsewhere in the area. Gnammas and runoff gnammas should be characterised by Individual Provisioning, as visits will invariably be short, either due to limited water storage (low winter rainfall years), the timing of regular seasonal movements (average winter rainfall) or greater freedom in selecting residential sites (high winter rainfall – see Chapter 8.3). High-input gnammas permit longer and more predictable visits, even under low rainfall conditions, so Place Provisioning may dominate at these locations. Further, temporal change should be evident – low rainfall should be accompanied by increased occupation intensity, while they would be occupied less intensively under high rainfall conditions. These shifts should be primarily reflected in artefact discard rates and the length of reduction sequences. Aggregation sites should only be found in the southern palaeovalley area. The archaeological signature at these sites will differ from the regular occupation sites described above. Location should be semi-predictable: a Granite area with several large gnammas, ideally capturing some degree of runoff, that can be capped. Aggregation events may require more intensive use of low-ranked resources, so there may be some evidence of seed processing through grindstones and/or grinding 170 patches. While Place Provisioning is generally not beneficial for shorter term occupation, the intensity of occupation – in this case dictated by group size – may warrant stockpiling some raw material, but the short occupation span would discourage the use of bipolar techniques. This Place Provisioning should be accompanied by Individual Provisioning, reflecting groups that have travelled longer distances to attend. Exotic materials may be present, though this depends on the types of stone available in the visitors’ home territory. Conkey (1980) and Mitchell (2016) noted that aggregation sites often demonstrate a level of segregation, with different parts of the site used for specific activities. This may create distinct assemblages in each area, or at least distinct clusters of archaeological material. While aggregations were probably infrequent, their archaeological visibility should be reasonable due to large group sizes, potential spatial organisation and some use of Place Provisioning. Rainfall patterns dictate the frequency with which these events can be held, so discard rates may be linked to longer term precipitation records. For example, higher discard rates would indicate more frequent events, which could only occur when high rainfall years were common, presuming that group size was relatively consistent over time and that events could not be sustained under average rainfall conditions. 8.6 CONCLUSION Despite the lack of freshwater lakes and rivers, or potable groundwater, the study area could be occupied year-round. However, for much of the year groups would be restricted to a subset of the Granite Landscape Division, where high-input gnammas capture runoff from larger areas. From late spring to early autumn groups would occupy the central and northern part of the study area, with residential sites tethered to the few reliable water sources found therein. For the remainder of the year, people would reside in smaller groups in the southern palaeovalley, frequently moving their residential bases between different parts of the Saline and Woodland Landscape Divisions to make use of ephemeral soil water resources. Following the cessation of winter rain, occupation would be tethered to Granite. Inter-group aggregations could be held about once a decade in the southern palaeovalley in summers after heavy winter rain. The archaeological and technological signatures should also exhibit a pronounced north-south dichotomy. In the central and northern study area, 171 occupation was lengthy and predictable, favouring a Place Provisioning strategy. In the southern palaeovalley, residential mobility was increased, and occupation location was far less predictable, so Individual Provisioning should dominate, although some Place Provisioning may occur at aggregation sites and high-input gnammas. Throughout the study area, occupation intensity at high-input gnammas should demonstrate an inverse relationship with rainfall, reflecting tethering when water availability is limited, and greater freedom under higher rainfall conditions. 172 CHAPTER 9 ARCHAEOLOGICAL METHODS 9.1 INTRODUCTION This chapter outlines the methodology applied during collection and analysis of archaeological material. First, field methods are described, beginning with the siteselection process before considering the techniques used during archaeological excavation and the collection of surface artefacts. Lithic analysis procedures are then detailed, with particular focus on the problematic nature of quartz, before describing the classification system adopted and the attributes measured. The analysis of other collected materials – charcoal, sediment samples, ochre and glass – is then outlined. Finally, the data analysis methods are described, focussing on quantifying site-specific water sources and using artefact data to identify technological provisioning systems and varying levels of occupation intensity. 9.2 FIELD METHODS This section outlines the main field methods employed throughout this research. Derivations from these general methods, as well as more specific details (i.e. the location of individual test-pits), can be found in Chapters 10–12. 9.2.1 Site Selection As noted earlier (see Chapter 1), Mulka's Cave was the initial focus of this research because the extensive suite of rock art made it a potentially significant site, but the author’s previous interpretations were hampered by a lack of archaeological context (Rossi 2010, 2014a). To address this problem, an entirely new lithic analysis was required, combined with comparative data from the surrounding area. The Department of Planning, Lands and Heritage’s online database was consulted to identify registered sites and 'other heritage places' within 50 km of Mulka's Cave. The relevant reports and site files were then evaluated to determine if any of these locations had potential for undisturbed sub-surface archaeological content. Locations were discounted if on private property (for reasons of access), if they preserved limited or no archaeological content (i.e. lost through development, or if literature 173 noted a lack of artefacts), and where there was evidence of disturbance or erosion. Only a few open artefact scatters (ID 16787, 16791, 18508 – see Table 4.1) as well as two water sources (5120 Anderson Rocks, 5121 Anderson Rocks 2 aka Twine Reserve) met these criteria. In the case of 5120 and 5121, the presence/absence of artefacts could not be confirmed as there was no associated literature. Anderson Rocks and Twine Reserve (Figure 9.1) were prioritised because if these sites had been occupied, the tethering effect of water sources would have encouraged occupation within a more restricted area, thereby increasing the chance of encountering stratified archaeological deposits. Both sites were visited, and each exhibited a series of gnammas as well as obvious (Twine Reserve) or potential artefacts (Anderson Rocks), warranting small-scale excavation of both sites. Figure 9.1 Location of archaeological sites excavated by the author. Unfortunately, investigation at Twine Reserve had to be abandoned (see below). To offset this loss of data a more intensive investigation was planned at Anderson Rocks. Nevertheless, another site was still required to maximise comparative data. It was cost-prohibitive to consider another location that met Section 5 of the Aboriginal Heritage Act (1972), so the decision was made to search other areas for potential 174 sites. The prevalence of agriculture creates a distinct preservation bias in the study area, so the search was limited to areas of state-owned land – primarily large granite outcrops or salt lakes, and a few small wooded reserves. Several areas yielded no cultural material, or even potential cultural material, with the exception of Gibb Rock, a granite outcrop that lacked gnammas but preserved potentially flaked stone. Gibb Rock provided a mix of similarities and contrasts to the other two sites: all are in the Granite Landscape Division but have vastly different water sources. The three sites could then be used to address the newly emerged question regarding the use of occupation history at certain areas as a proxy for effective rainfall (Rossi 2014b). Twine Reserve (aka Anderson Rocks 2) It is necessary to include a brief discussion of the Twine Reserve investigation as it initially formed part of the site sample and some fieldwork was undertaken. Permission to test-pit was sought, and subsequently granted, under Section 16 of the Aboriginal Heritage Act (1972). As a condition of their support, the Ballardong people (Traditional Owners for inland southwestern Australia – see Figure 4.3) required three monitors to be on-site during the excavations, as well as the implementation of an Intellectual Property (IP) agreement before reporting of any results deriving from the excavations. The monitoring process required each Traditional Owner (TO) to be paid $500 per day for their time, in addition to travel allowances plus food and accommodation costs. On-site monitoring was not required during earlier research at Mulka's Cave so, perhaps naively, the author assumed that this would not be required for the current research, since all sites are administered by Ballardong. Monitoring costs had therefore not been allowed for during the planning and budgeting stage for this research, which was primarily selffunded. It should be noted that the amounts requested by the Ballardong are in-line with those specified by the Noongar Standard Heritage Agreements that were implemented in June 2015, as a part of the yet to be finalised South West Native Title Settlement. Available funds could only support a three-day field trip to Twine Reserve, with the designated Indigenous monitors. While such a short visit would permit a cursory investigation, at best, the archaeological content was deemed significant enough to warrant the expenditure. The trip was scheduled for November 2015 with the aim to 175 excavate one or two small (500 x 500 mm) test-pits, as shallow deposits were anticipated; unfortunately, various delays meant that little was actually achieved during the limited timeframe. The first pit was very time-consuming to excavate due to several pockets of heavily cemented soil (likely due to an ants’ nest encountered lower in the pit) and was eventually abandoned due to limited archaeological content. Before another pit could be initiated one of the monitors fell ill, and it was clear that it was in her best interests to return to Perth immediately – therefore, excavation could not continue under the conditions set down by the Ballardong people. Due to financial constraints, further excavations at Twine Reserve were impossible. For this reason, as well as the limited archaeological content within the single excavated test-pit, no attempt was made to develop the IP agreement; instead, the results of this excavation will not be discussed further. However, this site features in the broader conclusions made herein, since the data did not derive from the excavation, but rather from the author’s observations made at a time outside of the November 2015 visit. 9.2.2 Survey and Excavation Methods Before any fieldwork was undertaken on registered Aboriginal sites, permission was sought under Section 16 of the Aboriginal Heritage Act (1972). While this specifically relates to excavation, in practice it also applies to the removal of cultural objects from the ground surface (i.e. surface artefacts should not be collected without obtaining permission). No excavation or removal of cultural material occurred at a registered site before the issuing of a Section 16 permit – these permits generally allow the excavation of one square metre across a site. These same guidelines were adhered at other locations, even where permissions were not required, as there was a strong chance these would be registered as sites in the future, based on the results of this research. All sites were evaluated on foot, but systematic grid searches were impossible due to large areas and limited access. Instead, investigation focussed on accessible areas where archaeological content was likely, i.e. at the base of outcrops, areas where sediment had accumulated on the rock, and around any water sources. Noteworthy features (e.g. gnammas) were positioned via GPS and photographed; gnammas were also measured, including their intake areas, where present. Surface artefacts 176 or potential artefacts were treated differently depending on whether these occurred in identifiable concentrations (surface artefact scatters) or were more sparsely distributed across the site. GPS coordinates were used to define the extent of artefact concentrations; small and/or lower density scatters were often collected as a single unit, while larger and/or denser scatters were subdivided depending on the distribution of material (see Chapters 10–12). Elsewhere, individual artefacts/potential artefacts were positioned via a GPS reading and collected along with any other potential cultural material visible within a certain radius, generally less than three metres. These methods permitted a reasonable level of spatial control within the available timeframes. At all sites, 500 x 500 mm pits were excavated. This permitted greater intra-site spatial analysis while remaining within the one square metre limit. Pits were positioned in areas offering the best chance of deep, undisturbed and stratified archaeological content, generally in areas where surface artefacts (or potential artefacts) were densest and there were no obvious signs of disturbance such as drainage channels or animal burrows. Pit locations were recorded via GPS, and the pit aligned on cardinal compass points where possible. Excavation was conducted by the author and her secondary supervisor, R.E. Webb. Due to the lack of visible stratigraphy the deposit was removed in arbitrary 25 mm excavation units (XUs), but XU depth often varied according to the consistency of the sediment, excavator error, or the presence of intrusive elements. Sediment samples were only taken from selected pits and XUs, as no soil changes were apparent. Excavated sediment was weighed, sieved through nested 6 mm and 3 mm mesh, and the retained material bagged. This matter was subsequently sorted, separating natural debris from charcoal, potential artefactual stone and other items of interest (e.g. glass, ochre). Natural debris was retained but not studied further, while the remaining material was subject to the additional analyses outlined below. 9.3 LITHIC ANALYSIS Before outlining the lithic analysis methods, it is necessary to evaluate the dominant raw material in study area – quartz. The problems associated with this material, including difficulties identifying intentional flaking, heavily influence subsequent 177 methodology in terms of both the classification system adopted and the specific attributes selected for analysis. 9.3.1 ‘Horrid Little Bits of Stone’ – the Problem with Quartz Artefact Analysis Quartz (SiO2) is a mineral and has a crystalline structure; the size of crystals and the way in which they aggregate can vary widely. Macrocrystalline quartz, the type to which analysts refer when lamenting the difficulty of quartz artefact analysis, comprises those varieties with a crystal structure that is identifiable at a macroscopic level. Specimens may be aggregations of smaller irregularly formed (anhedral) crystals (often visible only as grains), creating a texture ranging from fine- to coarsegrained or, if space and environmental conditions permit, each individual quartz crystal can grow to such a size that its crystal structure is regular (euhedral) and easily visible to the naked eye (Rafferty 2012:231). Crystal size forms a continuum, and both large euhedral crystals and aggregations of anhedral crystals (of various sizes) may be found within the same vein, since crystal size is related to the rate at which the liquid SiO2 cools and the space available for crystal formation (see also Chapter 3.2). While Hallam’s (1977:169) characterisation of some quartz as ‘horrid little bits of stone’ was made in jest, this in fact reflects the attitudes of many analysts to macrocrystalline quartz due to the heterogeneity of the material and its unpredictable fracture patterns. This complicates typological analysis of flaked quartz as well as the subclassification of quartz materials. In isotropic materials (e.g. obsidian, single large quartz crystals), physical properties are identical in every direction. These materials flake predictably because the fracture path is dictated by core morphology and knapping technique (Cotterell and Kamminga 1987). However, for aggregations of quartz crystals, fracture follows an easier, predefined path along crystal boundaries – the larger the crystal facets, the more influence these have over the shape of a flake. Aggregations of small crystals may flake fairly well, as the fracture can largely follow its own path, while aggregations of larger crystals will flake less predicably and produce blocky, chunky flakes. Many varieties of quartz can complicate archaeological analyses, then, for the following reasons: 178 1. Flakes may lack regular flake landmarks, where detached along internal flaw planes or large crystal facets (Holdaway and Stern 2004:116–118; Witter 1990:43); 2. Quartz has a higher fragmentation rate than other materials, so flakes are frequently split as they are removed from the core; no single knapping variable controls fragmentation (Driscoll 2011a; Tallavaara et al. 2010). Broken flakes may be misclassified as flaked pieces, especially when material is of poor-quality, which will affect artefact counts; 3. It may be difficult to identify retouch or other features, especially on coarser grained material. These issues are at the core of the ongoing argument of whether quartz-dominated assemblages are best represented by an artefact typology designed specifically for quartz and accounting for its irregular flaking properties, or whether existing typologies designed for flint-like materials can be successfully applied (see summary in Driscoll 2010:78–80). Driscoll (2011b) found that even where experienced analysts were provided with a typology and description of quartz artefact types, and then asked to classify an experimentally produced assemblage accordingly, their classifications varied widely; clearly a universally adopted typology would not alleviate all the difficulties associated with quartz artefact analysis. The heterogeneity of macrocrystalline quartz makes subclassification difficult, as texture, colour and opacity all form a continuum rather than a series of discrete categories. In archaeological literature, quartz is generally divided into two categories based on colour and opacity, using some variation of the crystal quartz/clear quartz/rock crystal vs. milky quartz/vein quartz system. This may be a reflection of the gemmological classification system, which relies mostly on the colour of the material to subdivide quartz (e.g. rose quartz, smoky quartz), or a consequence of Dickson’s (1977:102) argument that translucency is the strongest indicator of how well quartz will flake. As noted above, however, it is crystal size and the presence of internal flaw planes that actually determines how well quartz flakes. Single quartz crystals may be transparent, translucent, or opaque – the traditional archaeological colour-based classification scheme does therefore not address how flakeable individual types are. Martinez Cortizas and Llana (1996) proposed four 179 morphostructural groups of quartz based on crystal size and the presence of internal flaw planes; these have been demonstrated to more accurately characterise flaking predictability (de Lombera-Hermida 2008, 2009). In practice, however, it is difficult to determine quartz formation processes at an artefactual level – indeed, Driscoll (2011a) demonstrated that it is often impossible to accurately distinguish grain size without microscopic analysis, perhaps explaining why the morphostructural classification system has not been widely adopted. 9.3.2 Stone Artefact Classification Classification systems for flaked stone artefacts can generally be characterised as either typological or materialist/technological. In the former system, artefacts are classified into different types based on morphological characteristics (e.g. Bird 1985), with a distinct emphasis placed on the quantity of retouched pieces or ‘formal tools’ (e.g. geometric microliths, backed blades, points). In a materialist system, pieces are classified into mutually exclusive categories based on the position within the stone reduction sequence, thereby examining the form of the artefact ‘in terms of the mechanisms by which it was created, rather than the presumed purpose for which it was created’ (Hiscock 2007:202). The main criticism of typological analyses is the underlying assumption that each morphological type represents a distinct and meaningful category to those who made and used the tools. In fact, it has been demonstrated that one particular morphological type – the geometric microlith – did not have one distinct function but was rather used for a multitude of everyday tasks (e.g. Robertson 2009; Robertson et al. 2009). Other backed artefacts have been shown to have similarly varied functions (Fullagar et al. 2009; McDonald et al. 2007; Robertson 2005; Robertson and Attenbrow 2008). Thus, despite the geometric microlith being a more aesthetically pleasing form (and one that requires a greater investment of time to manufacture), this does not necessitate its use solely for ‘higher’ functions. Similarly, Taipale et al. (2014) and Knutsson et al. (2015) demonstrated that quartz pieces previously classified as flaking debris had in fact been used as tools, without further modification, again questioning the inferred link between form and function. In contrast, a materialist approach classifies artefacts without reference to how (or indeed if) they were used, merely how they were created (Hiscock 2007). 180 A materialist approach was employed herein as these systems accurately separate different stages in the reduction sequence, so can be easily employed in concert with Kuhn's (1995) provisioning systems that rely on technological attributes rather than the presence of distinct artefact forms (see Chapter 2.4). The classification system was modelled on Hiscock's (2007), with some minor adjustments to better reflect the quartz-dominated assemblages found in the study area. The procedural key below was used to evaluate potentially flaked stone, separating unflaked stone from flaked fragments, cores, flakes and retouched flakes (Figure 9.2). Some unflaked stone was non-local – where these pieces were unlikely to have arrived on-site via natural processes, they were classified as manuports (Table 9.1). Figure 9.2 Procedural key used to classify flaked stone via the materialist approach developed by Hiscock (2007). Unflaked stone preserves no evidence of flake removals or having been detached from a larger piece. Note that ventral surfaces were identified via any of the normal landmarks, such as bulb of percussion, ripple marks, termination etc, following Holdaway and Stern (2004:108). Retouch scars must be initiated from or modify the ventral surface – see Table 9.1 and Appendix D.2 for more information. Note the amendments for quartz-dominated assemblages, including the use of impact points or shatter facets to identify flaked quartz, as well as flaked fragments incorporating those pieces where breakages or poor-quality material make it difficult to confirm the presence/absence of a ventral surface. Similarly, poor-quality material likely masked retouch in several instances – where such uncertainty was encountered, pieces were classified as in the lower category, as flakes. 181 Table 9.1 Artefact type definitions and the attributes recorded for each. Flaked fragments were separated based on their maximum dimension, and flakes by their length. Artefact Attributes to record Core Any piece that exhibits a negative flake scar and/or shatter facets, but no ventral surface. Flakes have been struck from the core, but it preserves no evidence of having been detached from a larger piece. 1. 2. 3. 4. 5. 6. 7. 8. 9. Flaked fragment Any piece preserving negative flake scars and/or shatter facets but cannot be confidently assigned as either a core or a flake – the presence/absence of a ventral surface is uncertain. This may incorporate broken pieces, or those on such poor-quality raw material that features are difficult to identify. Flaked fragments have clearly been modified by humans but cannot be classified further. Small fragments (< 10 mm): 1. Raw material; 2. Combined weight and no; Flake 1 Definition Any piece that has been detached from a core, and not retouched (though may exhibit other edge damage, breakage). In contrast to most definitions of a flake (e.g. Holdaway and Stern 2004), even those devised specifically for quartz, flakes do not need to preserve evidence of conchoidal fracture (e.g. bulb of percussion) or fracture plane propagation (impact point, zone of impact shatter), simply a ventral surface indicating a previous attachment to a larger piece. These may demonstrate edge battering or breakage and may or may not have been used. Retouched flake A flake that has had further flakes struck from it, i.e. where flakes have been initiated from and/or modify the ventral surface or the interface between the dorsal and ventral surface (Hiscock 2007). Scars must be complete, to distinguish retouch from previous flake removals. Pieces exhibiting only microscars or edge battering were not considered retouched, since these result from use or post-depositional processes. Manuport Pieces of non-local unmodified stone = for complete/nominally broken pieces only 2 1. 2. 3. 4. 5. 6. 7. Raw material; Individual weight; Maximum dimensions; No. negative flake scars; Grams/scar; No. platforms used; Bipolar; Unmodified surface; Other. Large fragments (≥ 10 mm): Raw material; Individual weight; Maximum dimensions; No. negative flake scars; Grams/scar; No. platforms used; Other. Small flakes (< 10 mm): 1. Raw material; 2. Combined weight and no; 3. Breakage frequency. 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. Large flakes (≥ 10 mm): Raw material; Individual weight; Oriented/max. dimensions; No. negative flake scars1; Mm/scar2; Flake scar orientation1; Unmodified surface1; Bipolar; Breakage; Edge damage; Other. 1. 2. 3. 4. 5. 6. 7. Raw material; Individual weight; Oriented/max. dimensions; Type of retouch; Edge damage; Breakage; Other. 1. Raw material; 2. Individual weight. = for orientable flakes only 182 Table 9.2 Stone artefact attributes – further information regarding methods can be found in Appendix D. Negative flake scars < 5 mm in maximum/percussion length have been omitted as they more likely result from retouch or use, and single scars can be very difficult to separate (and therefore quantify) on all but very high-quality quartz. The references provided point to a definition of a particular attribute, and its method of measurement. Attribute Summary method and/or variables Raw material Type: quartz (and quality), BIF, chalcedony, silcrete. Weight Individual or combined, depending on artefact type. Reference Oriented Percussion Length of fracture plane (i.e. following the flaking axis) Holdaway and dimensions length from platform to termination in a straight line. Stern 2004:138–140 Width at At halfway point of percussion length line, at right midpoint angle. Thickness Taken at the point where percussion length and width at midpoint at midpoint intersect. Maximum Maximum dimensions length The longest line that can be drawn between two points on the piece. Maximum width Longest line that can be drawn at right angle to maximum length. Maximum thickness Longest line that can be drawn at right angle to maximum thickness. Breakage For small flakes, the total number of broken pieces was noted. For large flakes and retouched artefacts, breakage type (transverse/longitudinal/marginal) and the type of fragment remaining were noted. Number of negative flake scars Complete and partial scars ≥ 5 mm percussion/max. length, excluding retouch/damage scars. Weight/length per scar No. flake scars in relation to artefact size, classified by percussion length (mm/scar) or weight (g/scar). Negative flake scar orientation Quadrats (1–4) from which negative flake scars ≥ 5 mm have been initiated. Number of platforms No. platforms used to create flake scars. Proportion of unmodified surface The proportion of unmodified surface evident on a core or a dorsal flake surface: < 25%, 25–50%, > 50– 75%, > 75%. Bipolar Presence/absence/uncertain. Edge damage Presence of edge damage noted, and general description (location, extent). Does not refer to retouch or recent/fresh breaks (sieving etc). Retouch type Backing, unifacial retouch. Other Other notable features, e.g. presence of hafting resin, original crystal facets, parallel flake scars. Holdaway and Stern 2004:138–140, 189 Hiscock 2002b Holdaway and Stern 2004:146–147. Holdaway and Stern 2004:144, 193 Holdaway and Stern 2004:158–160. 183 The benefits of the materialist approach are clear – it removes the subjectivity from flaked artefact classification and voids the argument about whether quartz artefacts are best classified with a flint- or quartz-based typology. Nevertheless, poor-quality quartz limited the visibility of certain features, likely artificially enhancing the number of flaked fragments (i.e. flaked artefacts that cannot be classified further) and reducing the number of retouched items where coarse textures mask retouch scars. While not exclusive to quartz, bipolar flaking can sometimes split cobbles so that positive and negative surfaces cannot be identified, i.e. both positive and negative scars are lacking (Holdaway and Stern 2004:196). These items would likely be classified as unflaked stone, unless other features were present. Nevertheless, misclassification is less problematic than within typological systems as it is the distribution of attributes that is the primary focus under a materialist approach, rather than the frequency of certain artefact types. 9.3.3 Attribute Analysis A number of metric and non-metric attributes were recorded for stone artefacts, depending on both the classification and size of each piece (Tables 9.1–9.2; see detailed methodological information in Appendix D). The primary aim of the attribute analysis was to quantify reduction intensity (i.e. how heavily materials have been worked and/or reworked) since this accurately separates Individual and Place Provisioning strategies (Kuhn 1995; see Chapter 2.4). It was also important to consider other features of relevance to Kuhn's (1995) technological provisioning systems, as well as occupation intensity and assemblage composition. Specific attributes were selected for their relevance to the aims above, as well as their ease of observation on (often poor-quality) quartz, which dominated assemblages. Most raw material types were fairly easily classified, since they represent those commonly found in southwestern Australia. Unfamiliar types were identified following standard non-invasive geological techniques, based on colour, texture, weight and structure, as well as comparison to type specimens or general texts (e.g. Blatt and Tracy 1996; Hefferan and O'Brien 2010; Perkins 2011). Four raw material types were recognised, as below: 184 o Quartz: This material varied widely from glassy smooth pieces to coarsetextured fragments with visible fracture planes, crystal aggregations and/or flaws (Figure 9.3); o Silcrete: A metasedimentary rock comprised of reworked grains cemented within a silica matric (Langford-Smith 1978:3). Grain size and angularity varied considerably. Specimens with larger, more angular grains would be more accurately classified as silicified granite but have been grouped with silcrete as they derive from the same weathering profile, at a greater depth (Taylor and Eggleton 2017); o BIF: Banded Iron Formation, chert and/or ironstone. These categories were combined because they cannot be visually distinguished from small samples, and may have derived from the same source rock, probably in the greenstone belt east of the study area (Figure 3.4); o Chalcedony: Chemically identical to quartz but forms under different conditions; it is differentiated from quartz by a waxy lustre. It possibly originates from chalcedonic silica caprock outside the study area (Figure 3.4: Czj). Raw material quality is a major factor that separates expedient technologies from transported toolkits, with higher quality materials selected for the latter. It was therefore necessary to identify these materials, since expedient strategies create considerably more archaeological debris that may mask signs of alternate provisioning systems. BIF and chalcedony were always high-quality, while silcrete never was; quartz, however, was incredibly variable. High-quality quartz was defined as the finer grained (or entirely smooth/glassy) material that lacked visible flaws/fracture planes/crystal aggregations (Figure 9.3). Considering the quartzdominated assemblages, the quality classification of quartz was integral. However, it must be treated with caution since, as noted above (see Chapter 9.3.1), quartz can be highly variable within a single vein/specimen. Therefore, a high-quality quartz piece may not always be indicative of the qualities of its parent core. Furthermore, it is not always possible to accurately distinguish quartz quality at a macroscopic level (Driscoll 2011a). 185 Figure 9.3 Variable quality of quartz within the study area. Quality increases from left to right. Top row: very poor. Second row: poor–moderate. Third row: moderate–good. Bottom row: high quality. White bars are 1 cm. Reduction intensity was measured using the standard indicators: negative flake scar and platform counts, proportion of unmodified surface (as quartz rarely displays a cortex) and frequency of bipolar reduction (Table 9.2). Flakes/cores with more unmodified surface represent earlier stages in the reduction sequence, while flake scar and platform counts increase as the sequence lengthens as does the incidence of bipolar reduction. Retouch intensity was not investigated – while several measures accurately characterise the loss of mass through retouch (e.g. Clarkson 2002; Kuhn 1990) the link to reduction intensity only exists where retouch was performed for edge renewal or artefact repair. In the study area, most retouched artefacts were backed, so the 186 retouch formed part of the initial manufacture sequence, and was not linked to ongoing maintenance to extend artefact use-life. Similarly, standard indicators of reduction intensity were not applied to retouched artefacts, as they had been heavily modified from their original form. Artefact size is often linked to reduction intensity, with larger pieces argued to represent earlier stages in the reduction sequence since as a core reduces in size, only smaller flakes may be detached from it (Clarkson and O'Connor 2014). However, Lin et al. (2016) demonstrated that neither core size nor reduction intensity patterned the size distribution of artefacts in a measurable way, because the products are so overwhelmingly small that any increase in large flakes has a negligible effect at an assemblage level. Therefore, artefact dimensions were only linked to reduction intensity in certain instances, i.e. where a core was so small that it could not be reduced further. Elsewhere, artefact size was accommodated in a different way. Early stage analysis indicated that flake scar counts were heavily biased by artefact size – larger artefacts (particularly flakes) were more likely to exhibit higher scar counts than smaller artefacts, since they had larger surfaces on which to preserve flake scar evidence. Furthermore, the same flake scar count may have different implications for reduction intensity according to the artefact size. For example, a core with six flake scars may indicate higher reduction intensity if it weighs only a few grams, but far less intensive reduction if it weighs 50 grams. For flakes, scar count was considered against percussion length, by dividing the total length by the number of scars, to derive a mm/scar value. The same technique was used for cores and large fragments, substituting weight for length since it better represents size for more complex three-dimensional pieces. By allowing for artefact size, these measures more accurately represent reduction intensity, and also reduce the bias introduced by comparing larger artefacts to smaller ones. Where present, a number of qualitative measures were also noted. Hafting resin – normally a dark, solid encrustation on one or more margins – is an important feature since it may indicate the use of a mobile, multifunctional toolkit incorporating redundant features (Hiscock 1994; Torrence 1983). The presence of original crystal facets indicates that a piece of quartz came from a large natural crystal. These crystals would be ideal as transported cores, since the material therein is often 187 homogenous and of higher quality than when it comprises an aggregation of smaller crystals. The presence of elongated, parallel flake scars was noted, as it may indicate more formal, standardised reduction techniques that permit raw material conservation. It was also important to quantify artefact breakage since it may be indicative of discard behaviour. Further, transverse snaps are commonly associated with trampling, which can be linked to greater occupation intensity (Hall and Love 1985; Hiscock 1988:22; Holdaway and Stern 2004:114). Severe edge damage was noted, as it may have prompted artefact discard, or resulted from trampling. 9.4 OTHER ANALYSED MATERIALS While lithics were the dominant form of cultural material found at the sampled sites, charcoal was present, as were small quantities of European glass and ochre. Sediment samples were also collected for contextual information, as soil conditions can impact the preservation of organics. Each of these material types, and the methods of analysis, are treated separately below. 9.4.1 Charcoal Charcoal was bagged by XU and weighed. Certain samples were selected for radiocarbon dating according to the concentration of cultural material within the same XU (i.e. to date peaks in discard rate, or the earliest/latest cultural material), the depth of the sample compared to the overall depth of deposit, and the quantity of charcoal preserved. Where sample size permitted, in situ samples were chosen over fragments recovered from sieve residue, as very small fragments can be blown into test-pits during excavation or moved downwards through the deposit by insect activity. Samples were sent to the Radiocarbon Dating Laboratory at the University of Waikato, New Zealand – most were sufficiently small that AMS (Accelerator Mass Spectrometry) dating was required. The samples selected and the returned age determinations are provided in Chapters 10–12. 9.4.2 Sediments Small sediment samples were collected from certain pits/XUs primarily for pH measurements, as a soil that is acid (pH < 7), neutral (pH 7) or alkaline (pH > 7) can influence the preservation of organics such as bone or charcoal (Barham and 188 Huckleberry 2014). It must be noted, however, that pH is just one of many variables that influence the preservation of organics. For example, bone preservation is dictated by intrinsic factors (those of the bone itself, i.e. density, age, porosity, whether it had been cooked etc.) as well as extrinsic or environmental factors, such as soil chemistry, pH, moisture content and temperature (Beisaw 1996:7–8; Gordon and Buikstra 1981). Nevertheless, pH provides a broad indication of how well organics may be preserved had they originally been present in archaeological deposits. While bone is better preserved in more alkaline environments (pH > 7), these same environments are less ideal for charcoal preservation (Barham and Huckleberry 2014; Braadbaart et al. 2009; Gordon and Buikstra 1981). Soil pH was tested with an over the counter colourimetric test kit. While such kits are less accurate than pH readings obtained via electrometric methods, they are commonly employed in archaeological analyses due to their accessibility (Barham and Huckleberry 2014). Sediment samples were tested after the completion of fieldwork, once materials had been transported to Perth. Specific samples and the results of these are provided in Chapters 10–12. It is important to note, however, that soil pH is a dynamic variable that results from the interplay of chemical and physical properties of the soil and can therefore change fairly rapidly – current levels may not be indicative of those in the past (Barham and Huckleberry 2014). 9.4.3 Ochre Ochre is a natural pigment whose colour is derived from iron-oxides, iron-sulphides, and other minerals suspended within a matrix, generally clay (Wallis et al. 2016). It is fairly easily identified by vibrant colours, greasy texture and small particle-size, and should produce a colour-streak when abraded against a surface (Wallis et al. 2016; Watts 2010). Ochre was sorted by colour family (i.e. reds, yellows), and fragments counted and weighed within each XU. No further analysis was undertaken, as sourcing requires specialist equipment and techniques, as well as reference samples (e.g. Green and Watling 2007; Scadding et al. 2015; Wallis et al. 2016). 9.4.4 Glass Glass was roughly categorised as modern and historical – obviously modern glass (i.e. Coca Cola bottles, beer bottles) was not collected unless it occurred below the 189 ground surface, where it may be indicative of post-depositional disturbance. Historical glass requires evaluation to determine if it had been intentionally flaked, as there is well-documented, continent-wide archaeological and historical evidence for Aboriginal people flaking glass, primarily bottle glass (e.g. Allen 1969; Goward 2011; Harrison 2000; Horne and Aiston 1928:11; Moore 1884:28–29; Paterson 1999; Tindale 1941; Williamson 2004). Glass may have been desirable due to its isotropism, meaning it flakes more predictably than most types of stone (Cotterell and Kamminga 1987). Where flaked glass is present, this indicates that Aboriginal people were visiting sites after European arrival and, where it can be determined, after the date the vessel was manufactured. Within the context of this research, worked glass has significant implications for site-use history, and is a valuable method of indirectly dating deposits and assemblages. Several factors impede the identification of intentionally flaked glass. First, the low fracture toughness means that glass breaks more easily than stone and is more susceptible to post-depositional damage, the results of which often mimic retouch (Allen and Jones 1980; Beaumont 1961; Conte and Romero 2008; Cotterell and Kamminga 1987). Furthermore, the inherent differences between glass bottles and stone cores mean that flaking methodology can be varied. Thick bottle bases can be used as cores (base core reduction) and flaked like stone, producing cores and flakes with the standard landmarks (Goward 2011:16; Harrison 2000). The alternate strategy, known as worked fragment reduction, involves simply smashing a bottle and choosing the fragments best suited to a particular task, often without modifying them further (Goward 2011:16; Harrison 2000). Unless they have been retouched into a formal tool type, the products of such a strategy can be incredibly difficult to identify – various studies have confirmed the use of expedient fragments through microwear and residue analysis, even where these pieces lack macroscopic edge damage (Goward 2011:22; Martindale and Jurakic 2006; Wolski and Loy 1999). Various attempts have been made to define criteria against which to evaluate potentially flaked glass (e.g. Allen and Jones 1980; Cooper and Bowdler 1988; Goward 2011; Harrison 2000; Paterson 1999; Runnells 1976; Veth and O'Connor 2005; Williamson 2004). The bulk of these studies are concerned with the base core reduction technique, particularly distinguishing retouch from post-depositional 190 damage. Despite decades of research, no set of universally applicable criteria has been formulated and context is often considered the most important criterion of all (e.g. Allen and Jones 1980; Beaumont 1961). Effectively, this means that the same piece might be considered artefactual if it was discovered in an isolated, out of the way location, while it would be evaluated more critically were it found in a high-traffic (foot or vehicular) area. Herein, glass was divided into modified and unmodified – further classification was unnecessary as the focus was primarily on flaked glass as a means of dating deposits or assemblages derived from Aboriginal occupation. Table 9.3 lists the criteria used to evaluate glass fragments – the criteria selected were those relevant to the small, fragmentary glass assemblage that was encountered. It is important to note, however, that the absence of microwear and residue analyses means that expediently used fragments (classified as modified glass) will be underrepresented in the assemblage. Glass assemblages were too small to provide meaningful data on reduction intensity – therefore, only basic descriptive attributes were recorded (Table 9.3). No attempt was made to date individual glass fragments. Table 9.3 Features used to identify modified and unmodified glass, and the attributes recorded for each. To be classified as modified glass, the piece needed to meet at least one of the criteria listed – all other pieces were classified as unmodified. Criteria are based on Allen and Jones (1980), Goward (2011:97), Harrison (2000), Holdaway and Stern (2004:30), Paterson (1999), Runnels (1976), Veth and O'Connor (2005) and Williamson (2004). Maximum dimensions were recorded via the same method used for stone artefacts (see Table 9.2 and Appendix D). Edge damage is only considered evidence of modification only where the site is relatively undisturbed, i.e. receives little foot or vehicular traffic. Type Modified glass Criteria o o o o Unworked glass o Features of a normal conchoidal flake (e.g. bulb of percussion, etc); Physical resemblance to known formal-tool type (e.g. geometric microlith); Edge damage (undisturbed sites only); Unmistakable/unequivocal retouch, characterised by: • Continuous flake removals (i.e. individual scars are not separated by unmodified areas), which may be overlapping; • Restricted to one or two edges; • Regular-sized flake removals; • Uniform direction of force (though backing may be bidirectional); • Consistent scar patina i.e. flake removals produced in a single event. None of the above. Attributes 1. 2. 3. 4. 1. Individual weight; Maximum dimensions; Colour; Type of modification. Combined weight. 191 9.5 DATA ANALYSIS Several steps were required in the data analysis process to accurately merge the theoretical model with archaeological data. First, the longevity of site-specific water sources had to be modelled. Then, a temporal framework had to be constructed to evaluate change over time. It was then necessary to merge artefact data with theoretical models by identifying technological signatures that should accompany various provisioning systems and the indicators of occupation intensity therein. 9.5.1 Quantifying Site-Specific Water Sources While the data from Chapters 6–7 are deemed representative of food resources and soil water available at each archaeological site, Chapter 8 demonstrated that the longevity and volume of water varied considerably according to gnamma morphology. Therefore, it was necessary to make some attempt to more accurately represent the quantity and longevity of water at each studied site, since this would directly impact occupation span under various rainfall conditions. For each gnamma, the following were considered: o Depth; o Surface area; o Shape (cylinder, elliptic cylinder, cone, inverted triangular prism – Figure 9.4); o Impact of above on evaporation coefficient; o Level of runoff (none, 2x, 5x). It was fairly straightforward to track the daily water balance for cylinders and elliptic cylinders, as each millimetre of depth consistently represents the same volume of water; this is easily calculated using the structure's surface area. These gnammas were modelled following the methods outlined in Chapter 5.4.3. Cones and inverted triangular prisms, however, narrow with depth. Therefore, a given quantity of water (e.g. a 5 mm rainfall event) will be stored over different vertical distances depending on the water level at the time of the fall. Similarly, 5 mm of daily evaporation will represent a different volume of water, depending on the depth at which water is stored. As a result, both rainfall and evaporation must be considered in terms of water depth and volume. Similarly, evaporation coefficients had to be altered for 192 these structures, since their sloping walls offer less protection by shadow-casting, and water is exposed to direct sunlight for a greater portion of the day. Figure 9.4 Shapes used to represent morphological variation between individual gnammas, and the mathematical formulae used to determine their total capacity and the volume of water with depth. 193 Using formulae shown in Figure 9.4, which assumes that structures narrow at a consistent rate, it was possible to construct an Excel spreadsheet to calculate the volume of water held by each structure as the depth of water changed. The daily water balance was then calculated using the following measures, in sequence: o Daily rainfall volume: surface area of structure x rainfall amount (allowing for runoff multiplier), then converted to litres; o Pre-evaporation daily water volume: above added to the previous day's volume (or zero, for the first model day); o Pre-evaporation water level: above converted to mm; o Post-evaporation water level: above minus daily evaporation with appropriate coefficient applied; o Post-evaporation daily water volume: above converted to litres. The water balance was tracked for each of the modelled years used previously, representing periods of low, average and high winter rainfall. The water-holding periods for each gnamma are provided in Chapters 10–12, along with the specific values used to calculate them. 9.5.2 Constructing a Temporal Framework A major goal of this research was to investigate change over time, which inherently requires good temporal control within test-pits. In stratigraphically homogenous deposits, the excavation unit (XU) forms the primary analytical unit. Standard archaeological practice dictates that all material contained within a single XU is considered contemporaneous, but it is likely a palimpsest of several occupation events. Where available, radiocarbon dates were treated as the basal date for their XU of origin. Calibration provides a date range within a certain confidence interval, but this can complicate analysis of temporal change. Therefore, the midpoint of the date range at 95.4% confidence (as cited on the data sheets returned by the University of Waikato) was used as an indication of the calibrated basal date for a particular XU. For consistency, all dates sourced before 2013 were recalibrated using OxCal 4.2.4 (Bronk Ramsey 2009) and SHCal13, southern hemisphere atmospheric data (Hogg et al. 2013). If an XU or surface accumulation contained historical glass, these were presumed to post-date European arrival in southwestern 194 Australia (1826), though many likely post-dated the 1860s, when European people were in the vicinity of Hyden (see Chapter 4.3). Due to the financial cost and sample size requirements of radiocarbon dating, it is rarely viable to obtain more than a few age determinations for a single excavated pit. It was therefore necessary to construct a depth-age curve that effectively fills in the blanks between radiocarbon dates by assuming a constant sedimentation rate – this allows a broad date to be assigned to undated XUs. A curve was constructed for each dated pit, but the data were not applied to undated portions of the site, as there is no reason to expect sedimentation to occur at the same rate. 9.5.3 Identifying Technological Strategies In order to interpret attribute data and archaeological remains, it was necessary to construct technological signatures outlining what should be found at a site characterised by Individual or Place Provisioning strategies. While there is a wealth of literature on the subject (e.g. Clarkson 2004, 2006, 2007; Graf 2010; Kuhn 1995; Mackay 2005; Mitchell 2016), the focus must be on those features preserved in the archaeological record of the study area and identifiable within the data collected during this research. It is also important to consider the subsidiary strategies that may occur within the larger technological system, such as the manufacture of expedient tools to supplement the transported toolkit required by the Individual Provisioning strategy, or ‘gearing up’ by preparing a toolkit for the long logistical forays that may be necessary under Place Provisioning (see also Chapter 2.4). The archaeological literature rarely addresses the signatures associated with these supplementary strategies, so some attempt was made to define these herein by considering the similarities and differences between Gearing Up or Short-term Expedient strategies (as they are referred to hereafter), and the dominant Individual or Place Provisioning strategies. It was first necessary to consider the major differences in technological organisation between each of the four systems (Figure 9.5), as this will impact the attributes that serve as key indicators. Raw material quality represents a good primary indicator, since higher quality materials are required for transported toolkits. In contrast, expedient technologies can accommodate poorer quality raw materials, which are 195 generally more abundant in the landscape. While high-quality material may be stockpiled under Place Provisioning, its limited distribution would discourage reduction in a more expedient manner, particularly since lower quality materials would be readily available. Instead, it would probably be reserved for certain artefacts, such as those required for logistical forays, when use-life must be extended. Therefore, it was assumed that high-quality materials derived from Individual Provisioning or Gearing Up, while the remainder represented Place Provisioning or Short-term Expedient strategies. Figure 9.5 Major differences in technological organisation between the four provisioning systems: Place Provisioning (PP), Individual Provisioning (IP), Gearing Up (GU), and Short-term Expedient (EXP). It was then possible to separate the paired strategies based on aspects of technological organisation. Because Place Provisioning accompanies longer site visits, reduction sequences are longer, reduction intensity is higher, and some raw material conservation may be required; the opposite is true under Short-term Expedient strategies. While Gearing Up is a form of Individual Provisioning, it differs 196 in some important ways that should be visible at an assemblage level. In contrast to traditional Individual Provisioning, the toolkit is not required for extended, ongoing use, but rather for a particular journey, after which people will return to the same base camp. Therefore, the location and availability of replacement stone is known; this should translate to lower levels of reduction intensity and raw material conservation. The reduction sequence is longer under Individual Provisioning but, due to the higher residential mobility, only a small part of the sequence will be represented at any particular site. Using the technological contrasts set up above, it was possible to formulate a list of attributes that should characterise assemblages arising from each provisioning system (Table 9.4). Only the most reliable indicators were selected, as sample size/bias was often an issue, especially for assemblages on high-quality raw materials. The focus is on attributes rather than artefact type frequency, but in some cases the latter was indicative. For example, retouched pieces will be rare under Place Provisioning, but could occur if they were required for a specific task. In contrast, they should never be found within Short-term Expedient strategies, since formal tools would have been available in the transported kit. These signatures were evaluated against the archaeological data. When assemblages from several XUs shared the same characteristics, these were grouped to increase sample size and improve the visibility of provisioning systems. Nevertheless, small sample sizes precluded the use of more complex statistical measures. Instead, the focus was on simple expressions of central tendency along with graphical representations of data. Indeed, as Cowgill (2015) notes, these simpler patterns and methods are often overlooked in favour of more complex statistical techniques, to the detriment of broader research questions. 197 Table 9.4 Artefact attribute indicators for dominant and supplementary technological provisioning systems, using the pairing system established in Figure 9.5. The indicators are only of value within paired systems (Individual Provisioning and Gearing Up, Place Provisioning and Short-term Expedient), since they describe the relative position of each strategy on a continuum, whose range of values will be different to that of the alternate pair. Attribute Place Provisioning Short-term Expedient Discard rate/assemblage size Larger Smaller Negative flake scar average Higher Lower Negative flake scar range Higher Lower Platform average Higher Lower Platform range Higher Lower g/scar Lower Higher mm/scar Lower Higher Bipolar core frequency Higher Lower Retouched pieces Some None Individual Provisioning Gearing Up Breakage rate Higher Lower Heavy edge damage rate Higher Lower Standardisation rate Higher Lower Discard of usable artefacts Lower Higher Bipolar core frequency Higher Lower Negative flake scar average Higher Lower Platform average Higher Lower g/scar Lower Higher mm/scar Lower Higher Unmodified core surface Lower Higher Attribute 198 9.5.4 Quantifying Occupation Intensity A major aim of this research was to evaluate occupation intensity over time. Intensity shifts in response to the number of visitors and visit duration, but these factors are difficult to separate, so intensity is best considered in terms of the number of peopledays spent at a site. The dominant provisioning systems inherently indicate different levels of occupation intensity: once visits become longer and more predictable, Place Provisioning replaces Individual Provisioning. However, this switch merely means that a certain threshold has been reached, as occupation may still further intensify within a Place Provisioning system. Considering this strategy produces the most archaeological debris, such shifts should be readily identifiable in the stone artefact assemblages. A number of indicators were considered: o Evidence of Gearing Up; o Proportion of bipolar cores; o Range of negative flake scar counts; o Range of platform counts; o Range of unmodified surface; o Post-depositional breakage; o Artefact discard rates. Most of these attributes reflect the length of the reduction sequence, so higher values indicate the longer sequences that occur during more intense occupation periods. Post-depositional breakage may be higher where activity is more intense, while Gearing Up indicates restricted residential mobility and the need for long logistical forays to supplement the local resource base. However, none of the indicators are reliable enough to be used in isolation. For example, it is often hard to separate post-depositional breakage from pieces that broke during use. Similarly, Gearing Up activities may have occurred in a different part of the site and so be absent in particular deposits, or may represent preparation for long trips unrelated to resource procurement (i.e. aggregation events). If at least four high-intensity indicators were identified, this was considered a sufficiently robust measure of occupation intensity. The lack of any specific indicator was not considered particularly important, as small pit sizes meant that some attributes (e.g. discard rates, bipolar cores) will be sensitive to sampling biases. 199 While the data above relates to individual assemblages, it is necessary to consider occupation intensity at a site-wide level, rather than treating each pit or area separately. For example, Individual Provisioning may dominate in one pit but if contemporaneous Place Provisioning assemblages were found in another part of the site, then the former does not represent a large-scale reduction in occupation intensity. Instead, it probably indicates that different parts of the site were used in different ways, or at different times of year. Considering this, it was possible to identify three levels of occupation intensity in the archaeological record: o Low: dominated by Individual Provisioning and/or Short-term Expedient strategy across all deposits representing a particular occupation period. Represents short, unpredictable occupation; o Moderate: Place Provisioning, with no/few high intensity indicators. Represents longer, more predictable occupation; o High: Place Provisioning with at least four high intensity indicators. Represents longest, most intense occupation. 9.6 CONCLUSION The main archaeological sample used in this research comprises three locations (Anderson Rocks, Gibb Rock and Mulka's Cave), all located within the Granite Landscape Division. A fourth site – Twine Reserve – had to be abandoned due to logistical and financial constraints. A number of 500 x 500 mm test-pits were excavated at each site to permit intra-site spatial analysis without exceeding the one square metre excavation limit. Lithic analysis methods were heavily influenced by the properties of quartz, a material that dominates the study area assemblages. Quartz is highly variable and can fracture irregularly, often lacking regular hallmarks of intentional flaking. The ambiguities of this material are minimised within a materialist approach rather than the standard typological classification schemes; under a materialist approach, attribute data is of greater importance than artefact type frequency. Attributes were chosen to reflect reduction intensity, as this is a key factor in separating Kuhn's (1995) Individual and Place Provisioning systems. Other materials (e.g. sediment, glass) are mostly of value in terms of quantifying the preservation environment or dating deposits and assemblages. The formulated 200 technological signatures, covering both dominant and subsidiary strategies, will allow attribute data to be interpreted within the framework of Kuhn's (1995) technological provisioning systems. Occupation intensity can be evaluated by considering the dominant provisioning system and the presence of high-intensity indicators, primarily based on the length of the reduction sequence. Site-specific water data will allow these data to be interpreted within the framework of resource availability. 201 CHAPTER 10 ANDERSON ROCKS 10.1 INTRODUCTION This chapter discusses the landscape and archaeological content of Anderson Rocks. To begin, the natural setting and characteristics of the outcrop are described, before consideration turns to the availability of water in pans, soils and – most significantly – the three gnammas. Field methods are then discussed, encompassing the surface collections and the three excavated pits. The charcoal assemblage is quantified, with specific reference to the influence of soil conditions on charcoal preservation and the radiocarbon age determinations derived from selected samples. The stone artefact assemblages are then described, focussing on the frequency and distribution of artefactual stone and the major trends within attribute data. Finally, the small ochre collection is briefly evaluated. 10.2 SITE DESCRIPTION Anderson Rocks is 31 km north of Hyden, within the Granite Landscape Division. This area is bordered by Thicket to the north and elsewhere by Heath, with the surrounding Mallee at least 2.5 km away (Figure 9.1). The outcrop itself reaches a height of 392 m ASL and covers approximately half a square kilometre. Several long quartz veins are visible on the outcrop, measuring from 4.5 to 16 cm wide (Figure 10.1a–b). Anderson Rocks was originally classified as an 'other heritage place (lodged)' on the DPLH register. This label applies to locations that are potentially Aboriginal sites but have yet to be evaluated. During the course of this research DAA (now DPLH) completed the evaluation and stated that Anderson Rocks did not meet the criteria of a site under Section 5 of the Aboriginal Heritage Act 1972. The area is therefore not subject to the conditions imposed by the AHA. 202 B A E C D F I G H Figure 10.1 Features seen at Anderson Rocks – all photographs by A.M. Rossi. Individual pictured is 1.64 m tall. A–B: North-facing view of quartz vein striking north-south up the outcrop. C: Large pan (10.9 m wide) on top of outcrop, facing east. Note the outflow channel leading downslope. D: Deep channels created by water flow – these link many pans together in an interconnected system. E–F: Pans linked by channels. Note the higher elevation of some pans that, when full, would overflow into the channels directing water to lower lying structures. G: South-east facing view of large pan that overflows downslope to gnammas (behind small circular patch of shrubs in top right of photo). This pan receives runoff from upper slopes and channel systems. H: North-west facing view of the two gnammas that receive runoff from upslope pan and channel system (which continues to flow after rain has ceased), as well as the surrounding slope. Note that the well-defined flow of water does not reach the third gnamma, which lies to the left of those pictured. I: South-east facing view of all three gnammas – the black staining indicating inflow channels to the left and central gnamma. The smallest gnamma (right) receives runoff only from its immediate surrounds. Largest gnamma (left) has a diameter of 1.43 m. 203 10.2.1 Water Availability There is an abundance of pans at Anderson Rocks, many of which are quite expansive and, when full, resemble small pools. These pans are interconnected system linked by deeply scoured channels created by water flow over millennia – these channels distribute runoff and spill-over between pans (Figure 10.1c–f). Nevertheless, the shallow depth of these pans (mostly < 100 mm deep) limit the longevity of water and, like the modelled runoff pans these would generally be dry by October (Figure 7.8). The single deeper pan (160 mm) holds water for slightly longer. As elsewhere in the Granite Landscape Division, soils may store water for a month or so beyond winter, but the precise longevity depends on the depth of the profile and the amount of runoff received. Three gnammas are present at Anderson Rocks; a few others are forming elsewhere on the rock but are currently too small to be considered as water sources. All three gnammas receive additional input through runoff, but the situation is further complicated for the larger two gnammas, which are part of the interconnected flow system noted above. In this case, runoff is not simply generated by a single slope or an overlying pan, but rather a system of slopes, pans and channels that all eventually lead to the gnammas (Figure 10.1g–i). The gradient of channels varies, and the storage capacity of individual pans dictates when they will overflow, releasing water into the system further down the outcrop. It is well outside the scope of this thesis (and the expertise of its author) to reconstruct the incredibly complex system of inflow and runoff, but it is important to note that the closest analogue – high-input gnammas (Table 10.1) – is an oversimplification and will overestimate the water level at some times, and underestimate it at others. The gnammas’ conical structures provide less protection from direct sunlight, so an evaporation coefficient of 0.65 was applied. When full, the Anderson Rocks gnammas hold 593 litres of water (Table 10.1). Due to the lower volume combined with increased runoff, the gnammas approach or reach maximum capacity during winter, under all modelled rainfall conditions. Gnammas reach their capacity earlier as the quantity of winter rainfall increases, but the post-winter longevity is dictated by the amount of spring rainfall received (Figure 10.2). The influence of strong post-winter falls is clearly evident from the 2009, 2011 204 and 2012 model periods, where more water was available at 1 November than after high winter rainfall. Isolated falls had a strong, albeit temporary, effect on water balances due to the considerable runoff received by two of the three gnammas. For example, the 11.6 mm fall on 13 December 2011 deposited over 150 L of water across the three structures, but this balance reduced fairly rapidly due to greater evaporative demand (Figure 10.2). Table 10.1 Dimensions, capacity, shape, evaporation coefficient (EC) and modelled runoff conditions for the three gnammas at Anderson Rocks. Depth and radius were measured for each structure and used to calculate capacity following the standard mathematical formula for the volume of a cone (see Figure 9.4). Structure Radius (mm) Depth (mm) EC Runoff Shape Capacity (L) Gnamma 1 420 222 0.65 2x Conical 41 Gnamma 2 605 401 0.65 5x Conical 153.7 Gnamma 3 715 744 0.65 5x Conical 398.3 TOTAL 593 10.3 SURVEY AND EXCAVATION METHODS In 2016, the entire south-eastern portion of the Anderson Rocks outcrop was searched for potential cultural material. Border vegetation was generally too thick to penetrate, so survey of the surrounding area was restricted to less densely vegetated pockets. The western portion of the outcrop was omitted as there was no vehicular access and thick vegetation separated the eastern and western outcroppings at ground level (Figure 10.3). Three low-density accumulations of surface material were found – Scatter 1 on a vegetated shoulder of the outcrop, Scatter 2 at the base of a small pan (that also contained a quartz vein) and Scatter 3 in an open vegetation formation at the base of the outcrop (Figures 10.3–10.4). These scatters were each collected either as a single unit or as a series of smaller collections that were later merged into distinct collections based on their location (see Chapter 9.2.2). A few isolated artefacts were found at other parts of the site and were combined into a single unit for analysis. 205 high winter rainfall 700.00 1992 1998 600.00 water volume (L) 500.00 400.00 300.00 200.00 100.00 1 Dec 1 Nov 1 Oct 1 Sep 1 Aug 1 Jul 1 Jun 1 May 1 Apr 1 Mar 1 Feb 1 Jan 0.00 average winter rainfall 700.00 2009 2011 600.00 water volume (L) 500.00 400.00 300.00 200.00 100.00 1 Dec 1 Nov 1 Oct 1 Sep 1 Aug 1 Jul 1 Jun 1 May 1 Apr 1 Mar 1 Feb 1 Jan 0.00 low winter rainfall 700.00 2010 2012 600.00 water volume (L) 500.00 400.00 300.00 200.00 100.00 1 Dec 1 Nov 1 Oct 1 Sep 1 Aug 1 Jul 1 Jun 1 May 1 Apr 1 Mar 1 Feb 1 Jan 0.00 Figure 10.2 Total volume of water stored within Anderson Rocks gnammas during high (top), average (middle) and low (bottom) winter rainfall years. 206 A B Figure 10.3 A: Aerial photograph of Anderson Rocks. White square indicates the area magnified in image B. B: Main surveyed area of Anderson Rocks, showing the location of gnammas, artefact scatters and pits. Note the thicker vegetation separating the eastern and western exposures. Features too small to outline at the scale of this figure have been indicated by arrows. Base imagery from Google Earth. Three 500 x 500 mm pits were excavated in 2016, following the standard methods outlined earlier, including the use of arbitrary excavation units (XUs) due to the lack of visible stratigraphy (see Chapter 9.2.2). AR1 was positioned within Scatter 1, in an area where vegetation was limited to low groundcover. The sediment was waterlogged due to recent rains and became muddier as excavation progressed. At 360 mm below the surface (the base of XU13) water began intruding into the pit, making continued excavation impossible. Large rock slabs were also present, indicating that bedrock was probably nearby, so the pit was terminated and backfilled. AR2 and AR3 were excavated within Scatter 3, approximately 40 m southeast of the gnammas in an open area surrounded by thicker scrub/heath vegetation (Figures 10.3, 10.4d). Small roots and insect burrows were encountered within each pit – these intrusions may have caused some vertical movement of very small pieces, but there was no evidence of major disturbances. AR2 was excavated to a depth of 600 mm (XU24); cultural material was still present, but time constraints forced termination and backfill of the pit. The author returned a few months later to excavate AR3, which was positioned as near to AR2 as possible to intercept the same cultural 207 deposits while remaining out of the backfilled disturbance zone. XU33 was culturally sterile and time constraints were again a factor, so the pit was terminated at -850 mm and backfilled. Due to their proximity, AR2 and AR3 are analysed as a single unit, referred to hereafter as pit AR2-3. XU depths were identical throughout AR2 and AR3, except in one instance where two AR2 XUs (18–19) had to be combined to accommodate a 50 mm XU in AR3. XU numbering follows AR3 since that was the deeper deposit. A C B D Figure 10.4 A: Vegetated area near Scatter 1 and pit AR1, facing north-east. Note that the scatter and pit were immediately downslope of the thick vegetation pictured, where only low ground cover was present. B: Scatter 2, located within a small pan bisected by a quartz vein, facing north. Pan measures 1.63 m long by 0.91 m wide (maximum dimensions). C: close-up of the quartz vein in Scatter 2. D: Location of Scatter 3 and pit AR2-3 within open vegetated area southeast of the gnammas, facing south-west. Note the granite slab marks the position of the backfilled pit. All photographs by A.M. Rossi. 208 10.4 ORGANICS AND DATING Charcoal was far more common in pit AR1, and this pit yielded more than three times as much charcoal (n=87.7 g) as AR2-3, despite the latter being a larger and deeper pit. Soil pH values indicate that this discrepancy may be due to a preservation bias. As noted elsewhere (see Chapter 9.4.2), alkaline environments (pH > 7) are less conducive to charcoal preservation. The sediment in AR1 averages pH of 5 while AR2-3 is much closer to, and even within, the alkaline end of the scale (Table 10.2). Despite those same alkaline soil conditions being more favourable for bone preservation, none was found. Within each pit, charcoal weight shows the normal decline with depth, indicating the more recent material had been better preserved (Figure 10.5). In AR1, charcoal was most abundant in XU4 and XU5, which yielded 15.65 and 20.8 g, respectively. Over half of the XUs preserved more than 5 g of charcoal. In contrast, just six XUs in AR2-3 had more than one gram of charcoal. Charcoal weight peaked in XU9 that yielded 2.65 g. Table 10.2 Soil pH values for selected XUs within AR1 and AR2-3. Pit XU pH AR1 1 5 5 5 10 5 13 5 1 6 5 6 10 6 15 6.5 20 7 25 6 30 6 33 6 AR2-3 209 weight (g) 0 5 10 weight (g) 15 20 25 0 AR1 - XU 1 2 3 4 5 6 7 8 9 10 11 12 AR2-3 - XU 13 1 2 3 4 5 6 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 Figure 10.5 Charcoal weight with depth in AR1 (left) and AR2-3 (right). Five samples from AR2-3 were dated, but all samples originated from the deeper pit, AR3. All were AMS dated due to small sample sizes, and only two in situ charcoal samples were large enough to permit radiometric dating (Table 10.3). The upper four dates are in sequence, with XU3 returning an age of 164 ± 20 BP, and XU25 5928 ± 20 BP. Interestingly, XU3 and XU10 are separated by at least 150 mm of vertical deposit, yet their ages differ by just 45 radiocarbon years. The XU3 sample was derived from a large chunk of charcoal in the north-eastern portion of the pit – this fragment continued into the pit wall and unexcavated deposit beyond. Charcoal was only removed and collected where it occurred within the AR3 pit boundary, so it is possible that a small fragment of the remainder became dislodged during excavation and was introduced into XU10. The lowest sample, from XU32, returned an age of 2471 ± 20 BP, and is out of sequence with all other dates (Table 10.3). It is possible that the small sample of charcoal had moved vertically down the pit, possibly due to the root and insect activity noted above. Those dates considered more reliable (Wk44677, Wk-44679 and Wk-44680) were used to construct a depth-age curve, which indicated that the lowest cultural material in AR2-3 dates to around 8200 cal. BP (Figure 10.6). Sedimentation rates were highest in the top half of the pit and slowed below XU20. 210 Table 10.3 Charcoal samples and their returned radiocarbon ages, from pit AR2-3. Dates are arranged by XU depth. Asterisk beside sample weight indicates in-situ sample – all others derived from sieve residue. Note that no 13C measurements were provided with results. XU Depth (mm) Sample number Weight (g) %Modern C14 age BP Calibrated age BP 3 50–75 Wk-44677 0.25* 98 ± 0.2 164 ± 20 266–0 (68.2%) 275–0 (95.4%) 10 225–250 Wk-44678 0.1 97.5 ± 0.2 207 ± 20 280–149 (68.2%) 290–74 (95.4%) 20 500–525 Wk-44679 0.15 62.8 ± 0.2 3734 ± 20 4083–3982 (68.2%) 4145–3926 (95.4%) 25 625–650 Wk-44680 0.1* 47.8 ± 0.1 5928 ± 20 6735–6671 (68.2%) 6780–6645 (95.4%) 32 800–825 Wk-44681 0.1 73.5 ± 0.2 2471 ± 20 2675–2362 (68.2%) 2699–2353 (95.4%) cal. BP 0 1000 2000 3000 4000 5000 6000 7000 8000 9000 0 50 100 150 200 250 300 depth (mm) 350 400 450 500 550 600 650 700 750 800 850 Figure 10.6 Depth-age curve for AR2-3. Solid line joins radiocarbon age determinations, while linear trendline (dotted line) was used to estimate the age of the lowest cultural material, from XU32 (-825 mm). 10.5 STONE ARTEFACT ASSEMBLAGE A total of 821 stone artefacts were recovered from Anderson Rocks. Most of these came from the three pits, particularly AR2-3 which contributed over 80% of artefacts (Table 10.4). The assemblage was heavily quartz dominated (96%), and most quartz 211 was of moderate quality (i.e. exhibiting some flaw planes or granularity in texture), but there were some examples of extremely poor and very fine-quality material. The non-quartz artefacts comprised seven silcrete, nine chalcedony and 16 BIF pieces. The results of each pit and surface collection are discussed below, beginning with the surface material. The focus is on the attributes integral to interpretations of each pit/area; full artefact data can be found in Appendix E. Table 10.4 Number of stone artefacts from Anderson Rocks pits and surface collections. Pit/Collection Quartz Silcrete BIF Chalcedony TOTAL AR1 95 3 2 0 100 AR2-3 647 3 14 8 672 Scatter 1 9 0 0 0 9 Scatter 2 9 0 0 0 9 Scatter 3 24 1 0 1 26 Other surface material 5 0 0 0 5 789 7 16 9 821 TOTAL 10.5.1 Surface Collections Just 49 artefacts were recovered from surface collections at Anderson Rocks, more than half from Scatter 3. Scatters 1 and 2 each yielded nine pieces, and five were recovered from other parts of the site. Most of the artefacts were made on quartz (n=47), and the two non-local materials (silcrete and chalcedony) occurred within Scatter 3. Similarly, all of the high-quality materials were present in Scatter 3, where they outnumbered artefacts made on lower quality stone. All artefact types were present at Scatter 3, while the remaining collections preserved more restricted assemblages (Table 10.5). Table 10.5 Artefact types in surface collections at Anderson Rocks. Figures in parentheses refer to high-quality materials. Coll. = collection. Sc. = Scatter. OSM = other surface material, from outside defined scatters. Core Small fragment Large fragment Small flake Large flake Retouch Manuport TOTAL Sc.1 2 1 1 0 3 1 (1) 0 8 (1) Sc.2 1 0 4 0 4 0 0 9 Sc.3 1 (2) (2) 3 (1) 5 (9) (2) 1 10 (16) OSM 1 0 2 0 1 1 0 5 TOTAL 5 1 (2) 7 (2) 3 (1) 13 (9) 2 (3) 1 32 (17) Coll. 212 Cores Only five cores were present, all quartz of moderate–good quality. Two were found at Scatter 1, and one each in all other surface collections (Table 10.5). The largest core weighed 46.65 g, was made on the poorest quality material, and was found outside the defined scatters. Elsewhere, cores weighed no more than 5.3 g, and were of much finer textured quartz. Each core exhibited between five and nine negative flake scars, 0.38–5.18 g/scar, and the maximum values occurred on the largest core; all cores had less than 25% of their surface unmodified. Between two and four platforms were evident on each specimen, with the maximum recorded on three of the five cores. Bipolar cores were exclusively found in Scatter 1 and the surrounding material, while Scatters 2 and 3 yielded only freehand cores (Appendix E.1). Flaked Fragments Only three small fragments were present, unsurprising due to the bias against small items being noticed on the ground surface. All were quartz, and one was found in Scatter 1 and the remainder in Scatter 3. The Scatter 1 fragment was poorer quality quartz (quite granular in nature) while those from Scatter 3 were on higher quality material. Nine large fragments were found on the surface, all quartz. Four were recovered from Scatter 2, including three on very poor-quality raw material. All other collections yielded one or two examples, and high-quality material was restricted to Scatter 3. One notable example was found in the 'other surface material' – this piece was fairly rounded compared to other fragments, and the ridges between negative flake scars also appear to have been rounded. It possibly represents a core or fragment subsequently ingested by a bird (as a gizzard stone), which may explain its occurrence outside a defined scatter. As the identification was contentious, it was classified as a fragment, following the methods described earlier (see Chapter 9.3.2). It also preserved seven negative flake scars, while all other large fragments had no more than three (Appendix E.3). At Scatter 3, artefacts averaged 2.5 scars, 0.15 g/scar and two platforms. Elsewhere, reduction intensity was lower, when the large potential core is excluded (avg. 2 scars, 1.19 g/scar, 1.6 platforms). 213 Flakes Only four small flakes were present, all quartz and all from Scatter 3; two were made on high-quality quartz (Appendix E.4). A total of 22 large flakes were present, comprising 21 quartz and a single chalcedony flake. Large flakes were particularly common at Scatter 3, where 14 were found, including the nine on high-quality materials. All other collections yielded between one and four examples, and three from Scatter 2 were made on very poor-quality quartz (Table 10.5). Reduction intensity was similar across higher and lower quality materials. The latter averaged 2.9 scars, 6.21 mm/scar and 1.8 platforms, while high-quality materials averaged 2.6 scars, 6.12 mm/scar and 2.2 platforms. Broken (n=5) and heavily edge damaged flakes (n=3) were all confined to Scatter 3 (Appendix E.5). Retouched Flakes Five retouched flakes were found on the surface; all were backed and made on quartz. Three high-quality quartz specimens were present, including both pieces from Scatter 3 and another from Scatter 1. None of the retouched artefacts were broken or had suffered heavy edge damage (Appendix E.6). Manuports A single manuport was present at Scatter 3. This piece was a very weathered chunk of silcrete, weighing 27.3g, with a distinctly rounded surface. The material was quite coarse and poor quality, so the absence of flaking seems unsurprising. It may have been a hammerstone or grindstone, but no pitting or battering was evident. However, the coarse-textured material would make such features difficult to identify. In any case, the piece was certainly transported to the site by humans. 10.5.2 AR1 AR1 yielded 100 stone artefacts, with all types present except cores (Table 10.6). Generally, artefacts became more numerous with depth until their peak in XU9 (n=18), but the decline in XU6–7 is noteworthy. The basal portion of the pit (XU10– 13) contributed comparatively few stone artefacts, with no single XU yielding more than six pieces. Just four non-quartz artefacts were present (three silcrete, one BIF) in AR1, between XU4 and XU9 (Table 10.6). 214 Table 10.6 Stone artefact types by XU in AR1. Figures in parentheses refer to high-quality raw materials. XU Core Small fragment Large fragment Small flake Large flake Retouch TOTAL 1 2 3 4 5 6 7 8 9 10 11 12 13 TOTAL 0 0 0 0 0 0 0 0 0 0 0 0 0 0 3 3 3 1 (3) 7 3 7 8 (1) 8 2 4 1 1 51 (4) 0 0 1 0 0 0 0 1 4 1 0 0 2 9 0 1 4 (1) 1 (1) 1 0 1 (1) 4 (1) (1) (1) (3) 0 12 (9) 1 (1) 0 3 0 0 0 (1) 1 0 0 (1) 0 5 (3) 0 0 1 1 (1) 0 0 2 0 (1) (1) 0 0 4 (3) 4 4 (1) 9 5 (4) 8 (2) 4 7 12 (3) 17 (1) 3 (2) 4 (2) 1 (4) 3 81 (19) Flaked Fragments A total of 55 small fragments were present in AR1, comprising 54 quartz and one BIF. Unsurprisingly, many were poor-quality quartz, precluding their identification as either flakes, cores or retouched pieces; only four were made on high-quality materials. Numerically, small fragments were most abundant in XU7–9 (n=24) but contributed over 60% to each assemblage between XU5 and XU8 (Table 10.6). Just nine large fragments were present in AR1, all quartz of generally poor quality. These artefacts were most common in XU9 (n=4) but comprised a greater proportion of assemblages between XU9 and 13 (Table 10.6). Most weighed less than 0.6g, but one from XU3 weighed 10.45 g and was very poor-quality quartz. It may have been tested as a core and subsequently abandoned due to poor flaking properties; the presence/absence of a ventral surface could not be confirmed due to the material’s poor quality. Large fragments exhibited between one and three negative flake scars each, more commonly two or three, 0.29 g/scar (excluding the largest fragment) and one to two platforms. However, poor-quality raw material meant that scar and platform count could not be determined for nearly half of the specimens (Appendix E.3). 215 Flakes AR1 yielded 21 small flakes, which were most numerous in XU9 (n=5) and XU3 (n=4). They generally contributed between 10 and 30% to each assemblage, but a greater proportion were present in XU3 (> 40%) and XU12 (approx. 60%). Only nine small flakes were made on high-quality materials. These were concentrated in XU8– 12, and five were broken. Only eight large flakes were present in AR1, and five of these were from the upper four XUs (Table 10.6); five were on moderate quality quartz and silcrete, and three on high-quality quartz. Of the three high-quality flakes, one was broken and another exhibited heavy edge damage. These flakes averaged 2 negative flake scars, 6.31 mm/scar, and 1.5 platforms. The remaining flakes displayed 1–3 scars (avg. 2.2), 3.29–26.57 mm/scar (avg. 9.49 mm/scar) and 1–2 platforms (avg. 1.6) – none exhibited heavy edge damage, and only one was broken (Appendix E.5). Retouched Flakes Retouched flakes were the least common artefact type in AR1 – only seven were present, all quartz, and all backed. Three of these occurred between XU3 and XU5 and the remaining four between XU8 and XU11. Three were made on high-quality quartz, while the remainder were on moderate–good quality material. The latter were restricted to XU3–8, while high-quality materials were more common in the lower XUs (Table 10.6; Appendix E.6). Four retouched flakes were broken, including all three made on high-quality quartz. 10.5.3 AR2-3 A total of 672 stone artefacts were present in AR2-3, but the basal unit (XU33) was culturally sterile. Most artefacts were concentrated in XU12–23 (n=518), and only 34 were found in XU24–32 (Table 10.7). However, XU thickness and pit size varied with depth, so discard rates provide a more accurate indication of artefact concentrations (Figure 10.7). Cultural material accumulated fairly slowly in the earlier, basal deposits, with discard rates frequently below 5 artefacts/100 years. The highest rates occurred in XU17–20, around 2950–4035 cal. BP, and decreased until the European period (Figure 10.7). Artefacts were mostly quartz (n=647), but over half the 216 assemblage was made on high-quality materials (n=364) – these materials dominated in two artefact categories: small flakes and retouched pieces. High-quality materials outnumbered others in XU9–10, XU15–19 and XU29–32 (Table 10.7). BIF and chalcedony artefacts were particularly common in XU13–20 (n=14/22) but were also present elsewhere in the pit. The large artefact assemblage from AR2-3 means that change over time can be more easily addressed; as a result, some of the data are presented in greater detail. Table 10.7 Stone artefact types by XU in AR2-3. Figures in parentheses refer to high-quality raw materials. XU Core 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 TOTAL 0 0 0 0 0 1 0 (1) (1) 0 0 0 1 0 (1) 2 2 2 (1) 1 (1) 0 0 0 0 0 0 0 0 0 0 0 9 (5) Small fragment (1) (1) 1 0 (1) 5 (2) 1 0 1 (3) 1 2 1 (2) 0 5 2 (1) (2) 16 (9) 4 (1) 2 (3) 2 (3) 0 1 1 0 1 (1) 0 (1) 0 46 (31) Large fragment 0 1 0 0 1 1 (1) 0 0 0 1 0 0 2 0 0 (1) 3 6 1 0 0 2 0 1 0 0 0 0 0 0 0 19 (2) Small flake 0 1 (1) 4 (3) (3) 3 (6) 7 (1) 9 (1) 4 (4) (10) (8) (7) 7 (7) 2 (10) 10 (10) 12 (20) 8 (9) 6 (26) 16 (56) 3 (33) 10 (20) 15 (11) 5 (17) 10 (1) 1 5 1 (1) (2) 1 (1) (1) (3) (1) (2) 140 (275) Large flake 0 0 1 3 (1) 5 (3) 5 1 0 0 1 (1) 4 (1) 4 (1) 4 (2) 3 (1) 1 (5) 5 (1) 14 (5) 3 (5) 13 (4) 6 (2) 0 3 3 1 0 0 0 0 2 (1) (1) (1) 82 (35) Retouch TOTAL 0 0 0 0 0 0 0 0 0 (1) (1) 0 (1) 1 (1) 2 (1) (2) 4 (6) 1 (1) 2 (1) 1 (1) 0 1 0 0 0 0 0 0 0 0 0 12 (16) (1) 2 (2) 6 (3) 3 (3) 4 (8) 19 (6) 15 (2) 5 (5) 1 (11) (12) 3 (9) 13 (8) 8 (14) 17 (12) 20 (23) 15 (16) 13 (32) 55 (76) 17 (41) 29 (25) 22 (18) 7 (17) 16 (4) 4 8 2 (1) (2) 2 (1) (2) 2 (4) (3) (3) 308 (364) 217 no. artefacts/0.5 m2/ 100 yrs 0.00 5.00 10.00 15.00 20.00 25.00 30.00 35.00 1 2 3 4 5 6 7 8 9 10 11 12 13 14 AR2-3 XU 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 Figure 10.7 Artefact discard rates in AR2-3. The rate has been adjusted to artefacts/0.5 m2/100 years to allow for the difference in excavated area in the lower deposits, as only pit AR3 continued below 600 mm. Cores A total of 14 cores were present, all quartz, including five on high-quality quartz (Table 10.7). Cores were concentrated in XU15–21 (n=10), where each XU had one to two specimens; none were found below XU22. Size and reduction intensity varied between the standard and high-quality quartz cores. The latter averaged 5.4 scars, 0.51 g/scar, and 3.4 platforms. Four of the five cores weighed < 2.5 g and were bipolar – reduction technique could not be determined for the heaviest core; only one core had more than 25% of its surface unmodified. While other quartz cores had a higher scar average (n=5.89), other reduction intensity values were lower (avg. 0.95 g/scar, 3.1 platforms). These cores were considerably heavier, and only two weighed 218 < 2.5 g, both from XU16. Three had more than 25% of their surface unmodified, and five were bipolar, concentrated in XU16–20 (Appendix E.1). Flaked Fragments A total of 77 small fragments were recovered from AR2-3, all quartz, including 31 on high-quality quartz (Table 10.7). Small fragments were most common between XU15 and XU23, where 50 occurred. Half of these were found in XU18, which yielded five times as many small fragments as most other XUs; only five were present in XU24– 32. Large fragments were comparatively rare in AR2-3 – only 21 were recovered, accounting for just 3.1% of the entire artefact assemblage. All were on quartz, generally of moderate quality, and only two high-quality specimens were present. Large fragments were concentrated between XU14 and XU23, where 15 of the 21 specimens occurred, and especially common in XU18–19 (n=9); none were found below XU25. All large fragments displayed 1–4 negative flake scars, most commonly 2–3, and averaged 0.45 g/scar. Over 70% were dual-platform fragments, while the remainder had no more than three platforms (Appendix E.3). Small Flakes Small flakes were the most common artefact type in AR2-3 and were present in most XUs. Of the 415 small flakes recovered – 402 of which were quartz – 287 were found between XU14 and XU22 (Table 10.7). Over 65% of small flakes were made on high-quality material, so standard-quality small flakes were the minority in most XUs (Table 10.7). Over 50% of the small flakes were broken; separate breakage data were not available for standard- and high-quality materials, but the proportion of broken flakes with depth follows a similar pattern to the proportion of high-quality materials (Figure 10.8). This may indicate that those flakes were more likely to be broken than those made on lower quality material. 219 % high-quality 20 40 60 % broken 80 100 0.00 0 4 4 8 8 12 12 16 AR2-3 XU AR2-3 XU 0 0 20.00 40.00 60.00 80.00 100.00 16 20 20 24 24 28 28 32 32 Figure 10.8 Proportion of small flakes on high-quality material (left), and breakage rate for all small flakes (right) from AR2-3. Large flakes A total of 117 large flakes were recovered from AR2-3, including eight on non-local raw materials. Large flakes were most common between XU12 and XU21, where over 70% of the assemblage was found (n=84). They were especially concentrated in XU18–20, but were proportionally more significant in XU20, where they comprised over 30% of the assemblage. The distribution of high-quality materials (n=35) largely followed this same pattern, as they represented < 30% of the large flake assemblage (Table 10.7). High-quality materials are discussed separately to lower quality materials, below. Flakes on high-quality raw material displayed 1–5 flake scars, 1–3 platforms, and averaged 5.55 mm/scar. However, there was some variation with depth. The highest scar averages occurred in XU14–15, but values were also fairly high in XU20 and XU31–32 (Figure 10.9a). Platform averages also peaked in XU15, while the lowest averages were concentrated in XU21 and XU30–31. Higher scar averages were generally accompanied by higher platform averages, but this was not always the case (i.e. XU14 – Figure 10.9a). Mm/scar values were generally higher (indicating 220 lower reduction intensity) in the upper deposits, but sample sizes were small in those XU. Generally, most XUs averaged < 6 mm/scar, and values were below 4 mm/scar in XU14–15 (Figure 10.9c). Five flakes had long, parallel narrow flake scars, indicating they may have derived from formal, standardised cores. Two were present in XU21, and one each in XU6, XU14 and XU18. Only three large flakes had suffered serious edge damage, but the breakage rate was considerably higher, over 65% of those flakes where breakage/lack of breakage could be confidently identified. In fact, all flakes were broken in nine of the 13 XU that contributed breakage data – these maximum breakage rates were concentrated between XU11 and XU17 (Figure 10.9d). number 0.5 1 1.5 2 2.5 number 3 3.5 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 4 0 4.5 No. NFS No. platforms AR2-3 XU AR2-3 XU 0 4 6 8 10 1.5 2 2.5 3 3.5 4 4.5 No. NFS No. platforms % broken 12 14 16 0 18 high-quality other AR2-3 XU AR2-3 XU 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 2 1 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 mm/scar 0 0.5 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 10 20 30 40 50 60 70 80 90 100 high-quality other Figure 10.9 Selected attributes recorded on large quartz flakes, by XU, in AR2-3. A–B (top): Average negative flake scar (blue bars) and platform counts (orange bars) for high-quality materials (A – left) and lower quality materials (B – right). C–D: average mm/scar values (C – left) and proportion of broken flakes (D – right) for high-quality materials (black bars) and lower quality materials (grey bars). 221 Among the lower quality materials, individual values were more varied. Each exhibited 0–7 scars and 1–4 platforms, but flakes were generally larger (avg. 6.73 mm/scar). Flake scar and platform averages showed similar patterns – both peaked in XU8 and remained fairly high in XU12–15 (Figure 10.9b). Ranges were highest in XU18 and XU20, which yielded the largest flake assemblages. Mm/scar values were consistently higher than in the high-quality assemblage, and averages reached 16.72 mm/scar. However, most values were < 8 mm/scar; they were lowest in XU8, and peaked in XU16–17, and XU23 (Figure 10.9c). Two flakes from XU20 had a patina, possibly formed by weathering, on their dorsal surface. This coating covered some negative flake scars (distinguishing it from cortex or unmodified surface), while other flake removals had exposed fresher material beneath. The parent core may have been exposed for a considerable amount of time between flaking episodes. All unmodified surface categories were represented, and 12 flakes had more than 25% of their surface unmodified; six of these flakes came from XU20. A total of 32 flakes were bipolar – the technique was especially dominant in XU7 and XU12–14, but was also fairly common in XU18 and XU20, where sample sizes were larger. Ten flakes had suffered severe edge damage, but breakage rates were fairly low. Of the 18 XU that contributed data, seven yielded no broken flakes, and most had less than 50% of their flakes broken. Breakage rates were consistently low from XU18–21, and only whole flakes were present from XU23–25 (Figure 10.9d). Retouched Flakes A total of 28 retouched pieces were present in AR2-3, of which 16 were made on high-quality materials, including three on chalcedony. All retouched artefacts were found between XU10 and XU23, and over half (n=15) were from XU18–20. Ten were present in XU18, with no more than three in any other XU. Proportionally, however, retouched items were the most significant in XU16 where they comprised nearly 10% of the assemblage. Lower quality materials were particularly common in XU16–20 (n=9; Table 10.7). The high-quality materials were predominantly backed (n=14). The remaining two pieces were unifacially retouched, but this retouch also served to blunt an edge; it was not steep enough to be classified as backing. Only one of the high-quality 222 pieces had suffered severe edge damage, but seven were broken, including five from XU18. Among the lower quality materials, eight artefacts were backed, while four exhibited unifacial retouch. Only four of these retouched pieces were broken, and another exhibited significant edge damage (Appendix E.6). 10.6 OTHER CULTURAL MATERIAL Aside from stone artefacts, ochre was the only other cultural material present at Anderson Rocks. It was restricted to AR2-3 where it was fairly scarce, with a total of just 0.75 g recovered. Three colours were represented in varying quantity: brownishyellow (0.4 g), red (0.3 g) and the scarcest, a brownish-red (< 0.05 g). The samples are too small for any real analysis, but the vertical distribution is noteworthy. Ochre was primarily concentrated in XU18–22, where all but 0.15 g was recovered. Only one other XU – XU8 – yielded any ochre (Table 10.8). Table 10.8 Location, weight and colour of ochre samples from AR2-3. Pit XU Depth (mm) Weight (g) Colour AR2-3 8 175–200 0.15 red AR2-3 18 425–475 0.05 brownish-red AR2-3 21 525–550 < 0.05 brownish-red AR2-3 22 550–575 0.15 red AR2-3 22 550–575 0.4 brownish-yellow 10.7 CONCLUSION The most long-lived water source at Anderson Rocks are the three gnammas, two of which are also fed by the system of interconnected channels and pans that filter runoff down the outcrop. These gnammas fill to capacity even after low winter rainfall, and the longevity of water is determined by spring rainfall, which offsets water-loss through evaporation. Isolated falls (> 10 mm) create a considerable, but temporary, increase in water availability. Three pits were excavated (two of which intersect the same cultural deposits) at Anderson Rocks and a number of surface collections made around the site, but artefacts were most common in the excavations. Charcoal was scarce in AR2-3 but fairly abundant in AR1, probably due to soil conditions – the latter pit was not dated due to a comparative lack of artefacts and a shallower 223 deposit. The five radiocarbon dates from AR2-3 are largely in sequence, with the lowest reliable sample dating XU25 to 5928 ± 20 BP, around 6780–6645 cal. BP. The depth-age curve indicates that cultural material began to accumulate around 8200 cal. BP. The stone artefact assemblage was heavily quartz-dominated, but only AR2-3 provided a stratified sample large enough to evaluate change over time. High-quality quartz was more common, especially among small flakes and retouched pieces. Small flakes and small fragments dominated the assemblage in the upper portion of AR2-3, with few cores or retouched pieces present; all artefacts had less than 25% of their surface unmodified. Most artefacts were discarded in the middle XUs (XU12– 23), and the bulk of non-quartz artefacts were also found here, along with most of the cores and retouched pieces. Artefact discard rates were highest, particularly in XU17–20. Reduction intensity often peaked in these same XU, as measured by negative flake scar and platform counts, and weight/length per scar. Broken and damaged artefacts were also more common than elsewhere in the pit. Unmodified surface values were more varied and reached their maximum levels in the middle XU, among the lower quality materials. Most of the ochre – which was confined to pit AR2-3, was also found within these XU. In the lower portion of the pit, artefacts were comparatively scarce – no cores were present, and few large fragments or retouched pieces were found. 224 CHAPTER 11 GIBB ROCK 11.1 INTRODUCTION This chapter focusses on the physical characteristics of Gibb Rock and the archaeological material derived therein. To begin, Gibb Rock and its environmental setting are outlined, with specific relation to the Landscape Divisions represented in the wider area and the availability of water at the site itself. The survey, surface collection and excavation methods are then detailed. Finally, the archaeological material is quantified and described, beginning with organic material before turning to the major archaeological content – the stone artefact assemblage – and the attribute data collected. 11.2 SITE DESCRIPTION Gibb Rock is approximately 40 km NNE of Hyden, in the central part of the study area where Thicket and Heath meet (Figure 9.1). The rock and its immediate surrounds fall within the Granite Landscape Division, but beyond this lies an extensive boundary of Thicket. Vegetation shifts to Mallee (with pockets of Woodland – see Figure 3.14) around 800 m from Gibb Rock on the western and southern sides. The outcrop itself peaks at 416 m ASL and covers an area approximately 0.68 km2. The granite is fairly coarse-grained, and large chunks of quartz were observed weathering from the parent rock in several locations. Gibb Rock is currently a water reserve and does not appear on the DPLH register. 11.2.1 Water Availability No gnammas are present at Gibb Rock, but shallow pans are common (Figure 11.1a) – however, stored water would be short-lived (see Chapter 7.3.4). At present, a portion of the outcrop has been fenced off and a wall erected to capture runoff and funnel it to storage tanks for agricultural use (Figure 11.1b). Previously, water would have flowed down the outcrop and been stored in the surrounding soils. These would have held water for longer than pans, but the specific water-holding period would be 225 defined by the depth of the profile. In any case, water would be available during and immediately following winter rains but would be scarce thereafter. A B Figure 11.1 A: Shallow pans on eastern side of Gibb Rock; B: Fenced off wall to capture runoff from the outcrop, facing west. Both photographs by A.M. Rossi. 11.3 SURVEY AND EXCAVATION METHODS Gibb Rock fieldwork was conducted in October 2016. Due to accessibility issues, it was not possible to search the nearly 4 km perimeter of Gibb Rock. Much of the rock itself was fenced off, as noted above, and no attempt was made to enter this area. Instead, survey focussed on the central eastern side of the outcrop, where vehicle tracks provided easy access (Figure 11.2). The area north of the carpark was searched, and no potential cultural material was noted, but thick vegetation impeded access and visibility (Figure 11.2). The southern portion was more easily accessible as thick vegetation was interspersed with more sparsely vegetated areas. Two such areas had potential surface artefact concentrations – the first (Scatter 1) where a pocket of soil and vegetation had developed on the lower portion of the outcrop itself, and the second (Scatter 2) in a sandy area at the base of the outcrop (Figures 11.2, 11.3a–b). 226 A B Figure 11.2 A: Aerial photograph of Gibb Rock. White rectangle indicates area magnified in image B. Note the extensive wall constructed to filter runoff to the large circular tank, near the car park on the eastern side of the outcrop. B: Location of Scatter 1 and Scatter 2 (within white boundaries), and pit GR1 (red arrow). Note the thicker vegetation north of the car park area. Base imagery from Google Earth. At Scatter 1, there were several areas where quartz was naturally weathering from the bedrock (Figure 11.3c). This quartz was quite variable in quality and very friable on the weathered surface but had a finer, closer texture beneath. If it was quarried, rather than collected after weathering from the rock, it may have been a source of usable material. The potential artefacts were irregularly distributed and comparatively sparse, so when a piece was encountered, a GPS reading was taken, and the piece collected along with any other potential cultural material within a twometre radius; these small collections were subsequently grouped due to low artefact density. Deposit was probed and proved shallow (approx. 150 mm) so there was no real potential for excavation at Scatter 1. 227 A B C Figure 11.3 A: Area around Scatter 1, facing south. B: Area around Scatter 2, facing east. C: quartz weathering from parent granite, in Scatter 1. All photographs by A.M. Rossi. 228 At Scatter 2, no quartz exposures were present, and there was a lower density of potential artefactual material, all concentrated within approximately 10 m of the base of the outcrop. Surface material was collected following the same methods for Scatter 1. Depth probing indicated deposit was at least 300 mm deep, so there was some potential for excavation. Therefore, a single 500 x 500 mm pit (GR1) was positioned within Scatter 2 near some of the more convincing potential artefacts. This pit was excavated following the standard methods noted earlier (see Chapter 9.2.2). The soil was coarse and sandy, making it difficult to maintain a consistent XU (excavation unit) depth, so XU thickness varied from 20 to 40 mm. As the pit progressed, medium sized roots began to intrude, and few potential artefacts were noted in situ or in the sieve. As a result, GR1 was terminated at a depth of 130 mm below the ground surface, and subsequently backfilled. All sieve residue from GR1, along with material obtained from surface collections, was returned to Perth for analysis. 11.4 ORGANICS AND DATING A total of 1.6 g of charcoal was present in GR1, with no more than 0.4 g in any single XU. Charcoal weight declined from XU1–4, before increasing again in the basal unit (Figure 11.4). Due to the limited depth of deposit and the dearth of archaeological content, no attempt was made to date material from GR1. As a result, no pH readings were taken, since the main aim of that test was to determine the impact of soil chemistry on the preservation of organics (see Chapter 9.4.2). weight (g) 0 0.1 0.2 0.3 0.4 0.5 1 GR1 - XU 2 3 4 5 Figure 11.4 Weight of charcoal with depth in pit GR1. 229 11.5 STONE ARTEFACT ASSEMBLAGE A total of 88 artefacts were recovered from Gibb Rock, comprising 87 quartz and one silcrete. None of the quartz was high quality, and most was extremely variable, coarse-grained, and had visible crystal aggregations (Figure 9.3: top left). Most artefacts came from surface collections that contributed nearly 70% of artefacts (Table 11.1). The results of the pit (GR1) and surface collections are discussed separately below but, in most cases, small sample sizes were a considerable constraint on attribute analysis. Raw data are available in Appendix F. Table 11.1 Number of stone artefacts from the Gibb Rock pit and surface collections. Pit/Collection Quartz Silcrete TOTAL Scatter 1 53 0 53 Scatter 2 6 1 7 GR1 28 0 28 TOTAL 87 1 88 11.5.1 Surface Collections Of the 60 surface artefacts found at Gibb Rock, all but seven came from Scatter 1. However, Scatter 2 yielded the only non-quartz piece found at Gibb Rock (Table 11.1). Neither scatter preserved any retouched flakes, and Scatter 2 also had no flaked fragments; small flakes and small fragments were rare throughout (Table 11.2). Table 11.2 Frequency of artefact types found at Scatter 1 and Scatter 2. Collection Core Small fragment Large fragment Small flake Large flake Retouch TOTAL Scatter 1 2 1 21 3 26 0 53 Scatter 2 4 0 0 1 2 0 7 TOTAL 6 1 21 4 28 0 60 Cores Six cores were present, comprising five quartz and one silcrete. Despite lower overall artefact numbers, four of the cores came from Scatter 2, where cores comprised 57% of the artefact assemblage. Quartz cores were much heavier at 230 Scatter 1, ranging from 62.1 to 143g – at Scatter 2, however, all weighed under 12g, averaging 9.28g. The smallest Scatter 2 core was the only bipolar specimen. At Scatter 1, quartz cores averaged three negative flake scars, 43.5 g/scar and two platforms, while those at Scatter 2 averaged eight scars, 1.48 g/scar and four platforms. All quartz cores at Scatter 2 had < 25% unmodified surface, while those from Scatter 1 exhibited much higher values (50–75%, > 75%). The single silcrete core – from Scatter 2 – weighed 232g and had a single negative flake scar and platform, with > 75% of its surface unmodified (Appendix F.1). This piece may be a broken hammerstone but no pitting or battering was evident. However, the extremely rough texture may mask such features. Flaked Fragments All flaked fragments came from Scatter 1 and comprised one small and 21 large fragments (Table 11.2). The increased sample size in the latter meant that large fragments represented the best source of reduction intensity data from the surface collections, but only 14 provided negative flake scar and platform data. On average, each exhibited two negative flake scars, 1.84 g/scar and 1.5 platforms. One particular specimen, which had a smooth, fine texture compared to most Gibb Rock quartz (but still not high quality), exhibited seven scars and three platforms; all others had no more than three flake scars and one to two platforms (Appendix F.3). Flakes Small flakes were scarce on the surface, comprising three from Scatter 1 and a single example from Scatter 2. In contrast, a total of 28 large flakes were present, all but two from Scatter 1. None exhibited edge damage and just one broken flake was recovered, from Scatter 1. Bipolar and freehand reduction were equally common at Scatter 2, but bipolar flakes outnumbered freehand 2.6:1 at Scatter 1. At Scatter 1, flakes averaged 2.1 negative flake scars, 8.46 mm/scar and 1.6 platforms, but individual specimens displayed between zero and five scars and up to three platforms. At Scatter 2, one flake had no negative flake scars and a single platform, while the other was a dual platform flake with three scars (Appendix F.5). Unmodified surface was variable – each scatter yielded pieces in both the lowest (< 25%) and highest categories (> 75%). All four categories were represented at Scatter 1, but over 70% fell within the lowest unmodified surface category (< 25% – n=17). 231 11.5.2 GR1 A total of 28 artefacts, all quartz, were present in GR1 – artefactual stone was most abundant in the uppermost XU, which contributed 50% of the artefacts. There was an overall decline with depth thereafter, and the lowest two XU contributed just three artefacts. The extremely small sample size means that data were scarce, as the most numerous categories (small flakes, small fragments) contribute no usable attribute data. Few large flakes were found in GR1, and no cores or retouched pieces were present (Table 11.3). Table 11.3 Frequency of artefact types by XU in GR1. 0 Small fragment 4 Large fragment 0 Small flake 8 Large flake 2 0 3 0 1 1 3 0 4 0 2 4 0 1 0 5 0 1 TOTAL 0 13 XU Core 1 2 Retouch TOTAL 0 14 0 5 0 0 6 1 0 0 2 0 0 0 0 1 0 12 3 0 28 Small fragments were the most numerous category (n=13), and all but two were found in the upper three XUs, where most of the cultural material was concentrated (Table 11.3). Small flakes were also common (n=12) and were most abundant in XU1 (n=8). Only three large flakes were found, all in XU1–2. This was the only class of artefacts found in GR1 for which detailed attributes were recorded; sample size is therefore very problematic and average values must be treated with caution. On average, flakes had 2.5 flake scars, 4.58 mm/scar and originated from dual-platform cores. There was no evidence for bipolar reduction, and unmodified surface was consistently < 25%, but poor-quality raw material meant that each attribute could only be measured for a single flake (Appendix F5). 11.6 CONCLUSION Gibb Rock would have provided fairly easy access to three Landscape Divisions, and the food resources found therein. Granite and the surrounding Thicket are the dominant zones, but Mallee areas occur less than 1 km from the site. Water would be preserved in Gibb Rock pans and soils during winter but would be scarce soon 232 after rains ceased. While archaeological survey was impeded by thick vegetation and general accessibility issues, cultural material was found in two main areas – on the rock itself (Scatter 1), and at the base of the outcrop (Scatter 2). Excavation in the latter area provided few artefacts and limited organic content, so no attempt was made to date the deposit. The Gibb Rock stone artefact assemblage was small, and dominated by poor-quality quartz, possibly quarried locally from the exposures on the outcrop. Most artefacts came from Scatter 1, which yielded all artefact types except retouched pieces that were absent from the site as a whole. Generally, reduction intensity was variable, but attribute values were most often concentrated toward the lower end of the scale. 233 CHAPTER 12 MULKA’S CAVE 12.1 INTRODUCTION This chapter focuses on Mulka's Cave by first considering the natural setting of the site and its resource base, particularly the longevity of water in the deeper rock structures. The extensive body of previous research is outlined, from the earliest sources mentioning the site to later archaeological and rock art studies. Field methods are then described for the collections made during the 2006 development work as well as subsequent investigations permitted under Section 16 of the Aboriginal Heritage Act 1972. The charcoal assemblage is quantified, focussing on the resultant radiocarbon dates, including those derived from earlier research (Rossi 2010). The stone artefact assemblage is then discussed, with each of the eight pits/surface collections treated separately. Finally, a brief analysis of ochre and flaked glass is presented. 12.2 SITE DESCRIPTION Mulka's Cave is located within the Humps Nature reserve, 18 km north east of Hyden, within the southern part of the study area. The site itself falls within the Granite Landscape Division and is surrounded by vast areas of Mallee. Heath is accessible around 6.5 km to the north west, while Woodland and Saline areas lay around 15 km south (Figure 9.1). The Humps outcrop reaches 420 m ASL and covers an area approximately 0.5 km2, but other smaller granite exposures also occur within the reserve. Mulka's Cave is a registered site (ID 5842) that comprises surface artefact concentrations, gnammas, and a constructed lizard habitat ('lizard trap'). However, it is undoubtedly best known for the decorated boulder that houses 452 pigmented motifs in a variety of colours (Figure 12.1; Gunn 2006). These cultural and natural features all occur on the eastern side of The Humps (Figure 12.2). 234 B A D C E F Figure 12.1 Cultural and natural features present at The Humps. All photographs by A.M. Rossi. A: the main entrance to Mulka's Cave, facing west. Note the raised platforms either side of the entrance, created when erosion removed the sediments below. B: Two paintings inside the entrance to Mulka's Cave, note the yellow handstencil inside the circular motif. Lefthand painting measures 127 x 36 cm, righthand painting 83 cm wide (Gunn 2004:26). C: Raised walkway outside Mulka's Cave, facing south. D: Red and white pigmented handstencils and handprints inside Mulka's Cave. E: View of raised walkway from the top of Mulka's Cave. F: Gnamma 1, measuring 1.4 m long and 0.85 m wide. Note the shadows cast by the gnamma walls, protecting the water from evaporation. 235 A B Figure 12.2 A: Aerial photograph of The Humps. White rectangle indicates the area magnified in image B. B: Location of natural and cultural features around the Humps, including Mulka's Cave itself, the Humps Scatter, Camping Area Scatter, lizard trap and the main water sources. Features too small to outline at the scale of this figure are indicated by arrows. Pit locations for Mulka's Cave and the Camping Area are illustrated in Figure 12.4 and Figure 12.6. Locations for cleft rockhole and Gnamma 5 are approximations, based on Figure 3 from Gunn (2006:22). Base imagery from Google Earth. 12.2.1 Water Availability As at most granite outcrops, several shallow pans are present, and soils at the base of the rock would store water for several weeks after winter rains ceased (Figure 7.8). The most significant water sources, however, are the deeper rock structures. Five gnammas are present, four on flat exposures of granite north of the decorated boulder and the fifth mid-slope on the western side of the outcrop (Figure 12.2). The distance from the sloping outcrop means that four gnammas receive little runoff – at best, from an overlying pan. Some small, flat slabs of granite were found near the gnammas, and may have been used as lids to limit water loss through evaporation. One further rock structure exists, a cleft rockhole that has developed on a fault line on The Humps outcrop (Figure 12.2) – this structure is fed by the considerable runoff generated by the rock. The rockhole is interesting as its depth exceeds that of most gnammas, yet the surface area is more than five times greater than the largest gnamma. It also has a v-shaped profile, making its shape analogous to an inverted triangular prism (Figure 9.4). In smaller, straight-sided gnammas, water is protected by shadows cast by the structure’s walls except when the sun is directly overhead; protection is even greater when the structures are not full to capacity. In contrast, the 236 cleft rockhole's large surface area and gently sloping sides means that protection is much more limited, so an evaporation coefficient of 0.8 was applied (Table 12.1). Table 12.1 Surface dimensions, capacity, shape, evaporation coefficient (EC) and modelled runoff conditions for the six gnammas at Mulka's Cave. Dimensions of five traditional gnammas after Webb and Rossi (2008), cleft rockhole from Gunn (2004, 2006). Volumetric capacities were calculated using standard mathematical formulae provided in Figure 9.4. Surface dimensions (mm) Depth (mm) EC Runoff Shape Capacity (L) Gnamma 1 Radius a = 700 Radius b = 425 1100 0.4 None elliptic cylinder 1028.09 Gnamma 2 Radius = 1000 1000 0.4 2x cylinder 3141.59 Gnamma 3 Radius a = 225 Radius b = 100 300 0.4 None elliptic cylinder 21.21 Gnamma 4 Radius = 100 200 0.4 None cylinder 6.28 Gnamma 5 Radius = 125 350 0.4 5x cylinder 17.18 Cleft rockhole Length = 8500 Width = 2000 800 0.8 5x triangular prism 6800.00 Structure TOTAL 11,014.35 The six rock structures have a combined capacity of just over 11,000 L, more than half of which derives from the cleft rockhole that holds 6800 L alone (Table 12.1). Regardless of rainfall conditions, capacity was never reached (Figure 12.3), because only the cleft rockhole and Gnamma 5 ever completely filled. Under high winter rainfall conditions, water volume peaked at 8000–8500 L, around 75% of total capacity; it was slightly lower after average winter rainfall at around 7600 L. The cleft rockhole completely filled under either rainfall condition, albeit at different rates, so the difference in overall volume reflects the quantities of water stored in the gnammas. It is important to distinguish between these sources since gnammas could be capped, to limit evaporation, while the cleft rockhole could not. Gnammas held 1250–1650 L after high winter rainfall, and around 900–950 L when winter rains were average. In all cases, water balance steadily decreased from October, unless strong spring falls arrested the decline. By 31 December, 1500–4000 L of water was available, of which between 165 and 740 L was stored in gnammas. Even after low winter rainfall, water volume peaked at around 6000–6800 L; the cleft rockhole did not completely fill, and gnammas always held less than 400 L. All structures took longer to accumulate water and began to lose it soon after winter. By 31 December, 237 gnammas were almost dry, and 650–1400 L was available in the cleft rockhole (Figure 12.3). high winter rainfall 9000.00 1992 1998 8000.00 7000.00 water volume (L) 6000.00 5000.00 4000.00 3000.00 2000.00 1000.00 average winter rainfall 9000.00 1 Dec 1 Nov 1 Oct 1 Sep 1 Aug 1 Jul 1 Jun 1 May 1 Apr 1 Mar 1 Feb 1 Jan 0.00 2009 2011 8000.00 7000.00 water volume (L) 6000.00 5000.00 4000.00 3000.00 2000.00 1000.00 1 Dec 1 Nov 1 Oct 1 Sep 1 Aug 1 Jul 1 Jun 1 May 1 Apr 1 Mar 1 Feb 1 Jan 0.00 low winter rainfall 8000.00 2010 2012 7000.00 6000.00 water volume (L) 5000.00 4000.00 3000.00 2000.00 1000.00 1 Dec 1 Nov 1 Oct 1 Sep 1 Aug 1 Jul 1 Jun 1 May 1 Apr 1 Mar 1 Feb 1 Jan 0.00 Figure 12.3 Total volume of water available at Mulka's Cave from the gnammas and cleft rockhole (solid line) and the gnammas only (dashed line), during high (top), average (middle) and low (bottom) winter rainfall years. 238 12.3 PREVIOUS RESEARCH The earliest published accounts of Mulka’s Cave date to the early 1950s (Davidson 1952; Day 1951; Serventy 1952; Willey 1950), but Davidson conducted his fieldwork much earlier, in 1938–1939. Willey (1950) and Day (1951) both published short newspaper articles that briefly describe the decorated boulder but these documents mainly focussed on their personal experiences at the site. These records are interesting from a historical perspective but are not relevant to this research. Serventy (1952) and Davidson (1952) both attempted more serious publications, mostly concerning the rock art, but both have been superseded by the later, far more comprehensive rock art recording undertaken by Gunn (2004, 2006). These early sources are not discussed further as summaries are available elsewhere (Rossi 2010:32–37). The first archaeological investigation was conducted in 1988, when Bowdler et al. (1989) excavated a 1 m2 pit inside the dripline of the main entrance to Mulka’s Cave. A total of 210 stone artefacts were identified, the lowest from XU13a that also yielded a radiocarbon date of 420 ± 50 BP (lab code not supplied); they believed XU14–17 (0.7–1.1m below the surface) to be archaeologically sterile. The artefacts were largely quartz, and comprised 139 fragments, 61 flakes, seven retouched pieces and three cores. The upper deposits (XU1–10a) were heavily disturbed, with glass and modern material found throughout. Only 31 artefacts were recovered from XU11–13a, which they believed to have some stratigraphic integrity (despite the presence of a filled drainage channel running through these strata), leading Bowdler et al. (1989:31) to conclude that Mulka’s Cave ‘was occupied, in a brief and intermittent fashion, by Aboriginal people since perhaps 500 years ago’. As part of his site management plan (Gunn 2003), Gunn (2004, 2006) conducted a thorough re-recording of the rock art at Mulka’s Cave. Echoing Davidson’s (1952) earlier remarks, Gunn also concluded that the rock art assemblage was noteworthy for the region, with regard to the range of colours and motif types, as well as the overall number of motifs. However, Gunn’s analysis revealed that the degree of discordance between this site and those in the surrounding area was even more marked than Davidson (1952) perceived. Gunn recorded a total of 452 motifs, while most Noongar sites yield fewer than 100 (Webb and Gunn 2004). The assemblage 239 comprised 275 handstencils, 40 sprayed areas, 23 handprints, 23 paintings, three drawings and an object stencil, in addition to many uninterpretable fragments. The designs were executed in red, orange, yellow, cream and white pigment. Despite their well-protected location, nearly 80% of motifs were in poor or very-poor condition, leading Gunn to conclude that the visible artwork began to be created between 3000–2000 years ago. Motif condition and superimposition indicated that the art was not created in a single episode, meaning the site had been repeatedly visited over the last few thousand years (Gunn 2004, 2006). While Gunn’s main focus was documenting the artwork, he also conducted an onsite analysis of 130 stone artefacts from the drainage channel outside the main entrance to Mulka’s Cave. The assemblage comprised 96% quartz, most of which was debris or amorphous flaked material that could not be further classified (n=110), with only 16 flakes and four retouched pieces recovered. Gunn (2004, 2006) concluded that Mulka’s Cave was a recurrent focus for occupation over the last 3000 years and was used for both ritual and general activities. He argued that people would primarily have utilised the space outside the main entrance to the boulder, where the surface artefacts were found. Rossi’s (2010, 2014a) reinvestigation of the site was prompted by the conflicting interpretations offered by Bowdler et al. (1989) and Gunn (2004, 2006), specifically regarding the antiquity and intensity of occupation. Most of the earlier findings are largely irrelevant and superseded by the current research, so will not be discussed further. There are, however, some aspects that warrant discussion. Through comparison of photographs, plans and sketches created over several decades, it was demonstrated that heavy visitation since the 1980s resulted in the loss of 10–15 m3 of sediment from inside Mulka’s Cave, representing one vertical metre at the cave mouth (Rossi 2010; Rossi and Webb 2007, 2008; Webb and Rossi 2008). This sediment was redeposited onto the slope outside the main entrance, meaning the multitude of artefacts found in this area have no spatial or temporal control, and are of limited analytical value. Furthermore, Rossi (2010, 2014a) reanalysed the material excavated by Bowdler et al. (1989), including their large samples of 'natural debris'. While the assemblage data cannot be incorporated in the current analysis (due to the typological methodology employed at the time), Rossi (2010, 2014a) identified 240 nearly three times as many artefacts (n=604 compared to 210), including several from XU14–17, which Bowdler et al. (1989) considered archaeologically sterile. 12.4 SURVEY AND EXCAVATION METHODS All fieldwork at Mulka's Cave was conducted between 2006 and 2008 as part of Rossi's (2010) MA research. Work took place in two stages, that associated with the 2006 site development, and that permitted between 2007 and 2008 for research purposes. As per Gunn's (2003, 2004) recommendations, a raised walkway was installed outside the main entrance to Mulka's Cave in 2006, to prevent further degradation of the slope. Under Section 18 of the Aboriginal Heritage Act 1972, all ground disturbance had to be monitored for archaeological content. At this time, walkway postholes were excavated, and surface collections made in other areas of the site that would be impacted by development. In 2007, permission to excavate was sought and subsequently granted under S16 of the AHA, (permit no. 418). The one square-metre allowance was executed via the excavation of four 500 x 500 mm pits, allowing investigation in an undisturbed area around the cave mouth, as well as other areas of interest identified during development work. Much of the material collected during the 2006 site works were unusable, as erosion and/or earlier site works has voided spatial and temporal controls. The out of context material is not incorporated within the current study, but full assemblage details are available in Rossi (2010). The remaining pits and surface collections focus on three parts of the site, the Main Cave Area, the Camping Area, and the Humps outcrop itself. Each of these areas, and the specific methodologies employed are discussed below. 12.4.1 Main Cave Area The Main Cave Area was the focus of the 2006 work but, as noted above, most of the material is decontextualised and of no further use in the current analysis. Only one area suffered limited impact – the raised platforms outside the cave mouth, created when erosion scoured out the sediments below. These platforms receive no vehicular traffic and minimal foot traffic, noteworthy considering Mulka's Cave receives an estimated 80,000 visitors a year (Gunn 2006). Surface artefacts were 241 collected from these platforms during the 2006 development work. Artefact density was low, so each was collected as a single unit – these have been amalgamated for analysis (Entrance Platforms). The platform areas noted above also extended just inside the cave mouth, representing valuable interior deposits that had not been lost to erosion. As the 2006 development work progressed, the edges of these platforms were in danger of collapse, so they were sampled to evaluate the 420 BP date obtained by Bowdler et al. (1989). The author was not on-site at the time, but R.E. Webb cut a 200 x 200 mm column sample to a depth of -600 mm in 12, 50 mm XUs (Figure 12.4). Webb sorted the sieve residue on-site, and potential artefacts and dateable charcoal samples were passed on to the author for analysis – at present, the remaining charcoal samples cannot be located. Cultural material was comparatively scarce (and entirely lacking in the basal XU), probably due to the small pit size. Nevertheless, the assemblage is sufficient considering the main aim was to redate the interior deposits, while ensuring that archaeological content was present in dated XUs. After the Section 16 permit was granted, a final attempt was made to sample deposits in the Main Cave Area. In 2008, a single 500 x 500 mm pit (MC1) was excavated on a raised platform south of the cave entrance (Figure 12.4), following the standard procedures outlined earlier (Chapter 9.2.2). Archaeological content was limited, and MC1 was terminated after six XUs, at 300 mm below the surface. 242 Figure 12.4 Main Cave Area, facing west, showing the location of the Entrance Platforms (flat areas indicated by white arrows), Column Sample and pit MC1 (red arrows). Photograph: A.M. Rossi. 12.4.2 The Humps During the 2006 development work, most of the Humps outcrop was inspected for signs of Aboriginal usage. Artefacts were noted on one small shoulder of the outcrop that overlooked gnammas, where sediment and some vegetation (mostly ground cover with scattered low shrubs) had developed (Figure 12.5a–b). A low-density surface concentration was identified, spanning a 10 x 5 m area; in 2007, this scatter was collected as a single unit. Immediately thereafter, the area was probed to determine depth of deposit, which proved fairly consistent. A 500 x 500 mm pit (TH1) was therefore located on the flattest area in the centre of the scatter; it was terminated when bedrock was encountered at -320 mm, at the base of XU6. 243 A B C Figure 12.5 A: View of The Humps outcrop taken from the area around Gnammas 1 and 2, facing south west. Note the sparsely vegetated shoulder on which the Humps Scatter is located. B: The Humps Scatter, showing pit TH1, facing north east. Gnammas 1 and 2 occur on the flat granite exposure visible in the top right of the image. C: A portion of the Camping Area, facing west. Note the relatively clear area surrounded by thicker border vegetation – no artefacts were present in the latter area. 12.4.3 Camping Area The Camping area was identified in 2006 when members of the Mulka's Cave Aboriginal Steering Committee pointed out that as children they camped in a belt of trees between the gnammas and the decorated boulder, when brought to the site by their grandparents. The area was searched and a surface artefact scatter was located in a clearing bordered by larger shrubs (Figure 12.5c) – the more heavily vegetated area was also searched but yielded no artefacts. The Camping Area was not to be affected by the development works, so no further investigation occurred until the S16 permit was issued in 2007. At that time, the scatter was mapped, and 244 found to span an area measuring approximately 130 m long and 30–35 m wide. Within this scatter, six higher density concentrations were noted, each covering approximately 20 square metres (Figure 12.6: Clusters A–F). These clusters were collected individually, while the surrounding low-density material was collected as a single unit. Due to the extent of the scatter, two 500 x 500 mm pits were excavated within. CA1 was positioned within Cluster B and terminated at 450 mm below the surface (after XU9) at which point the sediment became cemented. CA2 was positioned at the centre of the scatter, between Cluster C and D. It was terminated after six XUs, at 300 mm below the surface, when saprolite was encountered. Figure 12.6 Camping Area artefact scatter, showing the extent of surface artefacts (light grey shading), the location of the high-density surface clusters (Clusters A–F, dashed circles) and pits CA1 and CA2. Note the scatter is surrounded by thick border vegetation (dark grey shading). 12.5 ORGANICS AND DATING Bowdler et al. (1989) recovered a considerable quantity of charcoal (over 1.2 kg) during their excavation inside Mulka's Cave. While the Column Sample total could not be quantified due to the loss of undated samples noted above, at least 5 g of charcoal was present, a not insubstantial quantity considering the small size of the pit. In contrast, none was present in MC1. A total of 3.75 g was found in TH1 and was fairly evenly distributed throughout the deposit, as four of the six XUs yielded 0.7–0.85g; the basal XU preserved just 0.2g (Figure 12.7). CA1 yielded 8.95 g of charcoal, mostly concentrated in XU1–2 (7.2 g). Charcoal was especially scarce towards the base of the pit – none was present in XU7, and XU8–9 preserved less than 0.05 g each (Figure 12.7). A similar amount was present in CA2, 8.8 g, despite 245 the pit being 150 mm shallower than CA1. Again, the bulk was from the upper deposits, with XU2 and XU3 yielding 2.8 and 3.75 g respectively. Charcoal quantity then declined to the base of the pit (Figure 12.7). charcoal weight (g) 0 1 2 3 4 5 1 2 CA1 - XU 3 4 5 6 7 8 charcoal weight (g) 0 0.5 1 1.5 2 2.5 3 3.5 4 1 CA2 - XU 2 3 4 5 6 charcoal weight (g) 0 0.2 0.4 0.6 0.8 1 1 TH1 - XU 2 3 4 5 6 Figure 12.7 Charcoal weight by XU within pits CA1 (top), CA2 (middle) and TH1 (bottom). Soil samples were unavailable for further analysis – these were passed to R.E. Webb during the author's MA research and could not be located thereafter. However, basic tests conducted at the time indicated that pH was generally < 5 (Rossi 2010:97). As noted elsewhere (see Chapter 9.4.2), such acid conditions are 246 conducive to charcoal preservation, so the lack of charcoal in MC1 cannot be explained in terms of soil pH. A total of 16 radiocarbon dates were obtained from Mulka's Cave, comprising 12 from the author's own excavations and four from the pit excavated by Bowdler et al. (1989); six were sourced as part of the author’s MA research (Rossi 2010). Dates were obtained from the Main Cave Area, Camping Area and The Humps. The dates from inside Mulka's Cave are complicated by the fact that they came from two separate pits in areas where the deposit was eroding at different rates. The top of each pit was first considered in reference to the 1952 ground level, as this represented the earliest visual record of the cave and the closest analogue to a depositional (vs. erosive) surface. This permitted Rossi (2014a) to correlate excavation depths across both pits allowing direct comparison of the dates from each (Figure 12.8; Table 12.2). All dates are in sequence except for Wk-35800 which is indistinguishable from modern carbon; it probably represents accidental inclusion of more recent material when the Column Sample was cut (see Rossi 2014a). The uppermost date (Wk-35798) lies within the disturbed deposits of the 1988 pit (XU1– 10a, see 12.3 above), so may be equally affected. The author had planned to avoid these deposits but had performed an incorrect correlation between the Column Sample and 1988 pit at the time of sample submission (Rossi 2014a). The dates from XU15 and XU16 of the 1988 pit, both of which pre-date 8000 cal. BP, overlap at one standard deviation (at 68.2% probability) and are statistically indistinguishable. When the in-sequence dates are considered, the depth-age curve indicates that cultural material began to accumulate around 9650 cal. BP (Figure 12.9). The earliest material derives from the 1988 pit, while that from the Column Sample postdates 6000 cal. BP. 247 Figure 12.8 Stratigraphic section (Bowdler et al. 1989: Figure 4.2a) superimposed with their XU levels (blue), sediment column (orange) and the approximate position of the 2006 Column Sample (pink). XUs numbered by each author. Rossi (2014a) after Bowdler et al. (1989). 248 Table 12.2 Radiocarbon dates from inside Mulka's Cave. Asterisk denotes AMS date. CS = Column Sample (cut in 2006), P88 = Bowdler et al. (1989) pit, excavated in 1988. Dates have been arranged by depth below the 2006 ground surface using Figure 12.8. Note that Bowdler et al. (1989) excavated by stratigraphic unit, so the thickness of each unit varied in different parts of the pit. The maximum depths have been cited below. All CS samples came from sieve residue, while the location of P88 samples are unknown. XU Pit Depth (mm) Sample no. Sample wt (g) C14 age BP Calibrated age BP Reference CS 3 100–150 Wk-35798* ~2 200 ± 25 281–145 (68.2%) 290–0 (95.4%) Rossi 2014a P88 12 125–300 Wk-29186 not available 1061 ± 31 961–914 (68.2%) 970–805 (95.4%) Rossi 2014a CS 7 300–350 Wk-35799* ~2 2118 ± 25 2082–2009 (68.2%) Rossi 2014a 2148–1940 (95.4%) P88 14 350–475 Wk-29187 not available 4285 ± 36 4858–4711 (68.2%) Rossi 2014a 4870–4626 (95.4%) P88 15 475–575 Wk-27113* ~2 7386 ± 30 8194–8050 (68.2%) Rossi 2010, 8299–8026 (95.4%) 2014a CS 11 500–550 Wk-35800* ~1 Modern P88 16 575–675 Wk-29188 not available Modern 7284 ± 41 8156–8002 (68.2%) Rossi 2014a 8166–7973 (95.4%) cal. BP cal. BP 0 Rossi 2014a 0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000 11000 1000 2000 3000 4000 5000 6000 7000 8000 0 0 CA1 MC CA2 50 100 TH1 100 200 150 300 500 depth (mm) depth (mm) 200 400 250 300 600 350 700 400 800 450 900 500 Figure 12.9 Depth-age curves for inside Mulka’s Cave (left) and other parts of the site (right). Solid lines join radiocarbon age determinations, while linear trendlines (dotted lines) were used to estimate the age of the lowest cultural material, represented by the end point of those lines. Inside Mulka’s Cave, Wk-27113 (P88 XU15) and Wk-29188 (P88 XU16) were statistically indistinguishable. The lower sample was selected as it was derived from a larger charcoal sample. 249 Two dates were sourced from the Humps pit (TH1), one basal (XU6) and the other from XU2. The basal sample returned an age of 1236 ± 20 BP, while XU2 was more recent, 454 ± 20 BP (Table 12.3). A total of seven dates were obtained for the Camping Area – four from CA1 and three from CA2. In CA1, all but the lowest date (XU8, 92 ± 30 BP) are in sequence. That date was on a single, small fragment of charcoal that may have been introduced into the pit during excavation. The lowest reliable sample, from XU6, dates to 1626 ± 21 BP (Table 12.3). The CA2 dates are all in sequence, ranging from 1340 ± 94 BP for XU2, to 5093 ± 30 BP for XU5. Depth-age curves indicate that artefacts began to accumulate in CA2 around 6700 cal. BP and from 2500 cal. BP in CA1. The vast differences in sedimentation rates is of note, as both pits are separated by < 60 m of flat, featureless terrain. Sediment in TH1 accumulated at a similar rate to CA1 (Figure 12.9). Table 12.3 Radiocarbon dates for the Humps and Camping Area pits. Asterisk denotes AMS date. All samples were collected from sieve residue. Pit TH1 CA1 CA2 XU Depth (mm) Sample no. 2 70–120 6 2 Sample wt (g) C14 age BP Wk-38220* 0.7 454 ± 20 504–472 (68.2%) 510–343 (95.4%) 300–320 Wk-38221* 0.2 1236 ± 20 1172–1066 (68.2%) 1180–1059 (95.4%) 3.9 326 ± 69 453–290 (68.2%) 500–147 (95.4%) Rossi 2010 Rossi 2010 40–90 Wk-24311 Calibrated age BP 5 200–250 Wk-24864* 0.3 1541 ± 30 1411–1349 (68.2%) 1470–1308 (95.4%) 6 250–300 Wk-38218* 0.25 1626 ± 21 1520–1432 (68.2%) 1533–1416 (95.4%) 8 350–400 Wk-24312* < 0.05 92 ± 30 2 50–100 2.8 4 160–220 Wk-38219* 5 220–270 Wk-24310* Wk-24309 Reference 239–24 (68.2%) 254–0 (95.4%) Rossi 2010 1340 ± 94 1300–1093 (68.2%) 1364–981 (95.4%) Rossi 2010 1.05 4308 ± 21 4855–4830 (68.2%) 4873–4711 (95.4%) 0.8 5093 ± 30 5891–5743 (68.2%) 5908 –5665 (95.4%) Rossi 2010 250 12.6 STONE ARTEFACT ASSEMBLAGE A total of 969 stone artefacts were recovered from Mulka's Cave, 702 of which derived from the Camping Area. In contrast, the Main Cave Area yielded just 90 artefacts. The assemblage was heavily quartz dominated, comprising over 96% by artefact number (Table 12.4). Most quartz was of fairly good quality, with fewer examples of the extremely poor material that was found elsewhere. Non-quartz materials consisted of silcrete (n=21), BIF (n=8) and chalcedony (n=3), all of which were confined to the Camping Area. The results of each pit and surface collection are briefly discussed below, arranged by site area and beginning with surface material. Due to the number of collections/pits (and the generally short depth of deposit), only selected data are tabulated or graphed – raw data are available in Appendix G. Table 12.4 Number of stone artefacts from Mulka’s Cave pits and surface collections. Assemblages have been grouped by site area. Main Cave Humps Camping Area Quartz Silcrete BIF Chalcedony TOTAL Entrance Platforms 8 0 0 0 8 Column Sample 63 0 0 0 63 MC1 19 0 0 0 19 Humps Scatter 21 0 0 0 21 TH1 156 0 0 0 156 CA Scatter 174 9 0 0 183 CA1 304 9 5 3 321 CA2 192 3 3 0 198 937 21 8 3 969 TOTAL 12.6.1 Entrance Platforms Just eight stone artefacts were recovered from the Entrance Platforms, comprising five large flakes, two large fragments and a core, all on lower quality quartz. The core was small (1.75 g), bipolar, and exhibited four scars detached from three platforms; < 25% of its surface was unmodified (Appendix G.1). The large fragments varied in weight from 0.9 to 4.25 g, but each exhibited a single scar and platform (Appendix G.3). Each large flake had 2–5 scars (avg. 3.4), 3.15–7.86 mm/scar (avg. 5.41) and 2–3 platforms (avg. 2.2). All had < 25% of their surface unmodified, and none were broken, but two were damaged (Appendix G.5) 251 12.6.2 Column Sample The 2006 Column Sample yielded 63 stone artefacts, all quartz, including 41 on high-quality materials. Small flakes dominated the assemblage (n=36), while cores and retouched pieces were rare; all cores, small fragments and retouched pieces were made on high-quality quartz. The limited pit size and small assemblage means that vertical trends should be treated with caution, so discard rates were not calculated. Overall, the number of stone artefacts increased with depth until the peak in XU9 (n=16) and declined thereafter; no artefacts were found in XU1, XU3 or XU12 (Table 12.5). While small sample size is problematic, the assemblage can be characterised in general terms. Table 12.5 Stone artefact types by XU from the Column Sample inside Mulka’s Cave. Figures in parentheses refer to high-quality quartz. 0 0 0 Small fragment 0 0 0 Large fragment 0 0 0 Small flake 0 1 (1) 0 Large flake 0 0 0 0 0 0 0 (1) 0 0 0 (1) 0 (1) 0 0 0 (2) 0 (2) (5) 0 1 (1) 1 1 0 (1) 0 0 3 (2) 3 2 4 (2) (6) (4) (5) (7) (1) 10 (26) 0 (1) 1 3 (1) 1 (1) 4 (3) 0 0 9 (6) XU Core 1 2 3 4 5 6 7 8 9 10 11 TOTAL Retouch TOTAL 0 0 0 0 1 (1) 0 0 0 0 0 0 (1) 0 0 (1) 3 3 (3) 6 (2) 4 (7) 1 (6) 4 (12) (7) (3) 22 (41) Just one core was found in the Column Sample, in XU8 (Table 12.5). It was made on high-quality material, a smooth-textured black opaque quartz. It weighed 3.25 g and exhibited six flake scars detached from four platforms. Less than 25% of the surface was unmodified, and a combination of feather and axial terminations indicates that both freehand and bipolar reduction techniques were used (Appendix G.1). A total of five small flaked fragments were present, all made on high-quality quartz. Four were found in XU9 and XU11, and the remaining small fragment was recovered from XU5. Large fragments were also comparatively rare; five were found in XU5–9, including two on high-quality quartz (Table 12.5). The high-quality material was more 252 intensively reduced, and averaged 4 flake scars, 0.18 g/scar and 2 platforms; the lower quality material averaged 2 scars, 0.5 g/scar and 1.5 platforms (Appendix G.3). Small flakes were the most common artefact type in the Column Sample, comprising more than 55% of the artefact assemblage (n=36). They were especially concentrated between XU6 and XU10, where 28 were found (Table 12.5). Lower quality materials were only present in XU2–6, while high-quality quartz was found throughout. None of the small flakes from XU2–6 were broken, and < 15% from XU10 were. Elsewhere, 60–100% of small flakes had been broken (Appendix G.4). A total of 15 large flakes were present, all between XU5 and XU9; numbers peaked in XU7 (n=4) and XU9 (n=7 – Table 12.5). Six flakes were made on high quality quartz, and averaged 2.3 negative flake scars, 5.05 mm/scar and 1.6 platforms. Individual values were highest in XU5–7, and one flake in XU5 exhibited several small partial flake scars, indicating it came from a highly reduced core. Only one flake was broken, and another had suffered heavy edge damage – both were found in XU9. The remaining nine flakes were made on lower quality material, and averaged 2.1 flake scars, 5.01 mm/scar and 1.8 platforms (Appendix G.5). Only one retouched piece – a backed artefact – was present, in XU9 (Table 12.5). The piece was made on high-quality quartz, was unbroken and exhibited no significant edge damage (Appendix G.6). 12.6.3 MC1 Just 19 stone artefacts were recovered from MC1, all quartz, including five on highquality quartz; all artefact types were represented except small fragments. XU1–2 were archaeologically sterile, and artefact numbers peaked in XU4 (n=7) and declined to the base of the pit (Table 12.6). The very small sample size means that the MC1 assemblage can only be characterised in very basic terms. Just one core was found, in XU4, made on a smoky clear quartz with a smooth, glassy surface (Table 12.6). This piece is noteworthy as it preserved four original crystal facets, representing the tip of a large individual crystal. The opposing surface was covered in negative flake scars and retained no trace of a ventral surface. This piece therefore represents the use of a large crystal as a core, or possibly a flake 253 that had been subsequently modified to the extent that the entire ventral surface was removed – had any trace of this surface remained, the piece would have been classified as retouched (see Chapter 9.3.2). The core was extremely small (0.65 g) so was presumably worked via the bipolar method. Three negative flake scars were preserved, all initiated from a single platform (Figure 12.10; Appendix G.1). Table 12.6 Stone artefact types by XU in MC1. Figures in parentheses refer to high-quality quartz. XU Core Small fragment Large fragment Small flake Large flake Retouch TOTAL 1 2 3 4 5 6 0 0 0 (1) 0 0 0 0 0 0 0 0 0 0 1 1 0 0 0 0 0 0 1 1 0 0 1 (1) 2 (1) 3 2 0 0 1 (1) 1 (1) 0 0 0 0 3 (2) 4 (3) 4 3 TOTAL (1) 0 2 2 8 (2) 2 (2) 14 (5) Figure 12.10 Small core from MC1 XU4. Note the presence of four original crystal facets (indicated by blue arrows). Two large fragments were present in MC1, one each in XU3 and XU4; both were on lower quality quartz (Table 12.6). The XU3 fragment was made on a greyish quartz of moderate texture and exhibited three scars detached from three platforms. In contrast, the XU4 specimen was made on extremely poor-quality quartz that comprised an aggregation of small crystals that gave it a very irregular texture. It weighed 8 g and exhibited just one negative flake scar (Appendix G.3). 254 Only two small flakes were found, in XU5–6. Both were made on lower quality quartz and had been broken (Appendix G.4). In contrast, large flakes were the most numerous artefact category, and comprised 10 of the 19 artefacts present; at least two large flakes were found in each XU (Table 12.6). Only two flakes were made on high-quality quartz, from XU3–4. They averaged three negative flake scars, 5.12 mm/scar and two platforms; one exhibited several overlapping small scars, indicating it derived from a highly reduced parent core. Neither flake was broken, but one had suffered severe edge damage. Eight flakes were made on lower quality material – reduction intensity was lower, averaging 2.1 flake scars, 12.28 mm/scar and two platforms. Only one of these flakes was broken, from XU6 (Appendix G.5). Only XU3 and XU4 both yielded one retouched piece made on high-quality quartz and another on lower quality material (Table 12.6). All were backed except one unifacially retouched piece from XU4. None of the retouched pieces were broken or exhibited major edge damage (Appendix G.6). 12.6.4 The Humps Scatter The Humps Scatter yielded 21 stone artefacts, all quartz, dominated by high-quality materials (n=14). The assemblage comprised nine large flakes, four small flakes, three cores, two large fragments, two retouched pieces and a small fragment. Two cores were made on high-quality quartz, while the remainder was fine-textured, but exhibited some flaws and fracture planes. All cores were small (2.2–3.05 g), bipolar, and preserved 6–7 flake scars detached from 2–3 platforms (Appendix G.1). None had more than 25% of their surface unmodified, and one high-quality quartz core preserved three original crystal facets, the largest used as the main platform. A single small fragment was found, made on high-quality quartz (Appendix G.2). Large fragments were almost as rare, with just two present, including one on lower quality material. The better quality piece was more intensively reduced (7 scars, 0.29 g/scar, 3 platforms) than the lower quality piece, but the latter still displayed five scars (0.43 g/scar) detached from two platforms (Appendix G.3). All four small flakes were unbroken; three were made on high-quality quartz while the fourth exhibited a blockier structure. Of the nine large flakes, only three were made on lower quality raw material. These averaged 4.7 flake scars, 4.96 mm/scar 255 and 2.3 platforms; unmodified surface was always < 25%. In contrast, the six highquality materials were less intensively reduced, averaging 2.8 scars, 6.74 mm/scar and 2.2 platforms; one flake had 25–50% of its surface unmodified. Three highquality flakes were broken, and none exhibited severe edge damage (Appendix G.5). Two retouched pieces were present, both exhibiting unifacial retouch in the form of flake removals on their ventral surfaces. One was made on glassy smooth quartz, while the other was a more granular variety. Neither was broken, and only the lower quality piece exhibited significant edge damage (Appendix G.6). 12.6.5 TH1 A total of 156 artefacts were excavated from TH1, all quartz, including just 24 on high-quality quartz (Table 12.7). Discard rates ranged from 7.6–21 artefacts/100 years, and values exceeded 12 artefacts/100 years in five of the six XU. The highest values occurred in XU2 and XU4, and the lowest in XU5 (Figure 12.11). Table 12.7 Stone artefact types by XU in TH1. Figures in parentheses refer to high-quality quartz. 0 0 0 0 0 Small fragment 4 (4) 8 (7) 10 13 6 Large fragment 3 3 4 6 3 Small flake 14 14 2 (4) 6 (3) 6 Large flake 3 (1) 3 2 1 3 0 0 3 44 (11) 0 19 2 (1) 44 (8) 0 12 (1) XU Core 1 2 3 4 5 6 TOTAL Retouch TOTAL 3 2 (3) 3 2 (1) 27 (5) 30 (7) 18 (7) 29 (3) 20 (1) 3 13 (4) 8 (1) 132 (24) no. artefacts/100 yrs 0.00 5.00 10.00 15.00 20.00 25.00 1 2 TH1 XU 3 4 5 6 Figure 12.11 Artefact discard rates for pit TH1. 256 Small fragments were the most common artefact type in TH1, with a total of 55 present. They were most numerous between XU2 and XU4, where 38 were found; just three were present in the basal XU, and high-quality materials were only found in XU1–2 (Table 12.7). A total of 19 large fragments were found in TH1, all on lower quality quartz. They were most common in XU4 (n=6), and none were present in XU6 (Table 12.7). Most fragments had 1–3 negative flake scars, and 1–2 platforms, while a single specimen from XU3 had five scars and three platforms. Reduction intensity peaked in XU1 (avg. 2.3 scars, 0.18 g/scar, 1.7 platforms) and showed a general decline with depth. Flake scar and platform averages were lowest in XU5 (avg. 1 scar, 1 platform), but those fragments were among the lightest (avg. 0.22 g/scar – Appendix G.3). Small flakes were the second most numerous artefact type in TH1, with a total of 52 recovered. XU1 and XU2 yielded 14 each, while all other XUs had three to nine specimens. High-quality materials were only found in XU3–4 and XU6 (Table 12.7). Large flakes were rare – only 13 were found, seven of which occurred in the upper two XU, including the single large flake made on high-quality material; no large flakes were present in XU6 (Table 12.7). The high-quality quartz flake was broken and displayed three flake scars (3.68 mm/scar) detached from a single platform. The flake scars were long, narrow and parallel, and the flake itself was fairly long and thin, so it may have derived from a more standardised core. The lower quality quartz flakes averaged 2.4 scars, 6.56 mm/scar and two platforms. Three had > 25% of their surface unmodified, including two from XU3. Reduction intensity was highest in XU1, where the single bipolar flake also occurred, and lowest in XU3 (Appendix G.5). A total of 17 retouched pieces were found in TH1, with two to three from each XU (Table 12.7). All were backed except two unifacially retouched pieces, from XU1 and XU4. The four specimens made on high-quality materials were restricted to XU3 (n=3) and XU5 (n=1). All of the high-quality quartz pieces were backed and only one, from XU3, was unbroken. The XU5 piece preserved traces of probable hafting resin, indicating it formed part of a composite tool. Nine of the 13 lower quality pieces were broken, including all pieces from XU5–6 (Appendix G.6). 257 12.6.6 Camping Area Scatter A total of 183 artefacts were recovered from the Camping Area Scatter, all quartz (including 50 on high-quality quartz) except for nine silcrete specimens. Artefacts were most numerous in the surrounding low-density material (n=70, including three silcrete), while the densest cluster, Cluster B, yielded 38. Clusters E and F were the sparsest, with just seven and six artefacts respectively. High-quality materials were proportionately most common in Clusters B and C (Table 12.8). Table 12.8 Stone artefact types in the Camping Area Scatter. Figures in parentheses refer to highquality materials. Note that ‘other CA’ refers to the low-density material found outside the clusters, but within the Camping Area Scatter. Area Core Small fragment Large fragment Small flake Large flake Retouch Manuport TOTAL Cluster A Cluster B Cluster C Cluster D Cluster E Cluster F Other CA TOTAL 0 4 (1) 0 1 (1) 0 1 1 7 (2) 0 0 (2) 0 0 0 1 1 (2) 3 2 (1) 3 2 1 1 11 (2) 23 (3) 3 (1) (3) (2) 1 (1) 0 (1) 3 (1) 7 (9) 10 (1) 19 (7) 10 (3) 11 (1) 3 (1) 3 36 (13) 92 (26) (4) (1) 0 1 1 (1) 0 (2) 2 (8) 1 0 0 0 0 0 0 1 17 (6) 25 (13) 13 (7) 16 (3) 5 (2) 5 (1) 52 (18) 133 (50) Cores Nine cores were present in the Camping Area Scatter, comprising seven quartz (two high quality) and two silcrete. Cores were especially common within Cluster B, which yielded five cores, including both silcrete specimens. No cores were present in Clusters A or E (Table 12.8). The two high-quality quartz cores were found in Clusters B and D; both weighed < 3g, were bipolar, and had been heavily reduced. They displayed ten flake scars each (0.24–0.27 g/scar), detached from 4–6 platforms. Reduction intensity was more varied among the lower quality materials. The two silcrete cores, both from Cluster B, were the largest, and weighed 14.15–60.2 g. The heavier was a single platform core with four negative flake scars and > 50–75% of its surface unmodified; the smaller displayed ten scars detached from four platforms and had < 25% unmodified surface. The latter was made on a finer grained silcrete, so the increased reduction intensity is not surprising. Lower quality quartz cores preserved 3–11 scars, 0.14–5.52 g/scar and 1–5 platforms; the lowest values were 258 all found in Cluster F, where the single core had > 75% of its surface unmodified. Three cores were bipolar, one each from Clusters B, D and F. The latter weighed 16.55 g so the flaking method was probably not related to core size, but may have been a response to the poor quality, blocky material (Appendix G.1). Flaked Fragments Only three small fragments were found, comprising two high-quality quartz pieces from Cluster C and one lower quality quartz piece from the surrounding area (Table 12.8). The 26 large fragments (all quartz) were primarily concentrated in the surrounding area, which yielded 13; all other Clusters had no more than three each. Only three large fragments were made on high-quality quartz, including two from the surrounding area and one from Cluster B. These fragments preserved 2–4 scars, 0.08–0.67 g/scar and two platforms. Reduction intensity was more varied among the lower quality quartz specimens. These exhibited 1–9 scars, 0.2–6.75 g/scar and 1–4 platforms. Reduction intensity values were greatest in Cluster E (6 scars, 3 platforms) and lowest in Cluster F (1 scar, 1 platform, 6.75 g/scar); Clusters B–D yielded values toward the higher end of the scale (Appendix G.3). One specimen from Cluster C preserved long, narrow, parallel flake scars, indicating it derived from a more standardised core. Flakes A total of 16 small flakes were found in the Camping Area Scatter, all quartz, with no more than four in any single collection; none were present in Cluster E. Small flakes on high-quality materials (n=9) outnumbered those on lower quality material (n=7), and Clusters B, C and F yielded only high-quality materials (Table 12.8). Both small flakes from Cluster C were broken, as well as three of four from Cluster A; unbroken flakes dominated elsewhere (Appendix G.4). Large flakes were the most common artefact type in the Camping Area Scatter, and a total of 118 were recovered, including six on silcrete. The largest assemblage (n=49) was present in the surrounding material, while large flakes were comparatively rare in Clusters E and F, where few artefacts were found. Only 26 large flakes were made on high-quality material – these were particularly common in Cluster B and the surrounding material (Table 12.8). These flakes exhibited 1–7 259 scars, 1–4 platforms and 1.68–11.66 mm/scar. All reduction intensity measures peaked in Cluster A, while Clusters D and E had the lowest values across one or more attributes (Figure 12.12a, c). Variation in individual values was greatest in Cluster B and the surrounding material, where sample sizes were largest. One flake from the surrounding material exhibited the same notching noted in the CA2 sample (see Chapter 12.6.8). Broken flakes were rare, and restricted to Cluster B (< 15%) and the surrounding material (< 17% – Figure 12.12d); just a single flake exhibited significant edge damage, from Cluster B (Appendix G.5). number 0 1 2 3 4 number 5 6 7 8 Cl. A 0 1 2 3 4 5 6 Cl. A No. NFS No. NFS No. platforms No. platforms Cl. B Cl. B Cl. C Cl. C Cl. D Cl. D Cl. E Cl. E Cl. F Cl. F Other CA Other CA mm/scar 0 Cl. A 7 1 2 3 4 % broken 5 6 7 high-quality 8 other 0 Cl. A Cl. B Cl. B Cl. C Cl. C Cl. D Cl. D Cl. E Cl. E Cl. F Cl. F Other CA Other CA 10 20 30 40 50 60 70 80 high-quality 90 100 other Figure 12.12 Selected attributes recorded on large quartz flakes within the Camping Area Scatter. Cl. = Cluster. A–B (top): Average negative flake scar (blue bars) and platform counts (orange bars) for high-quality materials (A – left) and lower quality materials (B – right). C–D: average mm/scar values (C – left) and proportion of broken flakes (D – right) for high-quality materials (black bars) and lower quality materials (grey bars). 260 Large flakes on lower quality materials displayed 1–9 scars, 2.32–12.72 mm/scar and 1–4 platforms. However, the collection was dominated by dual platform flakes (n=42), and those with 2–5 scars (n=80). Reduction intensity measures peaked in Cluster E and were lowest in Cluster F, but the small sample sizes must be considered; elsewhere, variation in average values was fairly subdued (Figure 12.12b–c). Clusters B–C and the surrounding material exhibited the greatest range in individual reduction intensity measures, but values were fairly diverse throughout. Broken flakes were more common than among the high-quality materials – while Clusters D–F yielded only whole flakes, the remainder had 11–25% of theirs broken (Figure 12.12d). Nineteen flakes exhibited significant edge damage; these were most common in Cluster B and the surrounding material, which each yielded five (Appendix G.5). Retouched Flakes A total of ten retouched flakes were found in the Camping Area Scatter. All were quartz, and all but one was backed. High-quality materials dominated, and just a single unifacially retouched piece and a backed artefact were made on lower quality materials. Retouched pieces were most common in Cluster A (n=4), while none were found in Clusters C or F (Table 12.8). The unifacially retouched piece was a large flake (53.65 g) on fairly rough-textured quartz. Several flakes had been removed from the ventral surface, in areas where flake morphology provided a good platform – this probably reflects the use of a flake as a core. None of the retouched pieces were broken, across higher or lower quality material. Three of the higher quality pieces exhibited more significant edge damage, and were restricted to Clusters A–B and the surrounding material (Appendix G.6). Manuports A single manuport was found in Cluster A, comprising a large unmodified chunk of silcrete weighing 25.4 g. It was rounded but showed no evidence of battering but may have been collected for future use as a core, hammerstone or grindstone. 261 12.6.7 CA1 A total of 321 artefacts were recovered from CA1, comprising 304 quartz (including 115 of high quality), nine silcrete, five BIF and three chalcedony (Tables 12.4 and 12.9). Non-quartz artefacts were concentrated in XU2–5 (n=15), and high-quality materials contributed around 45 to > 50% of assemblages from XU2–4 (Appendix G). Artefact discard rates were highly varied in CA1, ranging from < 2 to > 30 artefacts/100 years. Rates were fairly consistent from XU2–5, around 17–19 artefacts/100 years, and peaked in XU6. Rates then declined with depth until XU9, which registered the lowest value (Figure 12.13). Table 12.9 Stone artefact types by XU in CA1. Figures in parentheses refer to high-quality materials. XU Core Small fragment Large fragment Small flake Large flake Retouch TOTAL 1 0 4 0 3 (2) (2) 7 (4) 2 3 0 2 4 (1) 7 (3) 3 0 10 (9) 16 (17) 2 (5) 7 (14) 0 1 (1) 19 (15) 33(35) 4 5 6 7 8 9 TOTAL 3 0 0 0 0 0 5 0 7 8 8 (3) 2 (2) 1 41 (9) 3 6 1 8 2 (1) 0 23 (1) 9 (20) 22 (7) (14) 4 (7) 1 (2) 5 70 (76) 17 (7) 12 (2) 6 10 2 0 56 (30) 2 (3) (1) 0 0 0 0 3 (7) 34 (30) 47 (10) 15 (14) 30 (10) 7 (5) 6 198 (123) no. artefacts/100 yrs 0.00 5.00 10.00 15.00 20.00 25.00 30.00 35.00 40.00 1 2 3 CA1 XU 4 5 6 7 8 9 Figure 12.13 Artefact discard rates for pit CA1. 262 Cores Five cores (all quartz) were present in CA1, all between XU3 and XU4 (Table 12.9). All were quite fine-textured, but exhibited natural fracture planes and flaws, so were not classified as high-quality quartz. Two larger cores were present (16.15 and 82.15 g), one each in XU3 and XU4, and three smaller (< 4.5 g). The larger cores had higher flake scar (n=9–12) and platform counts (n=4–5), but reduction intensity was lower when the size of nodules was considered (n=1.35–9.13 g/scar). In contrast, the smaller cores were all bipolar, and averaged five scars, three platforms, and 0.5 g/scar. Two cores (the largest and one smaller), both from XU4, had 25–50% of their surface unmodified (Appendix G.1). Flaked Fragments A total of 50 small quartz fragments were present in CA1, over half of which were found between XU5 and XU7 (n=26), while XU4 yielded none. High-quality materials were only present in XU2–3 and XU7–8 (Table 12.9, Appendix G.2). CA1 also yielded 24 large fragments, all quartz. These artefacts were especially common in XU5 (n=6) and XU7 (n=8), while other XUs preserved no more than three. XU8 yielded the only large fragment on high-quality material (Table 12.9) – it weighed just 0.25 g and displayed five scars (0.05 g/scar). The remaining large fragments preserved 1–4 scars (avg. 2.1), 1–3 platforms (avg. 1.7) and 0.05–2 g/scar (avg. 0.61). Reduction intensity was highest in XU1, but XU7 preserved the greatest range of flake scar and platform counts (Appendix G.3). Flakes Small flakes were the most common artefact type in CA1, with a total of 146 present. They were particularly common in XU3–5, but numbers then declined towards the base of the pit (Table 12.9). Small flakes were one of the few categories where pieces made on high-quality materials (n=76, including 5 non-quartz) outnumbered those of lower quality. High-quality materials were particularly dominant in XU3–4, and XU6. Several examples from XU3 came from highly reduced parent cores, based on the frequency of small, overlapping flake scars. Overall, < 20% of small flakes were broken. Separate breakage data were not available (i.e. by raw material quality), but rates were often higher in those XU where high-quality materials were 263 more common, which may indicate those were more frequently broken (Figure 12.14). The breakage rate was highest in XU6 (n=35.71%), where all small flakes were made on high-quality material (Appendix G.4). % high-quality 0.00 20.00 40.00 60.00 % broken 80.00 100.00 0.00 0 1 1 2 2 3 3 4 4 40.00 60.00 80.00 100.00 XU CA1 XU 0 20.00 5 5 6 6 7 7 8 8 9 9 Figure 12.14 Proportion of small flakes on high-quality material (left), and breakage rate for all small flakes (right) from pit CA1. Large flakes were the second most common artefact type in CA1, with a total of 86 present, including 30 on high-quality materials (28 quartz, 2 BIF). Nearly 70% of large flakes (n=56) were found in XU3–5, and no high-quality materials were present below XU5 (Table 12.9). High-quality flakes displayed 1–7 flake scars (avg. 2.9), 2.39–11.52 mm/scar (avg. 5.08) and 1–4 platforms (avg. 1.9). Reduction intensity values peaked in XU4 and XU1 and were lowest in XU5 (Figure 12.15a, c). XU3 displayed the greatest range of individual values, as well as fairly high averages, and preserved the only piece with > 25% of its surface unmodified. No flakes were broken in XU1 or XU4, where reduction intensity values were highest, while half of those from XU5 were (Figure 12.15d). Only three flakes had suffered significant edge damage, one each from XU2, XU3 and XU5 (Appendix G.5). Two flakes from 264 XU3 preserved original crystal facets, while one from XU4 displayed long, narrow, parallel flake scars. Flakes on lower quality materials preserved 1–7 scars (avg. 3.2), 1.69–15.62 mm/scar (avg. 5.27) and 1–3 platforms (avg. 2), indicating a similar reduction intensity to high-quality materials, at a pit-wide level. Average values peaked in XU7–8, and XU1, but variation between individual XUs was much more subdued than in the high-quality assemblage (Figure 12.15a–c). Two flakes had 25–50% of their surface unmodified (one each from XU4 and XU6), while values were higher in XU5–6 (50–75%). Breakage rates were highest in XU2–3 (33–50%), while XU8 yielded no broken flakes (Figure 12.15d). Six had severe edge damage, including two each from XU4 and XU7 (Appendix G.5). number 0 0.5 1 1.5 number 2 2.5 3 3.5 0 4 No. NFS 1 0.5 1 1.5 2 2.5 3 3.5 1 No. platforms 2 2 3 3 4 4 CA1 XU CA1 XU No. platforms 5 5 6 6 7 7 8 8 9 9 mm/scar 0 1 1 2 3 % broken 4 5 6 7 0 high-quality 1 2 2 3 3 4 4 CA1 XU CA1 XU 10 20 30 40 50 60 70 80 90 100 high-quality other other 5 4 No. NFS 5 6 6 7 7 8 8 9 9 Figure 12.15 Selected attributes recorded on large quartz flakes, by XU, in CA1. A–B (top): Average negative flake scar (blue bars) and platform counts (orange bars) for high-quality materials (A – left) and lower quality materials (B – right). C–D: average mm/scar values (C – left) and proportion of broken flakes (D – right) for high-quality materials (black bars) and lower quality materials (grey bars). 265 In XU2, six of the seven large flakes (two on lower, four on higher quality material) and several of the smaller flakes preserved a crust on their dorsal surfaces. This crust overlay some negative flake scars but had been removed in other areas when subsequent flakes were detached (Figure 12.16). No flakes displayed any crusting on their ventral surfaces, so it was probably formed when flakes were still attached to their parent cores. Scar distribution indicates there was some length of time between flaking episodes, separated by the crust-forming event. Figure 12.16 Dorsal surface of three quartz flakes from CA1 XU2. Note the brown crust that covers most of the dorsal surface (including negative flake scars) and has been removed in other areas; this crust could not be removed by washing or rubbing. Crust was absent on ventral surfaces. White bar = 1 cm. Retouched Flakes A total of 10 retouched flakes were found in CA1, comprising nine quartz and one chalcedony. Retouched pieces were most common in XU4 (n=5), and none occurred below XU6. The three artefacts made on lower quality material were restricted to XU3–4 (Table 12.9); all were backed and unbroken, but one exhibited significant edge damage. The seven high-quality pieces comprised five backed and two unifacially retouched pieces. The latter occurred in XU1, where backed artefacts were not found. Both unifacial pieces terminated in a point, and had flakes removed from their ventral surfaces. Each also preserved a blunt margin of partial negative flake scars, created by prior flake removals when the piece was attached to its parent core – these pieces may have been used in the same way as intentionally backed pieces. One also preserved patches of a thick, resinous material near the 266 blunted margin, indicating these pieces had probably been hafted. Neither had been broken or heavily damaged, but two backed artefacts from XU4–5 had suffered transverse snaps (Appendix G.6). 12.6.8 CA2 A total of 198 stone artefacts were recovered from CA2, comprising 192 quartz, three silcrete and three BIF, across all artefact types; 83 pieces were made on highquality materials. High-quality materials were proportionally more common in XU1–2 but were present throughout the pit. They outnumbered lower quality material in just two categories: small flakes and retouched pieces (Table 12.10). Despite the fairly large assemblage, artefact discard rates were low, due to the extremely slow sedimentation rate. Values peaked at < 6 artefacts/100 years in XU2 and dropped as low as 0.99 artefacts/100 years in XU6 (Figure 12.17). Table 12.10 Stone artefact types by XU in CA2. Figures in parentheses refer to high-quality materials. XU Core Small fragment Large fragment Small flake Large flake Retouch Manuport TOTAL 1 2 3 4 5 6 TOTAL 0 0 0 1 0 0 1 5 6 7 10 (3) 10 (2) 1 (2) 39 (7) (2) 1 (1) 1 2 (1) 3 5 12 (4) 2 (5) 3 (15) 11 (18) 20 (5) 8 (3) (1) 44 (47) 2 (1) (1) 4 (4) 5 (7) 4 0 15 (13) (1) 1 (7) 1 (1) 1 (2) (1) 0 3 (12) 0 0 0 1 0 0 1 9 (9) 11 (24) 24 (23) 40 (18) 25 (6) 6 (3) 115 (83) no. artefacts/100 yrs 0.00 1.00 2.00 3.00 4.00 5.00 6.00 7.00 1 2 CA2 XU 3 4 5 6 Figure 12.17 Artefact discard rate in pit CA2. 267 Cores Just one core was present in CA2, recovered from XU4, where artefacts were the most numerous (Table 12.10). The core was made on fine-textured quartz, but natural fracture planes were evident. It weighed 23.15 g and exhibited 16 negative flake scars (1.45 g/scar) initiated from six different platforms. Less than 25% of the surface was unmodified, and the core was not bipolar (Appendix G.1). Flaked Fragments A total of 46 small fragments were found in CA2, all quartz. These artefacts were most numerous between XU4 and XU5, which together yielded over half (n=25) of the assemblage. Only seven were made on high-quality materials – these were restricted to XU4–6 (Table 12.10). All 16 large fragments were quartz, and were especially common in XU4–6, where eleven were found. Just four were made on high-quality materials, which were particularly frequent in XU1–2 (Table 12.10). On average, the high-quality fragments preserved 1.8 flake scars, 0.33 g/scar and 1.8 platforms. Most values were similar for lower quality materials, but flake scar averages were higher (avg. 2.4 scars, 0.34 g/scar, 1.8 platforms – Appendix G.3). Temporal trends could not be analysed as fragments were clustered in XU5–6, and small sample size plagued the upper deposits. Flakes Small flakes were the most abundant stone artefact type in CA2, comprising 88 quartz, two BIF and one silcrete; over half (n=47) were made on high-quality materials. Small flakes were especially common in XU3–4, where nearly 60% of the assemblage was concentrated, but most high-quality pieces came from XU2–3 (Table 12.10). Three of the high-quality flakes from XU3–4 exhibited long, narrow, parallel flake scars, indicating they had been struck from highly reduced, standardised parent cores. In XU1–5, breakage rates generally increased in concert with the proportion of high-quality materials, implying that these pieces were more often broken; XU6 was hampered by small sample size (Table 12.10, Figure 12.18). 268 % high-quality 20.00 40.00 60.00 % broken 80.00 100.00 0.00 0 1 1 2 2 3 CA2 XU CA2 XU 0.00 0 20.00 40.00 60.00 80.00 100.00 3 4 4 5 5 6 6 Figure 12.18 Proportion of small flakes on high-quality material (left), and breakage rate for all small flakes (right) from pit CA2. Of the 28 large flakes found in CA2, all were quartz except for one silcrete and one BIF flake from XU1–2; higher quality materials were almost as common as lower quality pieces. Large flakes were concentrated in XU3–4, which yielded a total of 20, while all other XUs preserved no more than four each (Table 12.10). Among the high-quality materials, flakes displayed 2–5 flake scars, 2.16–7.6 mm/scar and 1–3 platforms. All three values indicated that reduction intensity increased with depth, until its peak in XU4, but the limited samples in XU1–2 must be considered (Figure 12.19a, c). None of the flakes in XU1–2 were broken, but 50–67% of those from XU3–4 were (Figure 12.19d). The lower quality materials exhibited a greater range of reduction intensity values, and displayed 1–6 scars, 2.84–12.34 mm/scar and 1–3 platforms. These values peaked in different XUs – flake scar and platform averages were highest in XU3 and XU5, respectively, while XU1 produced the lowest mm/scar value (Figure 12.19b–c). Between 50 and 67% of flakes were bipolar in these same XU. XU3 also yielded the greatest range of individual flake scar and platform values (Appendix G.5). No lower 269 quality flakes were broken, but two had suffered significant edge damage (Figure 12.19d). number 0 0.5 1 1.5 number 2 2.5 3 0 3.5 0.5 1 1.5 2 2.5 3 4 No. NFS No. NFS 1 3.5 1 No. platforms 3 3 CA2 XU 2 CA2 XU 2 No. platforms 4 4 5 5 6 6 mm/scar 0 1 2 3 4 % broken 5 6 7 0 8 10 20 30 40 50 60 70 80 90 100 high-quality 1 1 other high-quality other 2 3 3 CA2 XU CA2 XU 2 4 4 5 5 6 6 Figure 12.19 Selected attributes recorded on large quartz flakes, by XU, in CA2. A–B (top): Average negative flake scar (blue bars) and platform counts (orange bars) for high-quality materials (A – left) and lower quality materials (B – right). C–D: average mm/scar values (C – left) and proportion of broken flakes (D – right) for high-quality materials (black bars) and lower quality materials (grey bars). Several of the high-quality quartz flakes were noteworthy for various reasons. Two specimens from XU1 and XU3 exhibit original crystal facets, indicating the use of a large crystal as a core. Four specimens from XU4 exhibited interesting flake scar patterns, in the form of long, narrow, parallel scars, and/or a series of uniformly sized and distributed partial scars that formed notches along a lateral margin. The latter may be produced from more formal cores with a series of narrow, parallel flake scars. If such a core is rotated and new flakes are struck – removing the old platform or base of the core – these flakes will retain a scalloped edge where the new flake 270 intersects with the existing scars (Figure 12.20). Interestingly, while the flake scar patterns indicated the production of long, narrow quartz flakes, only two were found, in XU3–4. One exhibited a length/width ratio of 4.84, and the other 2.7 despite having suffered a transverse snap (Appendix G.5). Figure 12.20 A (left): quartz flake from CA2 XU4. Lateral margins preserve notches formed by previous flake removals. Black bar = 1 cm. B (right): an idealised core that could produce notched flakes. Note the long, parallel flake scars, and the scalloped edge evident in cross section. If the platform or base of the core was removed, flakes would exhibit similar scalloping. Retouched Flakes A total of 15 retouched flakes were present, all but three on high-quality quartz. All XUs yielded no more than three retouched artefacts, except XU2 which preserved eight; none were present in XU6 (Table 12.10). Nine of the retouched pieces were backed, while the remainder exhibited unifacial retouch. The lower quality materials were exclusively unifacially retouched – two had scalloped retouch along one or more margins, while the other was a larger flake (4.1 g) that had subsequently been used as a core. None were broken or exhibited severe edge damage. The highquality assemblage comprised nine backed pieces and three that had been unifacially retouched; one of the latter preserved traces of possible hafting resin. Only three of the high-quality quartz pieces were broken, all backed, from XU2 (n=1) and XU4 (n=2). No unbroken pieces had suffered significant edge damage (Appendix G.6). 271 Manuports A single silcrete manuport was found in CA2, in XU4. The piece weighed over 13 g and had a very weathered, rounded surface. It may preserve weathered negative flake scars, but the coarseness of the material made this difficult to confirm. The piece lacked the battering associated with hammerstones, but again the poor texture may mask these features. 12.7 OTHER CULTURAL MATERIAL Some ochre and flaked glass was also found at Mulka's Cave. Ochre was relatively scarce, totalling 0.6 g, and was only present in CA1 (XU7) and CA2 (XU3–4). Three colours were present: red, brownish-red, and brownish-yellow. The latter was most abundant, comprising 0.4 g, while red was the scarcest (0.05g – Table 12.11). The sample size was too small for any further analysis. Table 12.11 Location, weight and colour of ochre samples from Mulka's Cave. Pit XU XU depth (mm) Wt (g) Colour CA2 3 100–160 0.15 Brownish-red CA2 4 160–220 0.05 Red CA1 7 300–350 0.4 Brownish-yellow Unworked glass was not recovered in any of the excavations discussed herein, so is not considered further. Just one example of flaked glass occurred, on the Entrance Platforms outside Mulka's Cave. Several fragments of unworked glass were found in the same area, but the colour of these pieces (clear and amber) indicates they derived from different source vessels to the worked piece. The flaked piece was pale green and had flake scars along one margin, effectively blunting it, while the opposing margin remained sharp (Figure 12.21). Such a dichotomous distribution is unlikely to result from post depositional processes, and indeed the Entrance Platforms receive minimal foot traffic (see Chapter 12.4.1). There were no distinguishing features on the glass, so it can be dated only in the broadest sense, as post European arrival in Western Australia. The piece weighed 1 g and measured 17 mm long and 13.59 mm wide. Three large (≥ 10 mm) negative flake scars were present, each initiated from a different platform. 272 Figure 12.21 Flaked glass from Entrance Platforms at Mulka's Cave. Note the small negative flake scars concentrated along the right distal margin, while the other remains sharp and intact. Stippling represents unmodified surface, black bar = 1 cm. 12.8 CONCLUSION Water was fairly abundant at Mulka’s Cave, regardless of winter rainfall quantity, due to the properties of the cleft rockhole. It filled later and held water for shorter periods as winter rainfall decreased, but post-winter longevity was strongly influenced by spring falls. In contrast to the other sites investigated herein, Mulka's Cave has a history of research beginning in the 1930s. Much of the earlier work concentrated on the rock art while later, more detailed studies (e.g. Bowdler et al. 1989; Gunn 2004, 2006) produced conflicting results. The author focussed on three main parts of the site: The Main Cave Area, the Humps Scatter, and the Camping Area. The former was problematic as many of the collections are unusable because erosion of the rockshelter sediments has redistributed archaeological material. Inside Mulka's Cave, occupation commenced around 9650 cal. BP, while the Camping Area deposits are slightly more recent – CA2 dated to 6700 cal. BP and CA1 to 2500 cal. BP. Artefacts began accumulating on the Humps outcrop after 1200 cal. BP. The different dates for CA1 and CA2 illustrate the problems with applying dates from one pit to another, even when they are located in the same part of a site. Over 950 stone artefacts were recovered – most came from the Camping Area, while cultural material was scarce in the Main Cave Area. The assemblage was heavily quartz dominated, with a small quantity of BIF, silcrete and chalcedony restricted to 273 the Camping Area, which also yielded the only ochre fragments. Artefacts on highquality material outnumbered those on lower quality material in only one area, the Column Sample, but high-quality materials were also fairly common in the Camping Area. Most subsurface stone artefact assemblages were dominated by small flakes, but small fragments were also fairly common. A number of pieces from around the site indicate the use of large quartz crystals as cores, and several other pieces, generally on high-quality quartz, came from highly reduced or standardised parent cores. A single flaked glass piece found on the Entrance Platforms indicates that Mulka's Cave was visited by Aboriginal people after Europeans arrived in Western Australia. 274 CHAPTER 13 IDENTIFYING TECHNOLOGICAL PROVISIONING SYSTEMS 13.1 INTRODUCTION The aim of this chapter is to identify the technological provisioning systems used at Gibb Rock, Anderson Rocks, and Mulka's Cave by considering the data presented in Chapters 10–12 against the archaeological signatures formulated in Chapter 9.5.3. Raw material quality and conservation, as well as the intensity and length of reduction sequences, are used to identify the dominant provisioning systems (Individual and Place Provisioning), as well as supplementary systems (Gearing Up and Short-term Expedient). Gibb Rock is evaluated first, as it provides a clear signature that forms a useful baseline against which to evaluate assemblages from other sites. The Anderson Rocks and Mulka's Cave assemblages are then considered. In both cases, larger sample sizes, varied pit/surface collection locations and the availability of radiocarbon dates allows spatio-temporal variation in provisioning systems to be evaluated. 13.2 GIBB ROCK Given the small artefact assemblages present at Gibb Rock, one of the best sources of data was the quality of quartz. As noted earlier (see Chapter 9.5.3), Place Provisioning strategies require collecting the best material available within the local foraging range (referred to hereafter as 'stockpile-quality material', i.e. moderate– good quality), and stockpiling it at a site where future occupation is predictable. In contrast, Individual Provisioning and Gearing Up require higher quality materials to prolong artefact use-lives and, especially under Individual Provisioning, to minimise waste by ensuring predictable flaking properties. The quartz used at Gibb Rock is, at best, variable, and at worst, incredibly poor. The outer, weathered surface is coarsetextured, friable and plagued by numerous natural fracture planes and flaws (Figure 9.3: top left). The internal material is closer textured, but still heavily flawed. This quartz would be wholly unsuitable for the manufacture of a mobile toolkit, and it is unlikely that such material would be collected and transported to the site as part of a Place Provisioning strategy, when the assemblages from nearby sites indicate that 275 much better quality quartz was readily available in the wider area. Instead, the assemblage seems to reflect the use of local raw material, as all artefacts are visually consistent with the material from on-site exposures; the use of local material occurs under Short-term Expedient strategies (see Chapter 2.4.1). Where the relevant data were available, they primarily indicated the lower reduction intensity that is characteristic of Short-term Expedient strategies. At Scatter 1, negative flake scar and platform averages were low across all artefact types (2-3 scars, 1.5–2 platforms). Individual values were especially low when considered in conjunction with artefact size – cores averaged 43.51 g/scar, large fragments 1.84 g/scar, and large flakes 8.46 mm/scar. In addition, all cores and over 13% of large flakes had more than 50% of their surface unmodified (Appendix F). Bipolar reduction was more than twice as common as freehand reduction, but rather than indicating increased reduction intensity or raw material conservation, the prevalence was probably linked to the poor quality of the local quartz; bipolar reduction is often helpful when flaking poor-quality quartz. Furthermore, pieces detached along natural fracture planes (which were abundant in the local quartz), can mimic the axial termination characteristic of quartz bipolar flakes. This is especially problematic where crushing cannot be observed due to poor texture. Only eight artefacts from GR1 and Scatter 2 provided reduction intensity data: large flakes and cores. Negative flake scar and platform averages from large flakes were similar to Scatter 1, but the flakes there were considerably larger. The quartz cores had been more intensively reduced, averaging eight flake scars and four platforms, on cores weighing < 12 g. These cores were made on better quality material than most at Gibb Rock, but still not predictable or fine-textured enough to form part of a transported toolkit. Instead, it seems visually consistent with the local quartz located below the weathering front. The increased reduction intensity may therefore indicate removal of the outer, weathered quartz to access the more closely textured material within. The cores were certainly flaked on-site then, but the scarcity of large flakes may indicate that the reduction event and core discard occurred in different parts of the site. Overall, despite some more intensive reduction than might be expected under a Short-term Expedient strategy, there is insufficient evidence of any other technological provisioning system. 276 13.3 ANDERSON ROCKS At Anderson Rocks, some assemblages were hampered by sample size issues, but technological systems were generally identifiable throughout, particularly when the Short-term Expedient signature from Gibb Rock was used as a basis for comparison. The surface material is discussed first, then the two test-pits: AR1 and AR2-3. A temporal timescale can only be defined for AR2-3; no dates were obtained for AR1 and there is no reason to assume that sedimentation proceeded at the same rate. 13.3.1 Surface Collections Three of the four surface collections (all but Scatter 2) provided fairly clear evidence of Place Provisioning. These assemblages were made on stockpile-quality material, comprising moderate- to good-quality quartz (generally with some flaws and variability in texture), as well as some silcrete. Negative flake scar and platform averages were higher than at Gibb Rock (avg. 3.1–7 scars, 1.8–3.5 platforms, combined across relevant artefacts from all three scatters), and artefacts were generally smaller. Individual values were often quite varied but more evenly distributed, rather than clustered toward the lower end of the range. Cores were almost exclusively small and heavily reduced, and three of the four were bipolar – one likely as a response to blocky raw material texture, and the remaining two due to their small size. In some cases, flake scars on cores outnumbered flakes found within the same surface collection. These cores clearly had a longer use-life and had been flaked in various parts of the site. The available evidence points to the longer on-site reduction sequences that characterise Place Provisioning (Figure 9.5). Scatters 1 and 3 also preserved evidence of more formal assemblages made on high-quality materials, primarily quartz and a small amount of chalcedony. Reduction intensity was similar to the Place Provisioning assemblage, across large fragments (avg. 2.5 scars, 2 platforms, 0.15 g/scar) and large flakes (avg 2.6 scars, 2.2 platforms, 6.12 mm/scar). No artefacts retained parallel flake scars, or any other indicators of more standardised reduction techniques. Finally, three retouched pieces were present, none of which were broken or heavily damaged. The moderate reduction intensity combined with the lack of raw material conservation and standardised reduction is more consistent with a Gearing Up strategy, used within the Place Provisioning system. 277 Scatter 2 preserved a different signature. As elsewhere, the assemblage was made on two contrasting materials but, in this case, these comprise a very poor-quality local quartz (the scatter was located within a pan that contains an exposed quartz vein) and a few items of stockpile-quality (i.e. moderate–good) quartz. The local material had been lightly reduced; pieces averaged 1.7 scars and 1–2 platforms. These values are similar those from Gibb Rock, but the Anderson Rocks artefacts were generally smaller, probably due to the limited width of the quartz vein. Raw material quarrying, as well as artefact manufacture, use and discard all occurred within a metre or so of the quartz source – this is indicative of a Short-term Expedient strategy, where artefacts were manufactured to meet an immediate need, and discarded thereafter. The remaining three artefacts at Scatter 2 were made on finer textured quartz that exhibited some variability and flaws. These artefacts comprised a single core, large fragment and large flake. All had been more intensively reduced than the local quartz (avg. 3–7 scars, 3 platforms), indicating they had previously been flaked elsewhere on-site, under a Place Provisioning strategy. As Scatter 2 was located within a pan, it is possible that these artefacts washed downslope via the interconnected channel system described earlier (see Chapter 10.2.1). If they were intentionally discarded here, they must post-date the local quartz assemblage, as they could otherwise have provided a source of better quality material for expedient artefact manufacture. 13.3.2 AR1 In AR1, XU1–9 preserved a different technological signature to the lower deposits, XU10–13. In the upper deposits, stockpile-quality material was most abundant, predominantly quartz with a small amount of silcrete. This assemblage was dominated by small fragments (n=43), which contribute no useful information on reduction intensity, but their prevalence does indicate a considerable amount of onsite reduction. Just six large fragments and five large flakes contributed useful data on reduction intensity. Negative flake scar and platform values were very similar to those at Gibb Rock (avg. 2–2.2 scars, 1.6–1.7 platforms), but the average weight/length per scar was higher in AR1 (avg. 3.76 g/scar, 9.49 mm/scar), indicating a lower reduction intensity. However, the range of individual values was similar, and the small sample size at AR1 makes the average values somewhat 278 misleading as they are easily skewed by outliers. Four retouched pieces (all backed) were present, on stockpile-quality material – only one was broken. These may have been required for specific on-site tasks, or been part of a Gearing Up assemblage, since shorter, predictable artefact use-life might not always require the use of highquality materials. Retouched pieces would not be manufactured under a Short-term Expedient strategy, since similar artefacts would be available within the transported toolkit. Therefore, these deposits probably represent Place Provisioning, with lower reduction intensity indicating that the area was used in a different, less intensive manner to other parts of Anderson Rocks. There may, of course, have been some mixing of contrasting signatures within the same archaeological deposits. The remainder of the XU1–9 assemblage was made on high-quality material and exhibited characteristics of both Individual Provisioning and Gearing Up. Flake scar and platform counts were similar to the Place Provisioning assemblage (as may be expected under Gearing Up), but artefacts were slightly smaller. There was no evidence of standardised reduction techniques that may occur under Individual Provisioning, but partial scars on one flake indicate that it came from a highly reduced parent core. Few artefacts were discarded, and the single retouched piece was broken, while both large flakes were not. The inconsistent reduction intensity and discard behaviour may indicate a mixture of both provisioning systems, or simply be an artefact of small sample size. The lower deposits (XU10–13) yielded much fewer artefacts, so sample size was more problematic. Nevertheless, in contrast to the upper deposits, the assemblage contained more even quantities of lower and higher quality materials. The lower quality material comprised eight small and three large fragments. The latter had higher flake scar averages than at Gibb Rock (2.5 scars, 0.17 g/scar), but showed less core rotation. The small sample makes it impossible to discern between a Shortterm Expedient strategy (using stockpiled materials) or Place Provisioning – the scarcity of artefacts may support the former, but sampling bias must also be considered. The high-quality quartz comprised five small flakes, two retouched pieces and one large flake; all were broken except two small flakes. The lack of artefacts and the dominance of broken pieces may be indicative of the raw material conservation that occurs under Individual Provisioning, since intentional discard (i.e. 279 not including waste products created by retouch) should be limited to broken or unusable items that cannot be repaired. 13.3.3 AR2-3 AR2-3 displayed several shifts in provisioning systems over time. Due to the depth of the pit, it has been divided into three sections for discussion: the upper (XU1–11), middle (XU12–21) and lower deposits (XU22–32). Upper deposits (XU1–11) XU1–3 (138–0 cal. BP) yielded a fairly even mix of stockpile- and higher quality material. The latter comprised small fragments and small flakes made on high-quality quartz, along with a single small BIF flake. This assemblage seems more consistent with reworking existing artefacts rather than manufacturing new ones, so is more indicative of Individual Provisioning. The stockpile-quality material comprised a few small flakes as well as a single large flake, small fragment and large fragment. Data on reduction intensity were scarce, so this assemblage could not be reliably compared to the archetypal Short-term Expedient assemblage from Gibb Rock. However, the scarcity of artefacts may indicate a Short-term Expedient strategy using stockpiled material. In contrast, XU4–8 (1220–138 cal. BP) preserved evidence of more intensive reduction of stockpiled material. Negative flake scar (3–4) and platform averages (2– 2.2) were higher across cores (n=1), large fragments (n=2) and large flakes (n=14), all of which were generally smaller than their Gibb Rock counterparts. Large flakes also exhibited a greater range of individual values, with up to seven scars and four platforms evident. These features are more characteristic of Place Provisioning, where longer reduction sequences occur. XU5–8 (1220–354 cal. BP) also yielded some artefacts on higher quality materials, comprising quartz and some BIF. While high-quality quartz may be stockpiled and used in a Gearing Up strategy, high-quality exotics (such as BIF) are more likely to be part of a toolkit where artefacts need to be preserved indefinitely. The BIF artefacts comprised a small whole flake (just under 10 mm in percussion length – too large to have been produced by retouch) as well as two large flakes. These must 280 have been struck from a transported core, which would invariably be part of a transported toolkit. All large flakes were relatively long and narrow, and scars of the same morphology were found on one large flake and the single core – this indicates the use of a more formal reduction technique. Two of the four large flakes were broken (and another was possibly broken), and the single core was very small (2.45 g), bipolar, and probably discarded because no further flakes could be removed from it. This use of exotic materials, standardised reduction techniques and selective discard are all characteristic of Individual Provisioning. XU9–11 (1870–1220 cal. BP) was dominated by high-quality material. Only four artefacts were made on stockpile-quality material, and only one yielded data on reduction intensity, so that sample was too small to evaluate. The higher quality material was primarily quartz, with one chalcedony piece also found. Artefacts comprised a core, three small fragments, 25 small flakes (mostly broken), one large flake, and two retouched pieces (both unbroken). Primary reduction clearly occurred on-site, as many of the broken small flakes would have measured > 10 mm when whole. Some level of secondary modification may be indicated by the small fragments, and those small flakes fine enough to have been detached by retouch. A single small flake exhibited long, narrow, parallel flake scars, but there were no other traces of standardised reduction. Data on reduction intensity were scarce, but the single core was very small (1.85 g), bipolar and preserved eight scars initiated from two platforms. The discard of unbroken retouched pieces is more consistent with Gearing Up, but Individual Provisioning is supported by the broken chalcedony flake, heavily exhausted core, and small flake with parallel scars. However, the latter two artefacts could be generated by Gearing Up, if there was some use of more formal, intensive reduction techniques to conserve high-quality material. It seems more likely, then, that the assemblage represents Gearing Up, but the small sample size makes the signature somewhat ambiguous. Middle deposits (XU12–21) In contrast to XU9–11, XU12–18 (3603–1870 cal. BP) exhibited clear evidence of Place Provisioning. There was abundant evidence of primary reduction, comprising 35 large flakes, seven cores, seven retouched pieces and five large fragments. Many of the small fragments (n=26) and small flakes (n=61) probably originated from 281 primary reduction, but some may have been produced by secondary modification. In addition to greater artefact numbers, reduction intensity was consistently high, averaging 2.7–5.9 scars, 2–3.1 platforms, 0.25–0.94 g/scar and 6.62 mm/scar. Individual values were also more varied than would be expected under a Short-term Expedient strategy. All cores weighed less than 11 g, and while the smaller ones had been flaked to exhaustion, larger ones retained a good amount of unmodified surface (25–50%) and preserved usable material. Bipolar reduction was common among cores (80%) and large flakes (45%), which may indicate some conservation of raw material at a core level. Finally, the seven retouched pieces may have been manufactured for a logistical foray (under Gearing Up), or they may simply have been required for specialist tasks. Higher quality materials were found throughout XU12–18, in varying quantities, but XU12–17 preserved a different signature to XU18. XU12–17 (3170–1870 cal. BP) yielded two large BIF flakes, three small BIF flakes and two small chalcedony flakes; all small flakes were whole and weighed < 0.05 g, so probably derived from retouch (Appendix E.4). Clearly, then, secondary modification of exotic materials occurred on-site, along with some degree of primary reduction. These activities are consistent with Individual Provisioning, if the presence of exotic materials can be considered a diagnostic feature of that strategy. The high-quality quartz assemblage from XU12–17 was much larger, and the number and range of artefacts indicated that a fair amount of primary reduction occurred. Reduction intensity was similar to the Place Provisioning assemblage (avg. 2.3–4 scars, 2–3 platforms), but the range of individual values was more restricted. Four flakes were quite long and narrow, and two had scars of the same shape, indicating some standardised reduction of high-quality quartz. These specimens had been more intensively reduced than the remainder of the high-quality quartz flakes (avg. 2.8 scars vs 1.8 scars), but the latter could rarely be oriented, so lacked platform data. The breakage rate of high-quality quartz flakes was very high – seven of nine were undoubtedly broken, compared to 39% of stockpile-quality quartz flakes. The latter averaged 5.2 mm thick for unbroken pieces and 2.4 mm for broken specimens, while the entire high-quality quartz assemblage averaged 2 mm thick. It is possible, then, that the higher breakage rate does not indicate selective discard, 282 but simply that a greater proportion of the assemblage was more susceptible to postdepositional breakage. The high-quality quartz assemblage clearly shows some evidence of Gearing Up, based on frequency of primary reduction, the lower reduction intensity, and the discard of unbroken retouched pieces. The more highly reduced standardised pieces could be part of the same strategy, or may support the Individual Provisioning strategy suggested by the exotic raw materials. In contrast, the entire high-quality assemblage from XU18 (3603–3170 cal. BP) appears to derive from Individual Provisioning. Reduction intensity data were somewhat limited, but values for large quartz flakes were higher (avg. 3.3 scars, 2.7 platforms) than even the more intensively reduced components from XU12–17. One large flake was long and narrow, and retained five long, narrow scars on its dorsal surface, evidence of standardised reduction techniques. One small flake retained an original crystal facet; large crystals would be the perfect choice for transported cores, since the material therein is generally homogenous and predictable, and the flat surfaces provide ideal platforms. Finally, there was evidence of preferential discard behaviour that was largely restricted to broken or non-functional items. Forty per cent of the large flakes were broken, and 83% of the retouched pieces. If these were broken post-deposition, the more numerous artefacts (flakes) would be expected to show a higher rate of breakage than the scarcer artefacts (retouched pieces), due to the greater likelihood of the former being encountered on the ground surface. Furthermore, the broken high-quality quartz flakes actually had a slightly greater average thickness (3.3 mm) than their unbroken counterparts (3.1 mm), an almost identical value to that of unbroken stockpile-quality flakes (3.4 mm). These data indicate that the high-quality artefacts were probably broken during use before being discarded. XU19 (3819–3603 cal. BP) was dominated by high-quality material (n=41), with a smaller quantity of stockpiled material present (n=17). The stockpiled material indicated Place Provisioning, as reduction intensity was higher than at Gibb Rock (avg. 2.5–3 scars, 2 platforms, 5.88 mm/scar, 0.71g/scar). In addition, a single backed piece was found; this piece would not be required under a Short-term Expedient strategy, as the required task could be performed with the transported toolkit. Among the higher quality material, reduction intensity was somewhat varied. 283 Values were higher for the single core, which was sufficiently small to discourage further reduction (4 scars, 4 platforms, 2.25 g, bipolar). Values were not particularly high for large flakes, and averaged 2.4 scars, 1.8 platforms and 6.53 mm/scar. Several flakes displayed smaller, overlapping flake scars on their dorsal surfaces, indicating they had been struck from heavily reduced parent cores, but there were no signs of standardised flaking techniques. Four of the five large flakes were broken, as was the single retouched piece. However, due to the small stockpile-quality assemblage, it was not possible to determine whether this reflected post-depositional breakage or preferential discard. As such, the high-quality assemblage may have derived from Individual Provisioning, Gearing Up, or a combination of both strategies. A Place Provisioning strategy clearly dominated in XU20–21 (4571–3819 cal. BP). The stockpile-quality material exhibited fairly high reduction intensity (avg. 2.5–8 scars, 2.1–4 platforms), and the single core was flaked by bipolar reduction due to its small size. Large flakes (n=19) showed extensive variation in individual flake scar (0–6) and platform counts (1–4), and seven had more than 25% of their surface unmodified. Several of these flakes had staining on their dorsal surface – this stain overlay some scars but had been lost in other areas due to subsequent flake removals. This indicates a reuse of older/stored material that is characteristic of Place Provisioning. Finally, three retouched pieces were present (all intact), indicating some Gearing Up, or the need for specialist tools for particular tasks. High-quality material was also well-represented in XU20–21, but reduction intensity data were only available from a single core and six large flakes. The core was quite heavily reduced (7 scars, 4 platforms, 0.88 g/scar) but, based on its weight (6.15 g), it could probably be reduced further. Large flakes averaged 3 scars and 1.6 platforms, which was fairly similar to the Place Provisioning assemblage. Two of these flakes had parallel flake scars, indicating some level of standardisation in previous flake removals, and most of the large flakes were themselves fairly long and narrow. Half of the large flakes were broken; this may represent selective discard, as these flakes were only slightly thinner, on average, than their unbroken counterparts. Two retouched pieces were also present – neither was broken or heavily damaged. Most of the assemblage seems more consistent with Gearing Up, since reduction intensity was not particularly high, and discard was not limited to 284 broken items or exhausted cores. Nevertheless, some isolated visits may have occurred under Individual Provisioning. Lower deposits (XU22–32) XU22–28 (7350–4571 cal. BP) yielded a smaller assemblage so the technological signatures were less clear. The deposit was dominated by stockpile-quality material, but small amounts of high-quality material were also present. The stockpiled material was fairly lightly reduced – average values were similar to those at Gibb Rock, except large fragments were considerably lighter (avg. 2–2.3 scars, 1.7–2 platforms, 8.84 mm/scar, 0.19 g/scar). This lower reduction intensity, combined with the overall scarcity of artefacts (and low discard rates), might indicate Short-term Expedient use of stockpiled material. A single retouched piece was present, which may indicate that some longer visits occurred, under a Place Provisioning strategy. Higher quality materials were represented by three small quartz fragments and 22 small flakes (19 quartz, 3 BIF). The absence of larger items, and the presence of exotic materials, is more consistent with artefact maintenance, and possibly some primary reduction of small transported cores, as would be the case under Individual Provisioning. The lowest deposits, XU29–32 (8200–7350 cal. BP) were associated with the smallest sample size. Stockpile-quality material was limited to two large flakes from XU30, so reduction intensity data were incredibly limited. Nevertheless, flake scar average was low (1.5) and one of the flakes had 25–50% of its surface unmodified. This appears to reflect Short-term Expedient use of stockpiled material. Higher quality quartz was more abundant, and comprised two small fragments, seven small flakes and three large flakes. Again, reduction intensity data were limited and of little value. However, there was clearly some on-site reduction of higher quality material, and some small flakes and fragments may have derived from secondary modification. Due to the small sample size, it is difficult to attribute this assemblage to Individual Provisioning or Gearing Up, but the scarcity of artefacts would seem to support the former, since reduction sequences would be shorter. 285 13.4 MULKA'S CAVE As in Chapter 12, the Mulka's Cave assemblages have been separated by site area: the Main Cave Area, the Humps, and the Camping Area. Sample sizes are variable both within and between areas, but most were large enough to permit fairly confident identification of technological systems. Radiocarbon dates were obtained from all three areas, allowing timeframes to be placed on most archaeological deposits, and the technological provisioning systems identified therein. 13.4.1 Main Cave Area In the Main Cave Area, small sample sizes hampered the identification of technological systems, particularly in MC1 and the Entrance Platforms. At the latter, only eight artefacts were present, all on stockpile-quality (moderate–good) material. Reduction intensity measures were generally higher than at Gibb Rock on the core and large flakes (avg. 3.4–4 scars, 2.2–3 platforms, 0.44 g/scar, 5.41 mm/scar), while large fragments were less intensively reduced (avg. 1 scar, 1 platform, 2.58 g/scar). All five large flakes had their entire dorsal surface covered by negative flake scars (i.e. no unmodified surface), meaning the parent cores had been previously flaked, elsewhere on-site. This reuse of stockpiled material is consistent with Place Provisioning, but the small number of artefacts means that reduction intensity measures can be heavily skewed by previous flaking events. The assemblage may represent a short site visit, when stockpiled materials were reduced in a Short-term Expedient manner, or simply indicate that this area of the site was used differently to others. Stockpile-quality material dominated MC1, but reduction intensity was fairly low, similar to that at Gibb Rock (avg. 2–2.1 scars, 2 platforms). While the lower reduction intensity is more consistent with a Short-term Expedient strategy, two retouched pieces were also found; as noted elsewhere, retouched pieces would not be produced under that strategy. The assemblage probably represents a level of Place Provisioning, but with a shorter reduction sequence than normal; this potentially indicates this part of the site was used less intensively or in a different manner to other parts of Mulka's Cave. High-quality materials, all quartz, were much less abundant. This assemblage comprised one core, two large flakes and two retouched pieces. The core was highly reduced (0.65 g), and made on a large natural quartz 286 crystal, an ideal choice for transported cores. The large flakes had been more intensively reduced than the stockpiled material (avg. 3 scars, 2 platforms, 5.12 mm/scar), and one preserved an abundance of small, partial scars, indicating it was struck from a highly reduced parent core. None of the large flakes or the retouched pieces were broken, but one large flake had suffered heavy edge damage. The sample size was too small to accurately separate Individual Provisioning from Gearing Up, and indeed each are supported by different features of the assemblage. The highly reduced transported core is more consistent with Individual Provisioning, while the discard of intact pieces is more synonymous with Gearing Up. Column Sample The Column Sample inside Mulka's Cave yielded the largest sample size in the Main Cave Area, so technological provisioning systems were more readily identified. In XU2–6 (1570–48 cal. BP), stockpile-quality material dominated over high-quality quartz, but the entire assemblage yielded limited data on reduction intensity, since it was dominated by small flakes (n=13), from which no relevant attributes were collected. The stockpiled assemblage comprised ten small flakes, two large fragments and one large flake; it may have originated from Place Provisioning or Short-term Expedient use of stockpiled material. The higher quality assemblage was even smaller, just three small flakes, a small fragment, large fragment and a large flake. While data were limited, reduction intensity measures were fairly high (4–6 scars, 3 platforms), especially when the size of pieces were considered (0.3 g/scar, 3.8 mm/scar). Furthermore, the dorsal surface of the large flake exhibited several small, partial scars, indicating it came from a highly reduced parent core – the flake was possibly detached to rejuvenate this same core. The reduction intensity and general scarcity of artefacts seems more consistent with Individual Provisioning than Gearing Up. XU7–9 (4250–1570 cal. BP) was dominated by high-quality material, but the abundance of large flakes on lower quality material means reduction intensity data were fairly abundant across both material types. The lower quality large fragment had been lightly reduced (1 scar, 1 platform), while flakes had similar flake scar and platform averages to Gibb Rock (2.1 scars, 1.8 platforms). However, by virtue of their small size, these flakes showed a greater reduction intensity (4.99 mm/scar). 287 Similarly, they demonstrated significant variation in individual scar, platform and unmodified surface values, considering the relatively small sample size. These features are more consistent with Place Provisioning, as reduction sequences are longer. The higher quality assemblage was generally consistent with Gearing Up – reduction intensity was similar to the Place Provisioning assemblage (2 scars, 1–1.3 platforms), and most large flakes/retouched pieces were discarded without being broken or heavily damaged. Only the single core, made on black smooth-textured quartz, had been more heavily reduced (6 scars, 4 platforms, 0.54 g/scar). No other artefacts were made on this same material, so the core may not have been flaked on-site, but the very small pit size must be taken into account. The lowest Column Sample deposits, XU10–11 (6000–4250 cal. BP), were associated with the smallest sample size, which comprised only eight small flakes and two small fragments, all on high-quality quartz. Reduction intensity data were not available, but the dominance of small artefacts indicates that on-site activities probably included artefact maintenance and/or flaking small, transported cores; this is more consistent with Individual Provisioning. 13.4.2 The Humps As in most other collections, the Humps Scatter yielded artefacts made on both higher and lower quality material, the latter suitable for stockpiling (n=14 and 7, respectively). The stockpiled material was heavily reduced (avg. 4.7–6 scars, 2–3 platforms, 0.37–0.43 g/scar, 4.96 mm/scar), so is more consistent with Place Provisioning than a Short-term Expedient strategy. Similarly, the presence of a single retouched piece is inconsistent with the latter. High-quality artefacts comprised six large flakes, three small flakes, two cores, one small fragment, one large fragment and one retouched piece. Based on the artefact types, some level of primary reduction clearly occurred in this area, and indeed all of the small flakes are too large to have been produced by retouch. Both cores were heavily reduced (avg. 6.5 scars, 2.5 platforms, 0.43 g/scar), bipolar, and probably discarded as they were too small to flake further. Reduction intensity was also fairly high for large fragments (7 scars, 3 platforms, 0.29 g/scar) but less so among large flakes (avg. 2.8 scars, 2.2 platforms, 6.74 mm/scar). Three of the large flakes were broken, while the retouched piece was intact. This assemblage seems more consistent with Gearing Up, based on the 288 discard behaviour as well as the frequency of on-site primary reduction. The higher reduction intensity probably reflects some effort to conserve better quality material, which would be encountered less regularly than stockpile-quality quartz. Stockpile-quality material dominated throughout TH1 (n=132), while higher quality material was comparatively scarce (n=24). The latter assemblage comprised 11 small fragments, eight small flakes, four retouched pieces and one large flake. Highquality material was most common between XU1–3 (601–0 cal. BP, n=14), while just five pieces were present in the lower half of the pit (XU4–6, 1120–601 cal. BP). The small flakes and fragments may have originated from secondary modification or artefact maintenance, and only the single large flake provided concrete evidence of on-site primary reduction. Three of four retouched pieces were broken, and one retained probable hafting resin, indicating it was part of a composite tool, which would be common in transported toolkits. The single large flake was also broken, and its long, narrow, parallel flake scars indicate that it came from a more standardised core. All of these features support Individual Provisioning. The TH1 stockpiled material comprised good mix of small (n=88) and large artefacts (n=44). Large fragments had a fairly low average scar and platform count (1.7 and 14, respectively) but showed a greater reduction intensity when their small size was considered (avg. 0.26 g/scar); large flakes were more intensively reduced than at Gibb Rock (avg. 2.4 scars, 2 platforms, 6.56 mm/scar). Overall, then, reduction intensity is more consistent with Place Provisioning. A total of 13 retouched pieces were also present – these may have been required for specific tasks under Place Provisioning, or represent some level of Gearing Up, since shorter artefact use-life may permit the use of slightly lower quality material. Nearly 82% of retouched pieces were broken, compared to just 40% of large flakes, which may indicate some selective discard. It is possible that some retouched pieces were part of Individual Provisioning assemblages (i.e. the definition of high-quality material was too narrow), but the Humps area may simply have been the focus of major retooling events prior to long logistical forays. 289 13.4.3 Camping Area The Camping Area preserved the largest stone artefact assemblages from Mulka's Cave, so technological provisioning systems were relatively easy to identify, in most cases. The surface material is discussed separately to each of the excavated testpits, CA1 and CA2. Camping Area Scatter Within the Camping Area Scatter, Clusters A–D yielded consistently similar signatures. Reduction intensity of stockpile-quality material was higher than at Gibb Rock across all measured artefacts, including cores (avg. 6–8.5 scars, 2–3 platforms), large fragments (avg. 2.7–5.5 scars, 1.3–2.5 platforms, 0.45–1.43 g/scar) and large flakes (avg. 3–3.8 scars, 1.7–2.3 platforms, 4.61–6.61 mm/scar). One large fragment from Cluster C exhibited long, parallel flake scars, indicating that formal reduction was not always confined to high-quality material. This higher reduction intensity, combined with greater artefact numbers, is characteristic of Place Provisioning. In addition, individual flake scar, platform and size values were quite varied, particularly in Cluster B, indicating longer reduction sequences. Two silcrete cores provided concrete evidence of reuse of stockpiled material – these cores displayed a total of 14 negative flake scars, but the entire Camping Area Scatter yielded only six large silcrete flakes. These cores must have been flaked elsewhere on site, or in the distant enough past that the resultant flakes were buried. This reuse of older material can skew reduction intensity data, but the effect is negligible on large, diverse assemblages such as those from Clusters A–D. High-quality raw material was much scarcer in Clusters A–D. Large flakes were more intensively reduced than the Place Provisioning assemblage (avg. 3–7 scars, 2–3 platforms, 2–6.1 mm/scar), but only one derived from a standardised core, based on the presence of long, parallel flake scars. Both cores were bipolar and displayed ten scars detached from 4–6 platforms, and were heavily reduced, weighing < 2.7 g. One of the large flakes weighed considerably more than others (4.15 g), so could have been used as a core if raw material conservation was integral, as it is under Individual Provisioning. Furthermore, none of the five retouched pieces were broken, and only one of the twelve large flakes was. Therefore, despite the smaller sample size, the discard behaviour seems more consistent with Gearing Up; the increased 290 reduction intensity and some standardised flaking may be indicative of attempts to conserve raw material at the core level. Clusters E and F both yielded much smaller assemblages (n=7 and 6, respectively), and indicated shorter term use of stockpiled material. In Cluster F, flake scar and platform averages were fairly low (1–3, 1–1.5 platforms, 5.52–6.75 g/scar), and the core had more than 75% of its surface unmodified. Values were considerably higher in Cluster E (avg. 5.5–6 scars, 3 platforms), but the small assemblage size indicates that much of this reduction happened elsewhere on-site, so does not accurately characterise the level of stone flaking in this area. A single retouched piece was present at Cluster E, a large flake (> 50 g) that had several flakes detached from its ventral surface. This probably reflects the use of an older, larger flake as a source of raw material, rather than retouching a flake for a specific purpose, so is not inconsistent with the Short-term Expedient use of stockpiled material. Higher quality material was scarce in both Clusters E and F, limited to just one small flake at the latter, and a single large flake and retouched piece at the former. It was difficult to make any conclusions on the basis of so few artefacts, but the large flake retained an original crystal facet, so came from a large crystal that could have been used as a transported core. The retouched piece was possibly broken which, in conjunction with the scarcity of artefacts, may indicate some level of Individual Provisioning. The surrounding, low-density material from outside the Clusters was more difficult to interpret. Stockpile-quality material was fairly heavily reduced (2.9–11 scars, 1.4–5 platforms, 0.82–1.3 g/scar, 5.76 mm/scar), which is more consistent with Place Provisioning. However, considering the low-density spread of the material (70 artefacts in 3900 m2), pieces probably preserved evidence of reduction that occurred elsewhere on site, rather than in the immediate vicinity of the discard location. Highquality material was scarcer and reduction intensity varied, but was generally slightly higher than the stockpiled assemblage (2.5–3.9 scars, 2–2.5 platforms, 0.37 g/scar, 4.39 mm/scar). Only two of the 13 large flakes were broken, and neither of the retouched pieces. Despite the limited data, this seems inconsistent with Individual Provisioning, so may represent some level of Gearing Up, or simply the use of higher quality material within a Place Provisioning strategy. 291 CA1 The sizeable assemblage of stockpile-quality material in CA1 preserved clear evidence of Place Provisioning throughout the entire occupation period (2500 cal. BP–recent past). Reduction intensity was consistently higher than at Gibb Rock, across all artefact types, when all XUs were grouped (avg. 2.1–7.2 scars, 1.7–3.6 platforms, 0.61–2.4 g/scar, 5.27 mm/scar); even when XUs were considered separately, reduction intensity always exceeded that of a Short-term Expedient strategy. There was considerable variation among individual flake scar and platform counts, as well as the amount of unmodified surface. Several cores had been flaked to exhaustion, while others retained usable material. This variation, in combination with higher average values, is indicative of the longer reduction sequences that characterise Place Provisioning. Higher quality material occurred in all but the basal unit, XU9 (2500–2158 cal. BP), but the provisioning systems varied over time. Gearing Up dominated in XU1 and 3– 4, while XU2 and 5–8 assemblages were more consistent with Individual Provisioning. In XU1 (144–0 cal. BP) and XU3–4 (1056–324 cal. BP), high-quality material was fairly abundant, but only large flakes (n=23) provided reduction intensity data. Average values were very similar to the stockpiled assemblage, but the flakes themselves were slightly smaller (avg. 3 scars, 1.9 platforms, 4.98 mm/scar). Discard behaviour showed no conservation of raw material at the tool/flake level, as breakage rates were incredibly low – only 15% (n=3) of large flakes (n=3) and 17% (n=1) of retouched pieces were broken. This lack of conservation is much more consistent with Gearing Up than Individual Provisioning. Three retouched pieces on stockpile-quality (from XU3–4) may also be part of this Gearing Up assemblage, if lower quality material was indeed suitable when artefacts did not have to be preserved indefinitely. Exotic raw materials were scarce, but the presence of three pieces in XU3 may indicate that some level of Individual Provisioning occurred at this time. Individual Provisioning is represented in XU2 and XU5–8, but these XUs preserve different evidence of the same strategy. In XU2 (324–144 cal. BP), the assemblage comprised one small fragment, nine small flakes (including two exotic raw materials) and five large flakes. A distinct crust was evident on the dorsal surface of all large 292 flakes, and many of the smaller flakes. Considering the absence of crusting on the ventral surfaces, it was clearly formed when the flakes were attached to the parent core. If the aim of reduction was to produce flakes for on-site use, one would expect a mixture of flakes, some with the crust (representing those with dorsal surface exposed during the crust-forming event), and others lacking it (those struck after the crust was removed). It seems more likely, then, that the flakes were produced when the core was trimmed, to remove the crusted surface and expose the unaffected material beneath. This indicates some desire to maximise utility of material per unit weight, by removing waste prior to transport. This is more consistent with Individual Provisioning, where greater mobility means that transport costs must be reduced wherever possible. Transported cores were an important part of Individual Provisioning, representing toolmaking potential that is otherwise lacking, considering the distance/time to next source is unknown; this reduces the risk of being without a certain tool when required. This same risk is absent on logistical forays, since they are shorter and target a specific range of resources before owners return to base camp, where raw material stocks are available. It is more efficient, then, for people to equip themselves with the tools required for the appropriate tasks, than to transport cores. XU5–8 (2158–1056 cal. BP) also preserved clear evidence of Individual Provisioning but, in this case particularly through selective artefact discard, which was primarily limited to small/unusable pieces and broken items. Large flakes were rare – only two were present, one of which was broken; the single retouched piece was also broken. The majority of the assemblage comprised small flakes (mostly unbroken) and fragments < 10 mm. These may have been produced by flaking small, transported cores, or from reworking existing artefacts. Finally, there were a greater number of artefacts made on exotic materials (n=7). These materials are more likely to be components of transported toolkits, since higher quality material flakes more predictably and prolongs artefact use-life. CA2 As in CA1, evidence of Place Provisioning was found throughout CA2. The stockpiled material comprised moderate to good quality quartz, as well as one piece of coarser grained silcrete. One core was recovered, which exhibited evidence of 293 extensive prior reduction (16 scars, 6 platforms), but still retained a good amount of usable material (23.15 g). Large fragments and large flakes also showed higher reduction intensity (2.4–3 scars, 1.8–2 platforms, 0.34 g/scar, 5.16 mm/scar) than would be expected under Short-term Expedient strategies, and a greater range of individual values, indicating longer reduction sequences. It is generally hard to detect Short-term Expedient use of stockpiled material among a Place Provisioning assemblage, since reduction intensity values are skewed by previous reduction events, and Place Provisioning creates far more archaeological debris. However, there was some evidence in the form of a unifacially retouched piece from XU4. This flake was fairly large and heavy (4.1 g), and the retouch comprised a single large flake removal from the ventral surface. If the aim was indeed to detach and use the new flake, rather than modify the form of the existing flake, it would seem inconsistent with Place Provisioning, as a new flake could be struck from the parent core, or other stockpiled material. It seems more consistent with the reuse of a previously discarded flake for a new purpose (in this case, as a core), which is more synonymous with a Short-term Expedient strategy. Nevertheless, Place Provisioning dominated throughout the entire occupation period, from 6700 cal. BP until the recent past. As in CA1, the character of high-quality assemblages varied over time. XU1–3 (2982–0 cal. BP) were dominated by evidence of Gearing Up. Large fragments and flakes exhibited lower reduction intensity than the Place Provisioning assemblage across most measures (avg. 1.3–2.3 scars, 1.3–1.7 platforms, 0.28 g/scar, 5.89 mm/scar), indicating shorter reduction sequences. Breakage rates were low – only two of six large flakes were broken, and one of nine retouched pieces. This is inconsistent with the conservation behaviour expected under Individual Provisioning. Four non-quartz pieces were found, as well as two small flakes that preserved evidence of more standardised reduction; these may indicate that some visits occurred under Individual Provisioning. In contrast to the upper deposits, XU4 (4792–2982 cal. BP) preserved clear evidence of Individual Provisioning. Reduction intensity was consistently higher than the Place Provisioning assemblage, across large fragments (3 scars, 3 platforms) and large flakes (avg. 3.3 scars, 2.7 platforms, 4.44 mm/scar). Five of the seven 294 large flakes exhibited evidence of standardised reduction, with more than one indicator present on several flakes. One retained long, narrow, parallel flake scars, while three flakes are long and narrow themselves. Another three have notches on their lateral margins, comprising regularly sized and spaced partial flake scars (distinct from retouch), created from previous flake removals when the standardised parent core was oriented differently (Figure 12.20). The fact that few large flakes were found indicates that much of this reduction occurred elsewhere, as would be the case under Individual Provisioning. One small flake also exhibited parallel flake scars, and another at 90 degrees to it, indicating standardised reduction and rotation of the parent core; the remainder of the small flakes are fine enough to have been produced by retouch. Both retouched pieces were broken, but only three of the seven large flakes were. However, the remainder may have been produced to rejuvenate the core, rather than to use the resultant flake, so their discard does not necessarily indicate a lack of raw material conservation. While greater reduction intensity could occur under Gearing Up, the selective discard and standardised reduction are much more characteristic of Individual Provisioning. XU5–6 (6700–4792 cal. BP) yielded much smaller assemblages on high-quality material; these comprised a single retouched piece, four small fragments and four small flakes (all whole). The retouched piece was not broken or heavily damaged; this is more consistent with Gearing Up, since functional pieces should not be discarded under Individual Provisioning. However, considering its extremely small size (0.05 g), it may represent artefact loss rather than intentional discard. Similarly, there was little evidence of primary reduction, since all small flakes and fragments could have derived from reworking existing artefacts or converting flakes into tools; the dominance of secondary reduction is more consistent with Individual Provisioning. Therefore, while the small sample size must be considered, the limited assemblage from XU5–6 is more likely a product of Individual Provisioning. 13.5 CONCLUSION Using the data provided in Chapters 10–12, it was possible to identify the various technological provisioning systems employed at the three studied sites. At Gibb Rock, assemblages were small and individual collections were often quite restricted. 295 A Short-term Expedient strategy dominated at a site-wide level, evident from the use of low-quality, local quartz, and lower reduction intensity measures. Provisioning systems were much more varied at Anderson Rocks and Mulka's Cave; different systems dominated at particular times, or in specific site areas. At Anderson Rocks, the first 3600 years were dominated by Individual Provisioning and Short-term Expedient strategies, evident from limited artefact discard and less intensive reduction of lower quality material. Around 4600 cal. BP there was a shift to Place Provisioning, which continued for the next 2700 years, but some Individual Provisioning and Gearing Up also occurred during this period. From 1870 cal. BP the dominant system shifted from Place to Individual Provisioning several times. The surface deposits were dominated by Place Provisioning, but a small assemblage on the outcrop indicated Short-term Expedient reduction of low-quality quartz from a local vein. In the Main Cave Area at Mulka's Cave, only the Column Sample provided a large enough assemblage to permit concrete identification of the technological provisioning systems in use. Individual Provisioning dominated in the earliest (6000– 4250 cal. BP) and latest deposits (1570–48 cal. BP), while Place Provisioning, often supplemented by Gearing Up, occurred between. On the Humps outcrop, the subsurface deposits indicated the simultaneous use of contrasting strategies, Individual and Place Provisioning, throughout the 1120-year occupation period. The surface assemblage only preserved evidence of Place Provisioning and Gearing Up. The latter was evident from extensive on-site reduction of high-quality materials, and the discard of unbroken, undamaged artefacts made on the same material. In the Camping Area, Place Provisioning dominated the surface and subsurface deposits throughout the entire 6700-year occupation period. Stockpiled materials were intensively reduced, and the variance in individual values indicated longer reduction sequences. Supplementary and contrasting strategies were identified in various deposits, based on assemblages made on higher quality raw materials: Individual Provisioning was more common in the earlier deposits, while Gearing Up occurred more frequently in the later assemblages. 296 CHAPTER 14 MOBILITY, SITE USE AND OCCUPATION INTENSITY 14.1 INTRODUCTION This chapter describes how Anderson Rocks, Gibb Rock and Mulka's Cave were used, and how site use changed over time. To begin, the resource and occupation model is used as a framework to interpret the technological provisioning systems identified in Chapter 13, considering both the dominant (Place Provisioning, Individual Provisioning) and supplementary systems (Gearing Up, Short-term Expedient). These data identify the season/s during which each site may have been visited, how different parts of the site were used, and the general level of residential mobility throughout the occupation period. Then, changes in occupation intensity over time are considered by identifying those periods of higher or lower intensity superimposed on the normal pattern evident within a selected pit; this can only be performed for dated deposits that yielded sufficiently large stone artefact assemblages. Finally, potential reasons for intensity shifts are identified, with particular links to water availability. 14.2 GIBB ROCK No samples from Gibb Rock were dated, but based on the limited depth of GR1, the cultural material probably dates to the last few hundred years; older material may be present in unsampled parts of the site. Further, the assemblages were too small to consider change over time. While limitations in intra-site sampling must be considered, the available evidence suggests that Gibb Rock was probably not a residential location, even in the short-term. There was no evidence of the Place Provisioning strategy that should accompany long stays, and no trace of toolkit maintenance that characterises Individual Provisioning – these renewal activities occur at residential sites where downtime is predictable and abundant (Clarkson 2007:15; Torrence 1983). The lack of Place Provisioning is not particularly surprising, given the absence of gnammas or rockholes that store water for longer periods. The only time the site could be visited for longer periods would be during winter, when surface water was freely available. As noted elsewhere (see Chapter 8.3.1), the 297 optimal response to more widespread water availability would be for people to spread out into the landscape, exploiting the areas that could not be visited during drier months. This increased mobility is incompatible with Place Provisioning. The lack of Individual Provisioning is more surprising, since it indicates that the site was not used as a residential base at all, even under a system of increased residential mobility. Short-term Expedient assemblages, such as those found at Gibb Rock, are created when a particular tasks could not be completed with the available tools. While this is often associated with Individual Provisioning, due to the use of a generalised mobile toolkit, it could also occur under Place Provisioning if an unanticipated task occurred away from the home base. It is therefore not possible to evaluate the general mobility of the Gibb Rock visitors based solely on the expedient assemblages preserved. It is only possible to conclude that the site was visited for very short periods, probably during foraging trips, and was not a residential location. If the predictive model is correct, these trips would have occurred in spring–summer, when people were based at a well-watered location nearby. This would explain the absence of Individual Provisioning assemblages, since the lack of water would make it impossible for people to camp onsite at that time. 14.3 ANDERSON ROCKS Pit AR2-3 indicates that Anderson Rocks was visited from 8200 cal. BP until the recent past. XU1–32 all yielded cultural material, indicating that there were no significant hiatuses in site occupation. However, it is impossible to address continuity of occupation at a finer scale with the available data. Nevertheless, it is possible to address questions of site use and changing occupation intensity over time, using the data outlined in Chapters 10 and 13. 14.3.1 Mobility and Site Use The technological provisioning systems identified in Chapter 13 provide important information on mobility, which can be combined with other indicators to determine when and how Anderson Rocks was used. Occupation evidence was clustered in two main site locations: most was in the vegetated area near the gnammas, while 298 smaller assemblages were present several hundred metres northwest, atop the outcrop. These site areas appear to have been used in different ways, so are discussed separately below. Gnamma area The depth of archaeological deposit in AR2-3 clearly indicates that this same area was regularly reoccupied over several thousand years. Considering the proximity to gnammas (approx. 50 m), this suggests that those structures provided the main source of water during these visits – otherwise, there would be no reason for occupation to be repeatedly tethered to this same location. It has been demonstrated that, as a long-lived water source, gnammas would not be used until more ephemeral sources (e.g. surface water, soil, pans) had been exhausted (see Chapter 8; Anderson 1984; Bird 1985; Gould 1968, 1969b; Tonkinson 1978:29). Therefore, it is unlikely that the bulk of occupation evidence was deposited in winter, or soon after, but instead when water was no longer widely available. Following the model derived earlier (see Chapter 8.3), Anderson Rocks should have been occupied in November and/or March–May, when the central study area was the optimal residential location. The gnammas would regularly hold water in November, but the precise quantity available depends on the amount of spring rainfall, which preserves the water balance by partially offsetting evaporation. At this time, occupation would be predictable, and extended stays would have been permitted under most rainfall conditions; this would favour Place Provisioning. Some visits could occur from March–May, but this would depend on the size and frequency of autumn rainfall events; these sporadic visits would favour Individual Provisioning. The evidence from AR2-3 and Scatter 3 indicated two fairly distinct occupation periods, with the transition between the two occurring around 4571 cal. BP, at the interface between XU21 and XU22. In the earlier deposits (8200–4571 cal. BP), Individual Provisioning was dominant, with some Short-term Expedient reduction of stockpiled materials. This indicates that visits were shorter and/or more sporadic, which necessitated the maintenance of a formal, transported toolkit, supplemented by expedient artefacts where required. While Individual Provisioning could suggest the site was visited in autumn rather than spring, it seems an unlikely explanation for 299 the 3600-year dominance of that strategy. It is more probable that a reduction in the normal water-holding period precluded long visits at this time, increasing residential mobility. However, the presence of stockpiled material as low as XU30 indicates that initial occupation may not have occurred under these conditions. Before 7775 cal. BP, the site must have been visited regularly enough, and for long enough periods, to warrant stockpiling material. It is therefore possible that there is additional occupation evidence (displaying the Place Provisioning signature) in the unsampled deposits, including those below XU33. The later period (4571 cal. BP–recent past) was dominated by Place Provisioning, but there were short periods where small sample size prevented the strategy being clearly identified. Nevertheless, in the last four and a half thousand years visits to Anderson Rocks must have occurred at predictable, regular intervals and for longer periods. As noted above, this could probably only occur during spring, as the gnammas' water balance declines swiftly from December. Several assemblages were characteristic of 'Gearing Up', which involves the manufacture of specialist tools before a logistical foray. This strategy is associated with longer site visits, when restricted mobility has limited access to certain resources, requiring an increase in logistical mobility. Gearing Up assemblages were particularly identifiable between 4571–3819 cal. BP (XU20–21) and 3170–1870 cal. BP (XU12–17), and in the recent past (Scatter 3). Despite dominance of Place Provisioning over the last 4500 years or more, there was evidence of the contrasting strategy (Individual Provisioning) within several of the same XUs (Figure 14.1). It was often difficult to separate the traces of Gearing Up and Individual Provisioning, but the latter was particularly evident in XU18 (3603– 3170 cal. BP), XU5–8 (1220–354 cal. BP) and XU1–3 (138–0 cal. BP). The intermixture of Place and Individual Provisioning could arise in one of two ways. It may indicate a short-term shift between the two strategies across a timescale short enough to preserve both in the same deposit – for example, a period of several years or more where longer visits were suspended, and Place Provisioning was replaced by Individual Provisioning. There is, however, an alternate possibility: that these contrasting strategies were complementary to a certain degree. If Place Provisioning reflects the normal, predictable occupation, then Individual Provisioning 300 may reflect site visits at other times of year. These out of season visits could occur after large isolated summer/autumn falls, when gnammas would hold water for a short period, allowing people to temporarily move further afield from their wellwatered refuges. Individual Provisioning was particularly common within the last 1200 years. This may indicate an increase in the conditions that facilitated out of season visits, or more frequent shifts in occupation patterns at this time. Figure 14.1 Technological provisioning systems (top) and occupation intensity (bottom) at Anderson Rocks, based on dated archaeological content from AR2-3 and surface material; the latter was presumed to date from the last 50 years or so. Where provisioning systems could not be confidently identified, these have been omitted. Outcrop Occupation evidence was less abundant on the outcrop itself, limited to assemblages from Scatters 1 and 2, and test-pit AR1. Scatter 2 preserved quite a distinct assemblage compared to others from Anderson Rocks, as artefacts were made on local raw material obtained from a narrow, exposed quartz vein. The entire artefact use-life, from quarry to discard, occurred at the raw material source. If people were camping on-site, at least some of these activities would probably occur 301 in other areas of the site, such as the main occupation zone. Similarly, the visit (and associated task) was possibly short-term enough that it was not worth collecting material from the stockpile, considering the availability of the local material. Therefore, the assemblage probably derived from people passing through, possibly on a foraging trip, while based at a site nearby. Scatter 1 and AR1 showed repeated occupation of the same small area atop the outcrop. AR1 was not dated, so temporal change can only be considered in terms of earlier and later deposits. The lower/earlier deposits (XU10–13) preserved evidence of Individual Provisioning, indicating shorter visits, and a higher degree of mobility at this time. This is unlikely to represent a widespread increase in mobility, based on the dominance of Place Provisioning in AR2-3 over the last 4500 years. Instead, it probably represents out of season visits to Anderson Rocks. It seems unlikely these visits would occur in winter (or any period of sustained rainfall), as runoff from the outcrop makes the area fairly damp and unpleasant. The upper deposits (XU1–9) exhibited less intensive reduction of stockpiled material. However, the presence of several retouched pieces on stockpile-quality quartz indicated the assemblages did not derive from a Short-term Expedient strategy. Instead, the area was probably used within the larger Place Provisioning system but in a different way. While the gnamma area was the focus of occupation, the outcrop itself may have been used for a more casual, restricted set of stone-working activities. Indeed, Scatter 1 preserved direct evidence of Place Provisioning and Gearing Up, indicating that longer visits had occurred on the outcrop in the recent past, at least. 14.3.2 Occupation Intensity Over Time Change over time could only be evaluated for AR2-3, as no other deposits were dated. However, its large assemblage size combined with the increased depth of cultural deposit means that periods of higher and lower intensity are readily identifiable. XU15–17, 18 and 20 all showed evidence of more intensive occupation, but in different ways. XU15–17 preserved a strong Gearing Up signature, which indicates that site visits were long enough to require logistical forays, for which specific tools were made. In addition, artefact discard rates were high, and there was a considerable degree of post-depositional flake breakage that can accompany more intense occupation periods. Most cores were incredibly small (≤ 1.6 g) bipolar, and 302 probably could not be flaked further – this indicates some degree of raw material conservation at the core level, which often accompanied longer visits. Artefact discard rates were also high in XU18 and XU20, and each exhibited varied individual flake scar, platform and unmodified surface values. These features are indicative of extended reduction sequences that accompany longer or more intense occupation periods. Further, all cores from XU18 were bipolar, and XU20 preserved strong evidence of Gearing Up. Less intense occupation is evident when Individual Provisioning or Short-term Expedient strategies dominate all deposits from a particular period. This indicates a shift from longer occupations (under Place Provisioning) to shorter and/or more sporadic visits. From 8200–4571 cal. BP (XU22–32), stockpiled materials were reduced far less intensively, and high-quality assemblages probably represented reworking of existing pieces rather than large-scale artefact manufacture. Artefact discard rates were also at their lowest in these deposits. As noted above, the presence of stockpiled material from 7775 cal. BP means that earlier or contemporaneous Place Provisioning episodes had occurred. Therefore, the short, sporadic visits preserved in the lower deposits do not reflect initial occupation of Anderson Rocks. It is also important to consider the reasons for interruption of the normal occupation patterns, specifically the less intensive occupation from 8200–4571 cal. BP and more intense periods from 4036–3819 and 3603–2520 cal. BP (Figure 14.1). As demonstrated above, regular use of the AR2-3 area was probably linked with the gnammas. As a result, the water preserved within them dictated the intensity of occupation that could be supported at any given time. Shifts in intensity would not be expected if gnammas were always capped at their peak, since they filled to capacity even under low rainfall conditions (Figure 10.2). Instead, intensity was probably influenced by the natural water-holding periods, which dictated the quantity of water available when people arrived on-site in late spring. As noted elsewhere (see Chapter 10.2.1), longevity of water was strongly influenced by the amount of spring rainfall. Strong spring falls frequently followed average or high winter rainfall, but rarely occurred when winter rainfall was low (BoM data). Therefore, lower intensity occupation may occur during low rainfall years, while higher intensity would only be 303 possible in average or high rainfall years when spring falls were strong. Any extension of the normal occupation period would be beneficial, as it would delay the move to the normal summer refuges; occupation was already tethered to those sites for long periods due to the scarcity of reliable water sources in the study area. 14.4 MULKA'S CAVE Mulka's Cave was visited from around 9650 cal. BP but most evidence post-dates 6700 cal. BP. Thereafter, the site was visited regularly until the recent past, but different parts of the site were used in different ways, and at different times. Changing intensity was especially evident in the Camping Area, where the pits were well-dated and yielded the largest stone artefact assemblages. 14.4.1 Mobility and Site Use Based on the artefact assemblages and technological provisioning systems identified therein, site use varied across Mulka's Cave. The Humps and Camping Area preserved similar evidence, so are considered together, while the Main Cave Area is treated separately. Site-wide data are then considered to evaluate potential evidence for inter-group gatherings at Mulka's Cave. Main Cave Area The Main Cave Area yielded the smallest assemblages, so technological signatures were often unclear. Nevertheless, artefact discard rates and other assemblage characteristics indicate that occupation was less intensive inside and around the rockshelter than in other parts of the site. It is important to note, however, that most occupation evidence from this area was out of context, so intensity may be underestimated by considering only the undisturbed portions of the assemblage. In any case, the rockshelter and the slope outside do not appear to have been the focus of activity at the site, contra to Gunn (2006). It may have been visited separately to, or in conjunction with, other parts of the site. The earliest deposits from the Column Sample (XU10–11, 6000–4250 cal. BP) inside Mulka's Cave preserved evidence of Individual Provisioning, indicating greater mobility at this time, at least in terms of visits to rockshelter itself. It is uncertain 304 whether the same signature is preserved in basal portions of the 1989 test-pit, but low artefact numbers may support this (Rossi 2014a). From 4250–1570 cal. BP Place Provisioning dominated, with some evidence of Gearing Up, which indicates that stays were more regular and longer, and that some visits were long enough to prompt logistical forays. Interestingly, it is during this period that the rock art was created, following Gunn's (2004, 2006) age estimate of 3000–2000 cal. BP. These longer site visits may have been associated with rock art creation, and indeed Gunn (2004, 2006) concluded that motif superimposition indicated that Mulka’s Cave had been visited regularly over the last few thousand years. However, the inferred link cannot be further evaluated with the available data. From 1570 cal. BP, Column Sample assemblages demonstrated a mix of Individual Provisioning and Short-term Expedient use of stockpiled material. At this time, then, greater mobility and shorter and/or less predictable visits again prevailed. While MC1 was not dated, the artefact-bearing layers (XU3–6) probably date to this same period. The assemblages indicated some Short-term Expedient artefact manufacture along with a formal strategy that could not be reliably identified, but had some features in common with Individual Provisioning. The uppermost deposits in the Column Sample and MC1 were both sterile, but surface artefacts, including one flaked glass piece, demonstrate that these areas were visited in the recent past, and that people were still using stockpiled material for expedient artefact manufacture. Camping Area and The Humps All three pits (CA1, CA2, TH1) showed repeated reoccupation of the same areas, but different time periods and activities were represented by each pit. The Camping Area was clearly the focus of occupation at Mulka's Cave, based on the quantity and spread of surface artefacts, as well as the stratified deposits yielding archaeological material. This fits well with the assertions made by Traditional Owners, who stated they used to camp there when they visited the site in their childhood. This, in turn, indicates some continuity in site use from the distant to the more recent past. The Camping Area was demonstrably in use during period the rock art was created but may also have preserved some more tangible links in some of its deposits. Ochre was found in three XUs deposited from 4792–1173 cal. BP, with most of the sample 305 post-dating 3000 cal. BP, when Gunn (2006) argued the rock art began to be created. However, ochre had a variety of uses, not all connected to art production. The Humps outcrop was also visited regularly, but for a far shorter period than other parts of the site. The lower reduction intensity (compared to the Camping Area) possibly indicates that it was primarily used as a satellite area for stone-working. This location may have been chosen for the excellent views it affords of the surrounding area, allowing occupants to monitor the movements of people, animals, rain clouds, or even bushfires. Throughout the Camping Area and The Humps deposits, Place Provisioning dominated from the earliest assemblages (6700 cal. BP) to the surface material, indicating that visits to the site had always been predictable and long enough to warrant stockpiling raw material. As a result, some inferences can be made regarding the main season of occupation. As noted elsewhere (Chapter 8.5.2), winter water abundance would favour Individual Provisioning, as people were free to move around the landscape and visit areas that were inaccessible during the drier months. Therefore, it would be suboptimal to stay at Mulka's Cave for long periods during winter. In contrast, it would be the ideal spot in September/October, when freedom was curtailed because all ephemeral sources of water had dried up, but the southern part of the study area was still the optimal foraging location. Similarly, while gnammas could be capped to preserve their water supply, the largest source – the cleft rockhole – could not. Due to its large surface area and sloping walls, it also lost water at a greater rate than gnammas. Considering the abundance (up to 6800 L) and ephemerality of this source, it would invariably have been used. It may have prompted some level of congregation, whereby people came together after winter to use this shorter lived supply before dispersing to the smaller, more reliable sources in summer. In this case, the dominance of Place Provisioning is not surprising. The rockhole provided a reliable source of water, even after low winter rainfall, which permitted long, predictably timed visits under all rainfall conditions. Several deposits also preserved evidence of Gearing Up, which results from restricted residential mobility and the need to travel further afield to procure specific resources. The strategy first appeared in CA2 XU3, but due to the slow rate of deposition, the date range was fairly broad (2982–1172 cal. BP). The basal XU of 306 CA1 dated to 2500 cal. BP, but Gearing Up assemblages were only present in deposits post-dating 1056 cal. BP. The strategy was well represented in surface material both in the Camping Area and The Humps; its absence in TH1 probably related to that part of the site being used in a different way, as noted above. Therefore, these longer visits, and the logistical forays they prompted, possibly occurred as early as around 3000 cal. BP, but became increasingly common in the last 1000 years or so, if the excavated deposits can be considered a representative sample. In contrast to Gearing Up, traces of Individual Provisioning were present throughout most of the 6700-year occupation period – only deposits from 2982–2158 cal. BP lacked clear evidence of the strategy (Figure 14.2). However, many individual XUs preserved no evidence of Individual Provisioning, and there were no deposits where this strategy was not accompanied by Place Provisioning. Nevertheless, Mulka's Cave was regularly visited by more mobile groups who relied on transported toolkits; such visits would have been short and less predictably timed. As noted above, the coexistence of Individual and Place Provisioning in the same archaeological deposit can indicate short-term shifts between the two strategies, or their use at different times of year. The former seems unlikely, considering the lengthy period over which both strategies are present, in contrast to the shorter periods of overlap at Anderson Rocks (Figure 14.1). Further, the rockhole would always preserve enough water for longer stays during the normal, post-winter occupation period, even in low rainfall years. It seems more likely, then, that Individual Provisioning represents visits outside the normal occupation periods associated with Place Provisioning. These visits may have occurred during summer or autumn as, if capped, the gnammas would have provided a reasonable source of water. This may have allowed people to temporarily move from their well-watered refuges in other parts of the study area, taking the pressure of the critical resources found there. Traces of Individual Provisioning were distinctly less common in the last 1000 years or so, particularly during the last 150 years (Figure 14.2). Despite the abundance of surface artefacts, only Cluster F preserved possible traces of the strategy. During this time, the predictable, longer visits still occurred, but out of season occupation was rare. It is possible that European arrival may have interrupted traditional occupation patterns, but this idea cannot be evaluated with the available data. 307 Figure 14.2 Technological provisioning systems (top) and occupation intensity (bottom) at Mulka’s Cave. Provisioning systems are shown for all dated pits and the combined surface assemblages, which were presumed to date from the last 50 years. CS = Column Sample. Where provisioning systems could not be confidently identified, these have been omitted. Occupation intensity is based on assemblages from the Camping Area. Evidence of aggregation at Mulka’s Cave Many have argued that Mulka's Cave had some ritual significance, often based on the quantity and diversity of its rock art, as well as particular motifs (Goode 2011; Gunn 2006; Macintyre et al. 1992; Webb and Rossi 2008). However, Rossi (2010:111–113) was unable to find any supporting evidence in the archaeological record, as assemblages from the Camping Area and Main Cave Area were fairly similar. While the aim of this research was not to re-address the question of ceremony at Mulka's Cave, it is worthwhile to consider the inferences that can be made based on the newly available data, specifically whether people gathered at the site, for ritual or social reasons. There are two main questions to consider – first, 308 whether the resource availability could support larger groups of people and, secondly, if the site preserves any evidence of this. As noted elsewhere (see Chapter 8.4), gatherings would be best held in the southern part of the study area, in December, after high winter rainfall (one year in ten). The timing and location would provide enough food and water for larger groups without impacting the survival of the regular occupants. Mulka's Cave fits the location criterion, but also provides a unique abundance of water. As demonstrated earlier (see Chapter 12.2.1), the cleft rockhole captures water easily, and stores it in great volume for a considerable period; this source could easily support larger groups of people. Under average or high rainfall conditions (eight years in ten), the structure held approx. 2900–4700 L of water at 1 December. Mulka's Cave would therefore permit greater freedom in terms of which years gatherings could occur. While large enough supplies would probably be available in low rainfall years, this water would be used for general occupation, due to reduced availability elsewhere (see Chapter 14.4.2). The rock art in Mulka's Cave began to be created around 3000–2000 years ago (Gunn 2004, 2006), but both the rockshelter and the Camping Area had been visited for thousands of years before this. Visits occurred on a regular basis both before and after the rock art creation, based on the evidence of Place Provisioning throughout the Camping Area deposits. Clearly, then, the entire occupation history cannot be linked to ceremonial events, if the rock art is considered indicative of ritual site use. However, the site still preserves some evidence that gatherings or ceremonial events may have occurred within the broader occupation scheme. Indeed Gunn (2006) argued that Mulka's Cave was probably used for a mix of general and ritual purposes. Kelleher (2002:270) claimed that ceremonial sites should preserve distinct activity areas, with the ritual areas separated from the main domestic areas; this should be visible in artefact density and assemblage characteristics. As demonstrated above, the Main Cave Area was used far less intensively than The Humps or the Camping Area. While Place Provisioning dominated elsewhere, the rockshelter assemblages more often displayed evidence of Individual Provisioning. This may indicate that it was used in a different way to the rest of the site, with less focus on day to day activities. However, the limited sample is problematic, and indeed the Column 309 Sample yielded some evidence of Place Provisioning from 4250–1570 cal. BP, so some level of general occupation did occur inside Mulka's Cave. It was also argued above that Individual Provisioning assemblages from the Camping Area probably indicated out of season visits. These visits may have occurred for ceremonial/gathering purposes; visitors may have camped in the designated area/s and maintained their transported toolkits as required. If gatherings did occur at Mulka's Cave, the general paucity of non-quartz materials may indicate that contact was primarily with other Noongar rather than desert groups from the east, who had more regular access to exotic materials (Figure 3.4). 14.4.2 Occupation Intensity Over Time Unlike Anderson Rocks, most deposits at Mulka's Cave were dated. Nevertheless, sample sizes were often too small to properly evaluate changing intensity over time. The Camping Area yielded the largest assemblages and, as the main focus of occupation, should preserve the best evidence of intensity shifts. At Anderson Rocks, periods of reduced occupation intensity were identified when Individual Provisioning or Short-term Expedient strategies dominated all deposits dating to a particular period. At Mulka's Cave, however, Place Provisioning was ubiquitous throughout the entire occupation period represented by the author's pits (Figure 14.2). While the lowest deposits in the Column Sample (XU10–11, 6000–4250 cal. BP) only preserved evidence of Individual Provisioning, contemporaneous deposit from the Camping Area indicate that Place Provisioning was in use at this time. In this case, the contrasting strategies relate to how different parts of the site were used rather than reflecting a large-scale shift in occupation intensity across the entire site complex. Only CA1 preserved evidence of increased occupation intensity, in XU3–4, that dated to 1056–324 cal. BP (Figure 14.2). Artefact discard rates were fairly high, and both deposits preserved clear evidence of Gearing Up. Negative flake scar and platform counts were their most varied in these XU, and 50–67% of cores were bipolar. All of these features indicate longer reduction sequences, which is a fairly reliable proxy for increased occupation span. Values were consistently higher and more varied in XU4 than XU3, and the former also preserved the largest number of artefacts with > 25% of their surface unmodified. Based on these data, site visits 310 were longer and/or more intense between 1056–324 cal. BP, particularly from 1056– 723 cal. BP. Interestingly, this overlaps with the period represented by TH1, where cultural material began to accumulate around 1120 cal. BP. This area may have been used as a response to increased intensity, prompting spatial segregation of some stone-working activities. At any site, longer and/or more intense visits are only possible when sufficient food and water resources are available. However, the capture properties and capacity of the cleft rockhole meant that, even under low rainfall conditions, more than 5700 L of water was available at the beginning of spring. As a result, the intensity of occupation was probably limited not by the availability of water, but by food resources, and the practical distance people can travel to forage and/or transport items to site. At some stage, depletion of surrounding food plants and animals will reduce foraging returns and increase transport costs – at this point, it becomes more profitable to move to a new location (Charnov 1976). Therefore, under average or high rainfall conditions, people would leave Mulka's Cave before the rockhole dried, since both water and more abundant food resources could be found elsewhere. However, under lower rainfall conditions, water would be scarcer and more patchily distributed across the group's territory, curtailing people's freedom to leave. The large capacity and ephemerality of the rockhole supply would instead encourage more intense occupations during the water-holding period. People would remain onsite for as long as possible (and possibly gather in larger numbers), despite lower foraging returns and increased transport costs. These longer stays would serve as a risk reduction strategy, by conserving water in other, more long-lived sources, until later in the season. These sources were already the focus of extended summer/autumn occupation within the normal seasonal route, so their conservation would be especially vital under low rainfall conditions. It seems plausible, then, the period of intense occupation at Mulka's Cave from 1056–324 cal. BP reflects the restricted availability of water in the wider area, which occurred under low rainfall conditions. 311 14.5 CONCLUSION The data collected earlier, in conjunction with the occupation model, allowed several conclusions to be made about how each site was used over time. Gibb Rock was visited in the more recent past, probably over the last few hundred years. The site was not a focus of residential occupation but was probably visited during spring/summer foraging trips when people were based at a location with reliable water. At Anderson Rocks, most occupation occurred when the gnammas held water, probably during spring. Visits were shorter in the earliest period (8200–4571 cal. BP), probably due to limited water availability, so people were more mobile and depended heavily on transported toolkits to meet their technological needs. For around the last four and a half thousand years, visits were longer, indicating that water was more abundant than during the earlier period. Occupation was particularly intense from 4036–3819 and 3603–2520 cal. BP, probably due to higher spring rainfall extending the water-holding period of the gnammas. The outcrop itself was used for stoneworking activities, but was also visited during foraging trips, when unexpected tasks were completed using quartz from an exposed vein. At Mulka's Cave, Place Provisioning occurred across the entire 6700-year occupation period represented by the author’s excavations. Visits were therefore predictable and long enough to warrant stockpiling raw material. This is unsurprising, as the cleft rockhole provided an abundant source of water, even under low rainfall conditions. The site was probably occupied in September–October, when water was becoming scarcer again, before people moved northwards to more optimal spring foraging grounds. Individual Provisioning was used throughout most of the occupation period, and may represent out of season visits, when capped gnammas would provide a reasonable supply of water, or people who travelled to the site for ritual or social gatherings. The Main Cave Area may have been the focus of those ceremonies, if indeed they occurred, but domestic activities were concentrated in the Camping Area. Occupation was more intense from 1056–324 cal. BP, when Mulka's Cave may have acted as a refuge, due to limited water availability elsewhere. 312 CHAPTER 15 CONCLUSION: OCCUPATION HISTORY OF THE STUDY AREA 15.1 INTRODUCTION This chapter ties together the different threads of this research to address the primary aim – to determine when and how people used the study area – and places these results in their broader context. Each of the four research questions are addressed in turn. The first focuses on the theoretical occupation model that was devised based on critical resource distribution and Optimal Foraging Theory, while the second defines the occupation period represented by the three excavated sites. The third considers mobility and occupation patterns revealed by the archaeological record, and how these fit within the theoretical model. The final research question quantifies temporal shifts in occupation intensity and relates them to water availability. Discussion then turns to the future research directions, before the main findings of the study are summarised, and its significance evaluated. 15.2 ADDRESSING THE RESEARCH QUESTIONS This research sought to answer one broad question: 'when and how did Indigenous people use the study area, and did the nature and/or intensity change over time?' This aim was addressed with reference to four specific research questions: 1. How were food and water distributed across the study area and, following Optimal Foraging Theory, how would seasonal and interannual resource availability influence human occupation patterns? 2. What is the timescale of Aboriginal occupation in the study area, and how does this compare to other parts of Australia? 3. What do the archaeological remains reveal about mobility and occupation patterns within the study area, and how do these results fit within the theoretical model? 4. How did occupation intensity vary over time, and can this variation be explained with reference to long-term shifts in water availability? 313 Each question is addressed separately below, with reference to the methods, source data and wider context and significance, as appropriate. 15.2.1 Theoretical Resource and Occupation Model This section addresses the first research question: ‘how were food and water distributed across the study area and, following Optimal Foraging Theory, how would seasonal and interannual resource availability influence human occupation patterns?’ To begin, the spatio-temporal distribution of food and water supplies are outlined for the study area. Then, the theoretical occupation model is discussed, focussing on the seasonal route and patterns of aggregation that would be permitted by the availability of critical resources. Finally, the results of the occupation model are placed in the context of similar data from southwestern Australia and other parts of the state. Distribution of plant and animal foods Plant foods were divided into seven categories, ranked by their return rates: gum and lerps/manna (high); flowers, fruit and storage organs (moderate); leaves/roots/stems and seeds (low). A total of 674 plant foods from 467 species were identified. Flowers were most numerous (n=252), but seeds (n=119) and gum (n=101) were also fairly common. Plant foods were most abundant in summer, which was the only season when all categories were available, including the two highly ranked ones; all storage organs were available, including those that consume their organ to reshoot after a period of dormancy, and fruit diversity also peaked. Food was scarcest in autumn, when highly ranked items were no longer available; moderately ranked foods were limited to a few perennial storage organs, flowers and fruits. Lower ranked items would have been more important at this time – leaves/roots/stems were available year-round, as were seeds from Hakea (n=30) and Allocasuarina (n=9). Plant foods became more abundant through winter to spring, as more species began to produce fruit, flowers and seeds; the entire suite of storage organs was again available by late spring. Overall, plant foods were generally the most diverse in Heath and Mallee. Saline yielded the most limited variety, but leaves/roots/stems were most diverse in this Division; storage organs were concentrated at Granite. However, species diversity is 314 not always indicative of the quantity of a particular food that may be available. For example, Thicket formations have fewer species than most other areas, but the three dominant genera all provide food items: Acacia provide edible seeds and gum, Allocasuarina have edible seeds that are available year-round, and Melaleuca yield edible flowers – these foods would be available in good supply. Similarly, lerps/manna derive from Eucalyptus, so would be abundant in Woodland and Mallee areas, where that genus dominates. Animal foods were categorised as follows: amphibians, birds, bird eggs, mammals, reptiles and reptile eggs; the latter only included eggs that were laid (rather than held inside the body), as they could be encountered independent of the animal. 142 animal species were identified, dominated by birds (n=89). Eggs were available from 85 birds and 23 reptiles. Temporal variation in animal foods was fairly subdued: amphibians, mammals and reptiles were present year-round, but their ease of capture and bodily condition varied. Female amphibians may have held eggs, mostly from March–June, while macropods (and possibly other herbivores) were in peak condition from June–September. Reptiles would have been difficult to catch in winter, when they are dormant, but more readily encountered when active during spring and summer; their eggs were available from September–December. Some migratory bird species visited the study area for short periods, but most birds were non-migratory. As a result, the minimum and maximum counts for a single Landscape rarely differed by more than ten species. Bird eggs were mostly available from August–October, and few were present in February–June. Overall, fauna was most diverse in Woodland and Mallee. Saline consistently yielded the least diverse assemblage, which was dominated by birds and their eggs. However, these would be available in good supply during winter and early spring, when migratory waterbirds travelled to the study area to breed. Large mammals were probably rare in Heath and Thicket areas due to the closely spaced vegetation. Distribution of water Generally, soil water was short-lived, and its presence centred on periods of sustained rainfall. Soils stored more water, for longer periods, after stronger winter rains. However, the distribution of falls was also important. If soils were dry for too long, the rainfall was often insufficient to recharge the profile – in those cases, Plant 315 Available Water (PAW) did not accumulate. Under low winter rainfall conditions, PAW was present in most soils for a few days to a few weeks. In contrast, Mallee soils never accumulated water, and Granite profiles held PAW until late winter, due to the additional input through runoff. Under average rainfall conditions, most soils retained PAW throughout most of winter; Mallee only held water intermittently and, along with Saline soils, was less likely to rewet once dry. Granite soils retained water until late October, and the 100 mm storage capacity was reached on several occasions. When winter rainfall was high, most soils held water in greater quantities and for a longer period, until early September; Granite soils held water until early November. Overall, Granite areas provided the best prospects for soil water, due to the influence of runoff shed from outcrops. The profile was frequently constrained by the upper storage limit, so benefitted from smaller falls spread over a long period, when capacity was less of a hindrance. Dry Granite profiles could accumulate PAW from isolated falls > 25 mm but would hold water for less than a week unless the falls were > 35 mm; such falls were comparatively rare (approx. 15% of all falls ≥ 15 mm). Heath, Thicket and Woodland soils held water for similar periods, while Mallee was the poorest performing profile – it only held PAW consistently under high rainfall conditions. Rock structures (pans and gnammas) were confined to the Granite Landscape Division. As they are closed systems, rock structures only lose water only by evaporation, so there was a simple relationship between the annual rainfall total and the number of days the structures held water. Falls were more effective in winter, when evaporation was lower, but water could accumulate relatively easily at other times of year. Pans held water for the shortest period, generally only during winter or other periods of sustained rainfall; water was only intermittently present when winter rainfall was low. Runoff pans fared slightly better, but still dried sooner than gnammas due to the storage limit imposed by their shallow depth. Nevertheless, they held water fairly consistently throughout winter, regardless of rainfall quantity. Both types of gnamma performed better than pans, especially under high rainfall conditions, as water could be stored over a greater depth and was lost more slowly. Runoff gnammas consistently outperformed gnammas and were only dry for 2–4 months of a year. Due to the larger water balances accumulated, they held water during long, rainless periods when other structures dried. 316 Due to the presence of rock structures, Granite was the only Landscape Division that regularly preserved water beyond early spring, even under high rainfall conditions. Structures would accumulate water after small, isolated falls, but this would generally last less than a week (Table 7.3; Figure 7.8). None of the structures held water beyond February, but this probably reflects the model limitations. Runoff gnammas were modelled fairly conservatively, by assuming they received twice the amount of rainfall. However, gnammas situated at the base of a slope, or at the end of a long inflow channel (high-input gnammas), would receive considerably more input. If these structures had the capacity to store this input, they could hold water for much longer periods. Occupation model Under OFT, higher ranked items should be pursued first, but lower ranked items will be incorporated when those are unavailable or limited in supply (Macarthur and Pianka 1966). Foraging patches can be similarly ranked, and those providing greater returns will be utilised first; where possible, people will move to a new foraging patch when resource depletion means that returns would be greater in a new patch (Charnov 1976; Macarthur and Pianka 1966). Finally, distance from camp dictates what resources can be profitably transported – higher ranked items can be transported greater distances before costs become prohibitive, while lower ranked items can only be moved over shorter distances (Orians and Pearson 1979). Considering these rules and the resource distribution data, the seasonal route below would be optimal under average rainfall conditions; variations under lower or higher rainfall are noted, where relevant. Summer: Larger groups would be tethered to Granite areas with capped runoff gnammas and/or high-input gnammas; group size would be limited by the volume of water available. Restricted residential mobility and predictable occupation location would favour Place Provisioning. Highranked foods were most abundant in Woodland, Mallee and Thicket areas, but people would also need access to a diversity of moderate- and lowranked plant resources and animals, since residential mobility was impeded. The optimal foraging patches would be in the northern study area, which is dominated by Thicket, and Mallee that preserves small pockets of 317 Woodland (Figure 3.14). Mallee and Woodland areas would have provided an abundance of fauna, and storage organs were readily available at Granite. Acacia seeds are low-ranked, despite their excellent nutritional content, but were abundant in Thicket. These could be profitably transported over short distances, so could be incorporated into the diet as required. After high winter rainfall (around one in every ten years), aggregation events could occur, ideally in December in the southern palaeovalley. Food would have been relatively plentiful, albeit less so than in other parts of the study area; water would also be available if gnammas had been capped after rains ceased. Importantly, increased consumption of critical resources would not affect normal summer occupation, since that occurred further north. Autumn: Water availability was still limited, probably restricted to highinput gnammas at particular Granite areas; Place Provisioning would still be the optimal technological system. Plant foods were far less abundant, and no highly ranked items were available; people would therefore need to broaden their diet and include more low-ranked resources (Macarthur and Pianka 1966). The central study area, where Thicket and Heath meet, would be optimal. Foraging returns would be highest in Heath and Mallee areas, where moderately ranked foods were most diverse, and Granite also preserved an abundance of storage organs. Fauna from Mallee and Woodland areas would be particularly vital during autumn, due to the more limited suite of plant foods available. If required, Allocasuarina seeds could be collected from nearby Thicket areas, but only if campsites were within fairly short distances. By mid-May, groups could begin to move to different Granite outcrops, as most rock structures begin to accumulate water from early rains. Winter: From late June, soil water was available in all Landscape Divisions, albeit intermittently in Mallee. Mallee was the only area that yielded diverse plant and animal foods during winter; plant foods were abundant in Heath, and fauna in Woodland, while migratory waterbirds were present in Saline 318 areas. The southern palaeovalley would be the optimal residential location, as it would provide access to all of these Landscape Divisions. As a response to widespread ephemeral water, people would disperse across the area in smaller groups, visiting areas that were inaccessible during the drier months. The increased mobility and unpredictable occupation location would favour Individual Provisioning, when technological needs are met by the use of a transported toolkit. However, by late winter, water was only available at Granite areas, as all other soils had dried. Under low rainfall conditions, this shift to Granite would have occurred sooner, while it could be delayed for several weeks under high rainfall conditions. Spring: By spring, water was only available in Granite areas, in soils and rock structures. Individual Provisioning would dominate in most cases, but some Place Provisioning may occur around larger rock structures where longer, more predictable visits occurred. Plant foods were more abundant, and reptiles were active after their winter dormancy. Heath, Mallee and Woodland areas provided the greatest diversity of plant/and or animal foods; many waterbirds were still present in early spring, but would probably be more common in the larger salt lakes in the eastern portion of the study area. Therefore, the southern palaeovalley would still represent the optimal foraging location. By November, most migratory waterbirds would have departed, so there would be fewer benefits associated with the southern part of the study area. People probably moved towards the central study area in preparation for a move further northward in summer. Under low rainfall conditions, small groups may have remained in the palaeovalley and spread out to utilise small supplies of water from capped gnammas, to reduce the strain on water sources in the central and northern study area. Conversely, if food resources were particularly abundant after strong winter rains, it may have been optimal for some groups to delay the northward move since, by summer, occupation would again be tethered to those few locations with reliable water. 319 Broader context and significance Anderson (1984) and Bird (1985) both reconstructed seasonal routes for different parts of southwestern Australia. Anderson (1984) believed that one group occupied the Swan Coastal Plain, and another the inland Darling Plateau. Reliable water and abundant aquatic resources allowed the Plain group to congregate on the coast in summer and autumn. In winter, they broke into smaller groups and dispersed along the plain and the forested area to the east; by late spring, people began to move back to the coast. She argued that inland groups had an irregular route and, since resources were less abundant, that these people were more mobile, and ranged over larger areas (Anderson 1984:34, 37). Bird (1985) studied a transect on the southern coast of Western Australia, inland to Lake Grace. She believed that water would dictate seasonal patterns of aggregation and dispersal. From February–May, people would have clustered in large groups around permanent, reliable water. From June–September, groups would fragment and spread out to use more ephemeral water. In October–January, intermediate water sources would be used, and camp sizes would vary according to the capacity of these sources (Bird 1985:176). The study area seasonal route is similar in several ways, but it occurred over a much smaller region than those considered by Anderson (1984) and Bird (1985). In each case, people dispersed during winter, as a response to less abundant food resources and more widely available water. There was some level of congregation in summer, but water sources in the study area probably could not support groups as large as those around more reliable, permanent water sources elsewhere in the southwest. Finally, the predictability of resources permitted a regular seasonal route, in contrast to the erratic occupation patterns that occurred in some parts of the Western Desert (Gould 1969b, 1991; Tonkinson 1978:29). In some parts of Western Australia, reliable water was available during seasons of food abundance, allowing different groups to gather for ceremonies or social events (Cane 1987; O'Connor and Prober 2010; Veth 1987). In contrast, unpredictable rainfall in the southern Western Desert meant that groups did not have regular seasonal routes, but instead followed the rain. Food could not be stockpiled for events, since it could be many years before groups were able to return to a particular location. Instead, events were held after strong rains, when water, plant and animal 320 foods were all readily available. As a result, these aggregations were held very infrequently (Gould 1968,1969b, 1984, 1991). The same was true in the study area, where aggregations would best be held the summer after strong winter rains, which occur around once every ten years. Water could be preserved in capped gnammas, to accompany the bountiful food supplies available in summer. Since summer gatherings were predicated on winter rains, and a location having the required structures to preserve this water, the location and timing of events was more predictable than in the southern Western Desert. As a result, some food could be stockpiled, including lower ranked resources that might be collected via embedded procurement, in a manner akin to flakeable stone. 15.2.2 Timescale of Aboriginal Occupation This section addresses the second research question: 'what is the timescale of Aboriginal occupation in the study area, and how does this compare to other parts of Australia?' This question is addressed in two parts, first by defining the occupation period, by identifying the oldest and most recent cultural material, with reference to the archaeological investigations at Anderson Rocks, Gibb Rock and Mulka's Cave, as detailed in Chapters 10–12. The second section relates to the wider context of these results, specifically how they compare to those from other parts of Australia, whether they can be considered representative of the study area as a whole (i.e. whether earlier material may be present at other sites), and why these data are significant. Timing of occupation Both dated sites (Anderson Rocks and Mulka's Cave) exhibit evidence of early Holocene occupation. Mulka's Cave preserved the earliest cultural material in the basal unit of the pit excavated by Bowdler et al. (1989); the author subsequently dated charcoal from various units (Rossi 2014a). The depth-age curve indicates that artefacts began to accumulate around 9650 cal. BP (Figure 12.9), but the earliest directly dated sample was deposited more than 1500 years later (1988 pit XU16, Wk-29188: 8166–7933 cal. BP at 95.4% – Chapter 12.5). Most occupation evidence was found in the Camping Area, and post-dated 6700 cal. BP (Figure 15.1). Considering the vastly different occupation periods represented by CA1 and CA2, 321 despite similar depths of deposit, it is possible that older cultural material is present in unsampled deposits of the Camping Area. Anderson Rocks demonstrated a similar antiquity of human occupation. In pit AR2-3, the lowest artefact-bearing layer (XU32) dated to around 8200 cal. BP (Figure 10.6). This early occupation evidence showed that people were meeting their technological needs through Individual Provisioning and, soon thereafter, Short-term Expedient flaking of stockpiled material. Stockpiling, or Place Provisioning, was the optimal solution when people knew they would be at a particular location regularly, and for long periods. Clearly, then, Anderson Rocks had been visited under such conditions before 7775 cal. BP. The site may preserve more occupation evidence below the culturally sterile XU33 or in other parts of the site; these deposits may be considerably older, or broadly contemporaneous to those from AR2-3 XU32. Figure 15.1 Occupation periods represented by subsurface archaeological deposits at Anderson Rocks (dark blue), Gibb Rock (red) and Mulka's Cave (light blue). CS = Column Sample, P88 = 1988 pit excavated by Bowdler et al. (1989). GR1 was not dated but based on the short depth of deposit the assemblage is probably no more than a few hundred years old. Note the presence of cultural material throughout the occupation periods, indicating no significant hiatuses in site visitation. It was also important to place some timeframe on the most recent occupation periods, particularly to determine whether occupation continued after European arrival. CA1, TH1, the Column Sample, and AR2-3 all yielded direct dates from 138– 48 cal. BP for XUs with basal depths of 75–150 mm (Tables 10.3, 12.2–12.3). Based on the deposition rates, much of the upper and surface material probably dates to the European period. Gibb Rock was not dated, but the assemblages probably date to within the last few hundred years – most material was found on the present ground surface, so some overlap with the European period is likely. 322 Rossi (2010) identified ten flaked glass pieces from Mulka's Cave. These constitute direct evidence that the rockshelter and the slope outside were used after European arrival. Only one of these glass flakes derived from the undisturbed collections that were reanalysed herein, so the differences in artefact typology and identification means the presence of the remaining nine pieces must be interpreted with caution. The same applies to flaked glass found in surface assemblages elsewhere in the study area (e.g. Quartermaine 2000 – see Chapter 4.2.1). Nevertheless, the evidence clearly indicates that occupation occurred after European arrival, and on a regular basis, contrary to the statements of early settlers (e.g. Meeking 1979; Mouritz 1986). Broader implications and significance While the initial colonisation date, route and conditions are still hotly debated, there is fairly reliable evidence that Aboriginal people had reached northern Australia by 65,000 cal. BP (Clarkson et al. 2017). People had certainly reached southwestern Australia by around 45,000 cal. BP, possibly as early as 50,000 cal. BP, if the more contentious dates from Devil's Lair are accepted (Allen and O'Connell 2014; Balme 2014; Pearce and Barbetti 1981; Turney et al. 2001). Most parts of the country – including the present-day arid zone – have sites pre-dating 40,000 cal. BP (e.g. Allen and O'Connell 2014; McDonald et al. 2018; Smith 2013:79–80). Clearly, then, if the three sampled sites can be considered representative, the study area has been occupied for a small portion of the continent's 65,000-year human history. The late occupation of the study area possibly comes down to water availability – all > 30,000 BP sites in the arid zone are associated with some sort of reliable water source (Bird et al. 2016; Smith 2013:78–91). In the study area, freshwater lakes and rivers are absent, and the groundwater saline/hypersaline. The only source of potable water was rainfall preserved temporarily in soil profiles and rock structures. Considering the ephemerality of water under modern rainfall conditions, it is unlikely that widespread occupation could occur during the drier conditions that characterised the last glacial period, beginning around 32 ka and terminating at the early Holocene (Petherick et al. 2013; Reeves et al. 2013). Therefore, while the present sample is limited, it seems unlikely that any Pleistocene sites or cultural deposits would be found in the study area, or the inland southwest. If present, they would probably 323 indicate short-lived incursions into the area, when climatic conditions permitted, rather than more regular, permanent use of the area. Despite the relative recency, the early Holocene dates from Anderson Rocks and Mulka's Cave are significant in the context of the inland southwest. As noted earlier (see Chapter 1.2), the area is vastly underrepresented in the national database of dated archaeological sites compiled by Williams and Smith (2012, 2013) and Williams et al. (2008, 2014). Prior to this research, Mulka's Cave was the only site in the area that provided radiocarbon determinations in association with cultural material. However, during their geomorphological analysis at Bennett Lake, Suzuki et al. (1982:70) claimed to have found artefacts in a unit they believe was deposited between 7500 and 4500 BP. However, when Bird (1985:111–113) visited the site, she was unable to locate any artefacts in undisturbed deposits. Based on geomorphological evidence, Bird (1985:115) believed that the surface scatters from her Inland Salt Lake Zone were exposed by the erosion of mid- to late Holocene units, so all assemblages post-dated 5000 BP. Mulka's Cave and Anderson Rocks, then, represent the first reliable evidence for early Holocene occupation of inland southwestern Australia, and provide clear evidence of its ongoing use until the European period. The Mulka's Cave data certainly refute the claims by Bowdler et al. (1989) that the site was only visited over the last few hundred years. 15.2.3 Mobility and Occupation Patterns This section addresses the third research question: ‘what do the archaeological remains reveal about mobility and occupation patterns within the study area, and how do these results fit within the theoretical model?’ This discussion is divided into three sections. The first comprises site histories for Gibb Rock, Anderson Rocks and Mulka's Cave, focussing on how Kuhn's (1995) technological provisioning systems indicate different levels of residential and logistical mobility. Then, the occupation model is evaluated against the results from the three excavated sites as well as data from other archaeological locations in the study area. The final section summarises the general mobility and occupation patterns from the study area and addresses the broader relevance of various conclusions. 324 Site histories At Gibb Rock, poor-quality local quartz was reduced in a Short-term Expedient manner; the resultant artefacts were used to supplement the transported toolkit to complete unanticipated tasks. Reduction intensity was low, and the assemblage was small, less complex, and lacked evidence of toolkit maintenance. This indicates that site visits were short, and no down-time was available (Clarkson 2007:15; Torrence 1983); Gibb Rock was therefore not a residential location. Instead, it was probably visited during foraging trips, while people were based at a nearby site. Since Gibb Rock was not a residential site, it is not possible to consider the level of mobility of the groups who visited. At Anderson Rocks, some cultural material was present on the outcrop, but most was concentrated around the gnammas that acted as a tether and led people to reoccupy the same area over more than 8000 years. These residential visits probably occurred during spring/early summer, when the gnammas held water that was not widely available elsewhere. Mobility varied over time – the earliest period began at unknown point in the past when visits became long and predictable enough to prompt stockpiling of raw material. From 8200–4571 cal. BP, Individual Provisioning and Short-term Expedient reduction of this stockpiled material indicates that visits were shorter. From 4571 cal. BP, visits became longer and more predictable, evident from more intensive reduction of stockpiled material, characteristic of Place Provisioning; this may indicate an increase in the gnammas water-holding period. Gearing Up assemblages indicate some increase in logistical mobility during the later period, at 4571–3819 cal. BP, 3170–1870 cal. BP, and during the recent past. Longer visits may have been more common at these times, requiring logistical forays to procure more distant resources. The coexistence of Place and Individual Provisioning assemblages, particularly over the last 1200 years, may indicate out of season visits (probably after summer/autumn isolated falls) and/or short-term shifts between contrasting strategies. Some satellite stone-working activities occurred atop the outcrop at Anderson Rocks. While these deposits were not dated, they likely accumulated over the last few thousand years; the longer occupation span during that period may have prompted some spatial segregation of activities. A small surface scatter indicates that the site 325 was also visited during foraging trips, when local quartz was used to manufacture expedient artefacts. The task must have been very short and easily accomplished, considering raw material stockpiles were available elsewhere on-site. The earliest occupation evidence from Mulka’s Cave derives from the rockshelter, where cultural material began accumulating around 9650 cal. BP. However, all available assemblages from the rockshelter post-date 6000 cal. BP. These are dominated by Individual Provisioning and Short-term Expedient strategies, despite the presence of Place Provisioning in contemporaneous deposits in other parts of the site. Longer visits occurred from 4250–1570 cal. BP; some may have been linked with rock art creation, which Gunn (2004, 2006) estimated began around 3000 years ago. The rockshelter was used less intensively than other parts of Mulka's Cave, but it was not possible to determine if it had any ritual significance as claimed by Goode (2011) and Macintyre et al. (1992). However, the Main Cave Area assemblages do not support Gunn's (2006) idea that the rockshelter and slope were the main focus of occupation, but the present analysis excluded the disturbed deposits that yielded most of the cultural material from that area. While the 420 BP date obtained by Bowdler et al. (1989) was clearly erroneous, their idea of intermittent visitation is broadly representative of much of the occupation period. The Camping Area was the focus of residential occupation, from 6700 cal. BP until the recent past. This indicates some continuity from the distant to the more recent past, when Traditional Owners camped in this same area when brought to site as children. Place Provisioning occurred throughout the deposits, so visits were long and predictable enough for raw material stockpiling to consistently represent the optimal technological strategy. This predictability, combined with the proximity of the Camping Area to the gnammas, indicates that visits were tied to the water-holding period of the rock structures. The cleft rockhole provided an abundant but ephemeral source of water under all rainfall conditions and may have been particularly valuable from mid–late spring when water was less widely available. Individual Provisioning occurred throughout most of the deposit – considering the reliability of the water sources, it probably represents out of season visits. Gearing Up assemblages all post-date 3000 cal. BP, indicating increased logistical mobility thereafter. 326 From 1120–1056 cal. BP three interesting patterns emerged, and persisted until the recent past: Gearing Up assemblages were more common, Individual Provisioning was used less frequently, and the Humps outcrop was first used as a satellite stoneworking area (Figure 15.1). This was probably a combined response to increased occupation intensity from 1056–324 cal. BP. Increased intensity would require more regular logistical forays to offset resource depletion, and out of season visits may be discouraged by the already lengthy period of tethered occupation. Finally, more intense visits may have required some spatial organisation, prompting an area to be dedicated to particular stone-working activities. Evaluating the occupation model While the sample is somewhat limited, is possible to evaluate the model against archaeological results from this research, and those of others. Briefly, the model summarised above (see Chapter 15.2.1) predicted that Place Provisioning would dominate in the central and northern study area, around reliable water sources. In the southern area, Individual Provisioning would be more frequently used, since residential mobility was greater in winter and part of spring, but some Place Provisioning might occur around certain rock structures that were the focus of occupation in mid- to late spring. Superficially, the studied sites fit this pattern. Realistically, however, the site histories summarised above are sensical without recourse to the model, based purely on the availability of water. Occupation was longer and more predictable in locations with more reliable water sources (Anderson Rocks, Mulka's Cave), while the site without water (Gibb Rock) was not used as a residential location. However, archaeological remains from the rest of study area permit the model to be further evaluated. Of the 17 other archaeological locations in the study area, 16 are in the southern palaeovalley. This may indicate widespread winter dispersal predicted by the model, but it is important to note that most were identified during cultural heritage surveys that accompanied development work, which is understandably more common near the Hyden township. It is difficult to place these archaeological locations in context, as none have been excavated or dated, and stone artefact data are very limited. Where available, these data can be difficult to integrate due to methodological 327 differences. As a result, discussion is limited to those sites that can be evaluated in context of the model or archaeological data collected during this research. Winmar (1996:28) claimed that Wave Rock was a place where various Noongar groups gathered. It is difficult to evaluate this idea based on the scant archaeological data, but some elements from the site complex may support the notion: the scarred trees may have denoted the special nature of the site; the single grindstone may indicate that low-ranked seeds were used to support larger groups of people; gnammas were present; and the site is in the southern part of the study area that the model identified as optimal. However, in the absence of limited artefact data, no concrete conclusions can be made. Four sites preserved gnammas: Wave Rock Rockholes (21386), Hyden (5844), Graham Rock (5610) and Twine Reserve (5121). The latter is in the central study area, while the remainder are in the southern palaeovalley. No artefacts were found at Graham Rock or the Wave Rock Rockholes, and artefacts were scarce in the wider area surrounding the latter (see Chapter 4.2.1). A few quartz fragments were present at Hyden (5844), which also preserved a gnamma > 1.5 m deep. While the volume of water captured would depend on the amount of runoff received, this source would preserve water for long periods, due to its depth. The paucity of cultural material at the southern gnamma sites supports the idea that this part of the study area was primarily used in winter, when surface water was more abundant. In contrast, Twine Reserve preserved extensive surface artefact scatters made on nonlocal quartz – probably indicative of Place Provisioning – and likely retains subsurface archaeological deposits. Timms (2013) noted seven gnammas at the site; the largest three range from 0.8–1.95 m in depth and hold almost 10,000 L of water between them. Based on the author's observations, these gnammas would collect runoff from a considerable expanse of rock. Twine Reserve would have been an ideal spot to camp during the drier months, possibly explaining the increased archaeological visibility; this also supports the model's predictions. Staub and Taylor (1999) investigated four salt lake sites: Kondinin Pipeline (KP)-1, KP-2, KP-3 and KP-5. Most artefacts were found around the lake edges, indicating these areas were primarily visited when the lakes were full, as predicted by the model. However, the presence of some artefacts on the lakebeds indicates some 328 limited use at other times of year. Stone artefact data are limited, but certain assemblage characteristics are of interpretive value. The assemblages were dominated by 'debris', which are probably analogous to the author's flaked fragments. In KP-5, the largest artefact sample studied, 59% of quartz artefacts were debris; similarly high values were present in most other samples. In AR2-3, fragments comprised 66.3% of the stockpile-quality assemblage, and only 33.7% of artefacts made on higher quality material. It seems likely, then, that most of the KP fragments were made on lower quality raw material. As a result, either Place Provisioning or a Short-term Expedient strategy must have been in use. Artefact totals are not available for any sites, but sample sizes indicate > 164 artefacts were found at KP-5, and > 390 at KP-1; artefact density was often greater than in the Camping Area clusters, and total artefact numbers were higher than in the entire Camping Area Scatter. However, it is unlikely the KP assemblages represent Place Provisioning, due to the infrequency of cores (0–3.7% compared to 5% in AR2-3 XU12–18), their uniformly small size (< 30 mm in maximum dimension), and the use of flakes as cores. This level of reduction and recycling would not be necessary under Place Provisioning or Short-term Expediency, since raw material would be readily available from a stockpiled or local source. Instead, the assemblages may represent a more casual form of stockpiling, whereby a limited quantity of material was accumulated during a particular visit; this material was used during subsequent visits, and discarded flakes/cores were flaked again rather than more material being collected. Therefore, as predicted by the model, occupations were probably not long or predictable enough to warrant regular stockpiling of raw material. The presence of a raw material source, in the form of previously discarded items, may have acted as a tether for future visits, so may explain why artefacts were clustered in particular areas along the shorelines. However, this increased archaeological visibility may also indicate that people sometimes gathered in greater numbers than predicted by the model, to exploit the seasonal abundance provided by waterbirds. Implications for mobility and occupation patterns throughout the study area In the study area resources were predictable, but not especially abundant (particularly water), so occupants probably lived in small groups as in other inland 329 areas (Hammond 1933:21, 23; Roe 1836; Smith 1993:287–288). It is not possible to evaluate territory size, but it would have been larger than in more resource-rich areas, and some boundaries may have been semi-permeable (Dyson-Hudson and Smith 1978; Harpending and Davis 1977). Overall, mobility was dictated by the availability of water – winter abundance allowed people to spread out, while the drier months tethered occupation to the few more reliable water sources; the length of occupation at individual sites shifted according to the availability of water on-site and elsewhere in the landscape. This created a system characterised by both scheduled/predictable and unscheduled/unpredictable moves. The former occurred when people travelled to a known water source, while unscheduled moves may occur during winter, or if large isolated falls permitted people to temporarily leave their summer/autumn camps. As a result, the same sites could be visited at several different times of year, and possibly for different reasons – while Mulka's Cave may have had a ceremonial function, general occupation clearly occurred there also. Certainly, the entire occupation and mobility system was more predictable than in parts of the Western Desert, where erratic and localised rainfall precluded normal seasonal routes (Gould 1969b, 1991; Tonkinson 1978:29). This regular, consistent use of the study area is in stark contrast to statements by early settlers, who claimed that people simply passed through on their way elsewhere (Meeking 1979; Mouritz 1986), or Landor’s and Lefroy's (1843) Aboriginal guides, who claimed that food and water were not available. The latter probably reflects a risk avoidance strategy, as the guides were not familiar with the distribution of food and water resources in the study area. While the bulk of mobility data derived from technological provisioning systems, raw material types also provided some insights. High-quality exotic raw materials (BIF, chalcedony) were only found at Mulka's Cave and Anderson Rocks; these materials probably derived from the greenstone belt beyond the eastern boundary of the study area, approximately 80–85 km from the sites (Figure 3.4). These materials were rare, but were more commonly found in deposits that also preserved evidence of Gearing Up (n=12/22 in AR2-3; n=3/8 in CA1, n=3/3 in CA2 – Appendices E and G). Considering the distance from source, it is unlikely that these exotic materials were collected during logistical forays to procure food resources. There are two possible explanations: first, that tools made on high-quality exotics formed part of the 330 permanently retained toolkit, and that maintenance activities occurred before a standard logistical foray along with the manufacture of other required artefacts on high-quality quartz. Alternatively, some quartz Gearing Up assemblages may reflect preparations for long travels for other reasons, perhaps social gatherings with people from further east. Small quantities of exotic materials, or artefacts, may have been transported back to the base camp and subsequently maintained. Regardless of how high-quality exotics fit into the broader technological scheme, and the extent of logistical mobility implied, the presence of these materials indicate that study area occupants had some access to the greenstone belt itself, or the people who lived there. The limited quantity of material means this access or contact was infrequent. It seems unlikely, then, that the greenstone belt was part of the study area group's territory or normal seasonal route. Contact with these people likely increased after 4000 cal. BP, when most of the exotic artefacts were discarded. If Tindale's (1974) cultural boundary is accurate, this may explain the general paucity of the high-quality exotics. While Noongar and desert groups had some social contact, it was infrequent (Macintyre et al. 1992). If this reasoning is correct, it indicates that the boundary separating Noongar from the desert groups had been located west of the greenstone belt, in the approximate position mapped by Tindale (1974), for the entire occupation period represented by the Anderson Rocks and Mulka's Cave deposits. This is in direct contrast to Gibbs and Veth's (2002) assertion that the boundary attained its present position only after European arrival arrested its westward progression. However, it is a very small sample of artefacts on which to base such an argument, which also relies on the exotic material having originated from the greenstone belt (or another area east of the study area), rather than an alternate source within Noongar territory. Nevertheless, if this theory is correct, and gatherings did occur within the study area, these must have been primarily with other Noongar groups, or people from areas where quartz was the dominant raw material used for artefact manufacture. 15.2.4 Occupation Intensity and Water Availability This section addresses the final research question: ‘how did occupation intensity vary over time, and can this variation be explained with reference to long-term shifts in water availability?’ The discussion begins with a consideration of the shifts in 331 intensity evident at each site, and how they might be dictated by water availability. While it was not the aim of this research, the collected data allowed Rossi's (2014b) earlier idea to be evaluated, namely the use of occupation intensity as a proxy for effective rainfall. These intensity data were used to construct an effective rainfall record for each site, which is then compared and refined into a single record for the study area and evaluated against the palaeoenvironmental record. Different patterns of occupation intensity were evident at each of the three studied sites based on the use of Individual or Place Provisioning and the presence of highintensity indicators (e.g. post-depositional flake breakage, longer reduction sequences). Intensity was consistently low at Gibb Rock, where Short-term Expedient strategies were ubiquitous. The deposits were not radiocarbon dated, but the assemblages probably date to the last few hundred years. Water was only available during winter, at which time Individual Provisioning was the optimal solution to permit the greater mobility required for people to access areas that could not be visited during the drier months. Therefore, it is unlikely any different pattern of intensity would be evident at Gibb Rock, even if older occupation deposits were found. At Anderson Rocks, low intensity occupation occurred from 8200–4571 cal. BP, when Individual Provisioning dominated. Thereafter, intensity was moderate, with bursts of higher intensity occupation lasting from a few hundred (4036–3819 cal. BP) to over 1000 years (3603–2520 cal. BP). A brief low-intensity period occurred from 138 cal. BP, but considering the abundance of Place Provisioning assemblages on the surface, this period probably only lasted around 75 years or so. The gnammas could only store < 600 L of water, so people would have stayed as long as was possible to reduce the length of time spent at their normal summer refuges, which were the focus of occupation for much longer periods. Therefore, the extreme variation in occupation intensity was linked to the natural water-holding period of the gnammas, which dictated the quantity of water available by the time people arrived on-site. Considering this, occupation intensity was linked to rainfall quantity as below: o Low intensity = low rainfall; o Moderate intensity = average or high rainfall; o High intensity = average or high rainfall, strong spring falls. 332 At Mulka's Cave, moderate occupation intensity dominated throughout the 6700-year occupation period represented by the author's excavations; this was probably due to the reliability of the cleft rockhole water supply. One period of higher intensity was identified, from 1056–324 cal. BP. The rockhole never held < 5000 L by spring, so food resources would invariably be exhausted while it still held water. It would be suboptimal to remain on-site thereafter, unless water was not widely available elsewhere – in that case, Mulka's Cave could be used as a refuge. Therefore, occupation intensity at Mulka's Cave responded to the availability of water in the wider area. This can be linked to effective rainfall as below: o Moderate intensity = average or high winter rainfall; o High intensity = low winter rainfall. A rainfall record was produced for each site based on the inferred response of occupation intensity to varying water availability (Figure 15.2). When these records are compared, several conflicting periods are evident. From 8200–4571 cal. BP and 138–50 cal. BP, low rainfall was indicated by the Anderson Rocks record, while the Mulka's Cave data identified these as periods of average–high rainfall. Intensity at Anderson Rocks was dictated by the availability of water on-site, while Mulka's Cave occupation responded to how abundant it was in the wider area. Further, the Anderson Rocks gnammas were more susceptible to drying than other structures, due to their limited capacity and conical shape. Therefore, rainfall may have been low enough to preclude long visits to the site, but not so low that Mulka's Cave had to be used as a refuge. From 1056–324 cal. BP, the Mulka's Cave record indicated low rainfall, while Anderson Rocks data implied average–high rainfall. At Anderson Rocks, both Individual and Place Provisioning assemblages date to this period. The latter were smaller than elsewhere in the pit, while the former were more readily identified than in most other XU. The dominance of Individual Provisioning probably indicates that low rainfall conditions prevailed for most of the period, while the small Place Provisioning assemblages may represent short periods of increased rainfall, when longer visits could occur. The latter would not be visible at Mulka's Cave, as such evidence would be numerically overwhelmed and thereby masked by the highintensity signature. 333 Figure 15.2 Occupation intensity (top) for each site and effective rainfall conditions for dated sites (middle) and the study area as a whole (bottom). Note the Gibb Rock assemblages were presumed to date to the last few hundred years, and all surface assemblages to the last 50 years. Considering the data presented above, it is possible to combine the Anderson Rocks and Mulka's Cave records to produce a higher resolution, more refined rainfall record for the study area. Nevertheless, due to the limitations inherent in each record, there are periods where it is not possible to quantify effective rainfall precisely, but merely narrow the range of possible conditions. Five rainfall conditions could be identified, by comparing and combining the records: 334 o Very low: high-intensity indicators at Mulka's Cave; o Very low–low: Individual Provisioning dominant at Anderson Rocks, no data from Mulka's Cave; o Low: Individual Provisioning dominant at Anderson Rocks; o Low–high: unclear Anderson Rocks signature, no high-intensity signatures at Mulka's Cave; o Moderate–high: moderate intensity at Anderson Rocks and/or Mulka's Cave. The combined rainfall record indicates that the mid-Holocene was dominated by lower rainfall conditions (Figure 15.2). For the next 2700 years, average or high rainfall conditions persisted, with two periods of probable higher spring falls (4036– 3819 and 3603–2520 cal. BP). During the last 1000 years, drier conditions again dominated, but rainfall temporarily increased for short periods. While the above pattern is interesting, it is difficult to integrate into the wider palaeoenvironmental record for southwestern Australia, primarily due to methodological limitations. The author had previously considered how occupation intensity might be used as a proxy for effective rainfall (Rossi 2014b), but the current research was not designed to test that theory. At best, it was hoped that the collected data might allow the idea to be further developed. As a result, the methods here were not designed to test rainfall variation over time, but instead to consider how different rainfall conditions (as represented by modern data) might have affected water availability and human occupation at various sites. There are three particularly significant points of difference between this research and palaeoenvironmental studies: 1. Palaeoenvironmental reconstructions generally characterise years as wetter or drier, while three categories were used herein: low, average and high. There will invariably be some overlap between the author's 'average' rainfall years, and those characterised as wetter or drier by others; 2. This study utilised modern rainfall data, as recorded by the Hyden meteorological station (no. 010568). As such, these data are only representative of periods of lower, average and higher rainfall that occur during the current climatic phase. Earlier periods may have had a wetter or 335 drier baseline, so palaeoenvironmental reconstructions consider the amount of effective rainfall compared to the modern climate. There is a difference of scale, then, between this research and palaeoenvironmental studies. The former will reflect smaller scale shifts in effective rainfall, while the latter are designed to identify larger scale changes; 3. Finally, gnammas are closed systems with no groundwater input. In contrast, most palaeoenvironmental proxies originate from waterbodies that interact with the regional groundwater table. As a result, each will respond to shifts in effective rainfall in different ways, and the results are not necessarily directly comparable. Keeping the above in mind, there are some notable similarities and differences between the rainfall record produced here and the palaeoenvironmental record from southwestern Australia. Interestingly, the record has little in common with those from inland southwestern Australia (Harrison 1993; Zheng et al. 2002) that show wetter conditions until 4000 BP (Figure 3.8); at that time Anderson Rocks record exhibited lower occupation intensity, indicative of drier conditions. In contrast, the same broad pattern of drier followed by wetter conditions is evident in three coastal records: those produced by Kendrick (1977), Semeniuk (1986) and Yassini and Kendrick (1988). The timing of their wet to dry transitions (4500–2800 BP) are not directly comparable, as none of their dates were calibrated, but they fall within a similar period to the 4571 cal. BP date identified here. Similarly, the drier period from 1056– 324 cal. BP fits fairly well with the record produced by Gouramanis et al. (2012), who identified drier conditions from 1400 cal. BP (Figure 3.8). However, these similarities should not be overstated. Considering the discordance of the southwestern palaeoenvironmental record, it would be difficult to produce a pattern that did not in some way resemble another. Indeed, if Pickett (1997) was correct, and the southwest has experienced no major Holocene climatic change, then the record produced for the study area may be capturing more minor fluctuations in effective rainfall. While the accuracy of the effective rainfall record is inconclusive, the idea itself shows some potential and warrants further consideration, as it represents a novel use of occupation intensity data. Such data are often interpreted through a palaeoenvironmental lens but, to the author's knowledge, have never been used as a proxy for palaeoenvironmental conditions themselves. 336 15.3 FUTURE RESEARCH DIRECTIONS This study has identified several areas that warrant future research. These fall into three main categories – the first two relate to the study area, and inland southwest as a whole, and involve increasing the body of archaeological data, and more accurately quantifying water availability. The third is more closely focussed on the assemblages collected from Anderson Rocks, Gibb Rock and Mulka's Cave, and outlines how certain methodological changes could maximise the quantity and reliability of available data. The primary aim for any future research should be increasing the quantity of archaeological data available for the study area, and the inland southwest in general. Rossi (2010, 2014a) had difficulty interpreting the Mulka’s Cave assemblage, primarily due to the lack of archaeological data from the surrounding area. The Gibb Rock assemblage typified the value of comparative data – this assemblage preserved a clear Short-term Expedient signature, and so provided quantifiable reduction intensity values that could be compared to those from other assemblages. This allowed the author to accurately distinguish between different provisioning systems made on lower quality raw material. If more data were available, the same method could be used to separate traces of Gearing Up from Individual Provisioning. At present, it is possible to theorise that reduction intensity should be higher under Individual Provisioning, but the smaller assemblages produced by both strategies means that these differences cannot be quantified. If high-quality materials were found at a site that had no reliable water, they would likely be part of an Individual Provisioning assemblage. While the small sample size may be problematic, it would provide some point of reference for reduction intensity values. As a result, excavation and surface collections should focus on sites without reliable water (probably around granite or salt lakes, due to the preservation bias associated with agricultural land), as well as those with larger rock structures, particularly in the central and northern study area. The latter represent the best opportunity to evaluate the antiquity of occupation in the study area, since the water sources imposed a strong tethering effect. Considering how the cleft rockhole shaped occupation of Mulka's Cave, Twine Reserve would be an ideal candidate for investigation, since the large gnammas represent a reliable water source. A more thorough survey of 337 Anderson Rocks and Gibb Rock would also be worthwhile, since access issues limited the portions of the site the author could visit. As demonstrated by this research, rainwater temporarily held in deeper rock structures was vital to year-round occupation of the study area. Therefore, it is pertinent to consider ways of more accurately quantifying these sources. Here, daily water balance was modelled using local rainfall data and an evaporation coefficient based on the inferred rate of water loss compared to Evaporation Pans and larger reservoirs. The coefficient is the main source of potential error and would be best refined by logging the ongoing water balance of a particular gnamma. Ideally, this gnamma would be deep, and receive no additional input from runoff. The daily/weekly water level could be measured and considered in conjunction with local rainfall data to determine the rate of water-loss through evaporation. This would also indicate how the evaporation rate varied with depth, since water held near the bottom of the gnamma would evaporate more slowly due to increased protection from the sun and wind. Ideally, different shaped structures could be studied to test the individual coefficients applied to those with sloping sides. Similarly, it would be beneficial to more accurately characterise site-specific water sources, with particular reference to the level of runoff received – this is dictated by the size of the area generating runoff, as well as the slope and texture of granite, and the amount of rainfall received (Fernie 1930); if these data were available, they could be factored into the model. Accurate data is vital for larger structures like the cleft rockhole at Mulka's Cave, whose water-holding properties strongly influenced interpretations. By more accurately characterising these water sources, particularly how they perform under various rainfall conditions, it may be possible to further develop the idea of using occupation intensity at particular sites as a proxy for effective rainfall. Finally, several methodological improvements could be applied to the cultural material analysed during this research. Ideally, more charcoal samples should be dated, allowing deposition and discard rates to be refined, and also to establish the age of certain deposits (e.g. pit AR1). Small flakes (< 10 mm long) dominated most assemblages, but limited data was collected from these artefacts. If the arbitrary boundary between small and large flakes was reduced to 5 mm, this would increase the sample size of larger flakes, for which more detailed attributes were recorded; as 338 a result of this change all flake scars on large flakes would have to be counted, excluding those produced by retouch or edge damage. The reduced flake length boundary would also more accurately separate flakes produced by retouch, which are less likely to measure > 5 mm. If small flakes were classified by raw material quality, and breakage rates recorded, the extent of retouch could be reliably determined based on waste products, which are more likely to enter the archaeological record than the retouched items themselves. The increased stone artefact data may allow more accurate identification of different provisioning systems, and aid in distinguishing Individual Provisioning from Gearing Up. Finally, the material excavated by Bowdler et al. (1989) warrants reanalysis, using the same methods as in this research, allowing for the changes suggested above. This would provide a more accurate indication of how the rockshelter at Mulka's Cave was used, and the increased artefact sample size may allow the question of ceremonial/ritual occupation to be addressed further. 15.4 CONCLUSION The main aim of this research was to determine when and how the study area was used by Aboriginal people. It was demonstrated that the study area was visited as early as 9650 cal. BP, and that regular occupation occurred from at least 7775 cal. BP and continued until after European arrival. Predictable resource distribution allowed year-round use of the study area and permitted a regular seasonal route – the intensity of occupation at particular sites depended on the water available there and elsewhere in the landscape. Summer and autumn occupation centred on larger gnammas in the northern and central study area, respectively. Residential mobility was limited, due to the restricted availability of water, so Place Provisioning dominated, and some Gearing Up occurred when stays were longer. People moved into the southern palaeovalley in winter when more abundant water allowed them to disperse more widely – this period of freedom was shorter under low rainfall conditions, and longer after strong winter rainfall. The increased residential mobility encouraged Individual Provisioning, supplemented by Short-term Expedient reduction of lower quality materials. Place Provisioning was restricted to Mulka's Cave, as the large ephemeral source of water held by the cleft rockhole permitted predictable spring occupation under all rainfall conditions. The source held sufficient 339 water for Mulka's Cave to be used as a refuge after very low winter rainfall, when water was scarce in the wider area. By late spring, groups moved northward to the central study area, when the few reliable water sources again limited residential mobility. Aggregation events, for ceremonial or social reasons, could have been held in the southern study area the summer after strong rainfall, when the required resources would have been available. However, this idea could not be addressed with the archaeological data. The quartz dominated assemblages indicate that, if these aggregation events did occur, they primarily involved Noongar groups. The paucity of high-quality exotics indicates the study area occupants had little access to the greenstone belt to the east, or the people who resided there; despite the proximity to desert groups, then, their influence may have been fairly limited. The occupation history summarised above is particularly significant since little archaeological or ethnohistorical data were available for the study area, or the inland southwest in general. The results demonstrate some similarities with parts of the southwest, where regular seasonal routes involved dispersal during winter, and parts of the Western Desert, where the lack of reliable water limited the frequency of aggregation events (Anderson 1984:34, 37; Bird 1985:176; Gould 1969b, 1984, 1991). After the author’s MA research (Rossi 2010), only six dates were available for inland southwestern Australia, all from Mulka's Cave (Williams and Smith 2013); no other sites yielded comparative artefact assemblages. At present, Anderson Rocks and Mulka's Cave yield a combined total of 22 radiocarbon dates, of which 17 were considered reliable. Artefact assemblages from Anderson Rocks and Gibb Rock both represent valuable comparative data that allowed Mulka's Cave to be placed in a regional context. Furthermore, Anderson Rocks had previously not met the criteria of a registered site under Section 5 of the Aboriginal Heritage Act (1972). As a result of this study, that ruling may need to be re-evaluated, and Gibb Rock may also be added to the register; the author will communicate these results to the appropriate authority. Overall, the results of this research represent a significant improvement in the number of known, dated archaeological sites. It has demonstrated that the study area was regularly occupied from the early to mid-Holocene transition and suggests that Pleistocene sites are unlikely to be found anywhere in the inland southwest, as 340 drier conditions would have limited the water held by ephemeral sources, which were vital to human occupation. While the value of archaeological and theoretical data for the inland southwest should not be underestimated, much of the significance of this research is in its approach, and apparent success thereof. This approach required compilation of a theoretical occupation model based on the distribution of critical resources, which was then merged with archaeological data. While opportunities to test the model were limited, its success is evident from the lack of conflict between model predictions and archaeological data. Particularly strong support was found in the southern study area, where most gnammas were associated with limited archaeological content – the model predicted that this part of the study area would be mostly used in winter and early spring, when water was more widely available. The model is a useful comparison to archaeological data, but is also an interpretive tool in its own right. It can be used to place specific sites in a wider context, based on the Landscape Division and part of the study area in which they occur. It also identifies the areas with archaeological potential for longer occupation sequences, due to the tethering effect of more reliable water sources in the central and northern study area. These inferences could be used to target particular locations for excavation, to evaluate/confirm the antiquity of occupation in the study area and provide valuable comparative stone artefact data. This research has highlighted the importance of considering spatio-temporal variation in the distribution of critical resources, and how this high-resolution data can permit detailed occupation models to be formulated. It has confirmed the value of such models, particularly in areas where comparative data are rare, and indicated how they can be integrated into an archaeological analysis, with neither archaeology nor theory taking precedence over the other. This study has devised novel ways to quantify the longevity of rainwater temporarily stored in soil profiles and rock structures – this approach could be applied to other areas where ephemeral water sources were integral to human survival. It has also demonstrated the value of floral and faunal surveys when compiling lists of plant and animal foods available prior to European arrival. This approach would be particularly useful in other agricultural areas, where extensive clearing prompts regular biodiversity audits of uncleared land. 341 Finally, this research has supported Rossi's (2014b) earlier assertions that, in some cases, occupation intensity can be uses as a proxy for effective rainfall; this identified periods of lower rainfall from 8200–4571 cal. BP, and throughout most of the last 1000 years. While the accuracy of the results is still in question, the method warrants further investigation, particularly due to the conflicting palaeoenvironmental records from southwestern Australia. 342 REFERENCES Abeysinghe, P.B. 2003 Silica resources of Western Australia. Western Australia Geological Survey, Mineral Resources Bulletin 21. Perth: Geological Survey of Western Australia. Abbott, I. 2002 Origin and spread of the cat, Felis catus, on mainland Australia, with a discussion of the magnitude of its early impact on native fauna. Wildlife Research 29:51–74. Abbott, I. 2003 Aboriginal fire regimes in south-west Western Australia: evidence from historical documents. In I. Abbott and N. Burrows (eds), Fire and ecosystems of south-west Western Australia: impacts and management, pp.119–146. Leiden: Backhuys Publishers. Abbott, I. 2008 Historical perspectives of the ecology of some conspicuous vertebrate species in south-west Western Australia. Conservation Science Western Australia 6(3):1–214. Allen, J. 1969 Archaeology and the history of Port Essington. Unpublished PhD thesis, School of Archaeology and Anthropology, Australian National University, Canberra. Allen, J. and R. Jones 1980 Oyster Cove: archaeological traces of the last Tasmanians and notes on the criteria for the authentication of flaked glass artefacts. Papers and Proceedings of the Royal Society of Tasmania 114:225–233. Allen, J. and J.F. O’Connell 2014 Both half right: updating the evidence for dating first human arrival in Sahul. Australian Archaeology 79:86–108. Allen, R.G., L.S. Pereira, D. Raes and M. Smith 1998 Crop evapotranspiration – guidelines for computing crop requirements. FAO Irrigation and Drainage paper 56. Rome: Food and Agriculture Organisation of the United Nations. Ambrose, S.H. and K.G. Lorenz 1990 Social and ecological models for the Middle Stone Age in Southern Africa. In P. Mellars (ed.), The emergence of modern humans: an archaeological perspective, pp.3–33. New York: Cornell University Press. Anand, R.R. and C.R.M Butt 2010 A guide for mineral exploration through the regolith in the Yilgarn Craton, Western Australia. Australian Journal of Earth Sciences 57(8):1015–1114. https://dx.doi.org/10.1080/08120099.2010.522823 Anand, R.R. and M. Paine 2002 Regolith geology of the Yilgarn Craton, Western Australia: implications for exploration. Australian Journal of Earth Sciences 49(1):3– 162. Anderson, J. 1984 Between plateau and plain: flexible responses to varied environments in south-western Australia. Occasional Papers in Prehistory 4. Canberra: Research School of Pacific Studies, Australian National University. 343 Andrefsky, W. 1994 Raw-material availability and the organization of technology. American Antiquity 59:21–34. Andrich, M.A. and J. Imberger 2013. The effect of land clearing on rainfall and fresh water sources in Western Australia: a multi-functional sustainability analysis. International Journal of Sustainable Development and World Ecology 2013. https://dx.doi.org/10.1080/13504509.2013.850752 Atkins, C. and Y-L. Ruan 2010–2016 Phloem transport. In R. Munns, S. Schmidt and C. Beveridge (eds), Plants in Action (2nd ed.). Retrieved from < http://plantsinaction.science.uq.edu.au/content/chapter-5-phloem-transport > Atlas of Living Australia (n.d.) Atlas of Living Australia [Computer Software]. Retrieved from < https://www.ala.org.au/ > Attenbrow, V. 2004 What’s Changing: Population Size or Land-use Patterns? The Archaeology of Upper Mangrove Creek, Sydney Basin. Terra Australis 21. Canberra: Australian National University. Australian Biological Resources Study 2009 Australian Faunal Directory [Computer Software]. Retrieved from < https://biodiversity.org.au/afd/home > Australian Government Department of Environment and Energy 2012 Interim Biogeographic Regionalisation for Australia (IBRA), Version 7. Retrieved 7 September 2018 from < http://www.environment.gov.au/land/nrs/science/ibra/australias-bioregions-maps > Balme, J. 2014 Devils Lair: occupation intensity and land-use. Australian Archaeology 79:179–186. Bamforth, D.B. 1986 Technological efficiency and tool curation. American Antiquity 51(1):38–50. Barham, A. and G. Huckleberry 2014 Sediments. In J. Balme and A. Paterson (eds), Archaeology in practice: a student guide to archaeological analyses (2nd ed.), pp.47– 84. West Sussex: Wiley-Blackwell. Barrows, T.T., S. Juggins, P. De Deckker, E. Calvo and C. Pelejero 2007 Long-term sea surface temperature and climate change in the Australian–New Zealand region. Paleoceanography 22:PA2215. https://dx.doi.org/10.1029/2006PA001328 Bates, D. 1938 The passing of the Aborigines. London: John Murray Ltd. Bayly, I.A.E. 1999 Review of how indigenous people managed for water in desert regions of Australia. Journal of the Royal Society of Western Australia 82(1):17–25. Bayly, I. 2015 The importance of gnammas for Indigenous Australians. Wildlife Australia 52(2):23–25. 344 Beard, J.S. 1972 Vegetation survey of Western Australia. The vegetation of the Southern Cross area, Western Australia. Map and Explanatory memoir, 1:250 000 series. Sydney: Vegmap Publications. Beard, J.S. 1975 The vegetation survey of Western Australia. Vegetatio 30(3):179– 187. Beard, J.S. 1976 Vegetation survey of Western Australia. The vegetation of the Boorabbin and Lake Johnston areas, Western Australia. Map and explanatory memoir, 1:250 000 series. Sydney: Vegmap Publications. Beard, J.S. 1979 Vegetation survey of Western Australia. The vegetation of the Hyden area, Western Australia. Map and explanatory memoir, 1:250 000 series. Sydney: Vegmap Publications. Beard, J.S. 1980 Vegetation survey of Western Australia. The vegetation of the Corrigin area, Western Australia. Map and explanatory memoir, 1:250 000 series. Sydney: Vegmap Publications. Beard, J.S. 1981 Vegetation survey of Western Australia. Swan. Series 1:1,000,000. Crawley: University of Western Australia Press. Beard, J.S. 1990 Plant life of Western Australia. Kenthurst: Kangaroo Press. Beard, J.S. 1999 Evolution of the river systems of the south-west drainage division, Western Australia. Journal of the Royal Society of Western Australia 82:147–164. Beard, J.S., A.R. Chapman and P. Gioia 2000 Species richness and endemism in the Western Australian flora. Journal of Biogeography 27(6):1257–1268. Beaton, J.M. 1982 Fire and water: aspects of Australian Aboriginal management of cycads. Archaeology in Oceania 17(1):51–58. Beaumont, P. 1961 The problem of glass implements. Bulletin of the South African Museums Association 7(8):161–163. Beck, W. 1992 Aboriginal preparation of Cycas seeds in Australia. Economic Botany 46(2):133–147. Beisaw, A. 1998 Differential preservation and recovery: taphonomy of bone preservation at the Thomas/Luckey site, Ashland, NY. Unpublished MA thesis, State University of New York, Binghamton. Bell, D.T. and J.E. Williams 1997 Eucalypt Physiology. In J.E. Williams and J.C.Z. Woinarski (eds), Eucalypt ecology: Individuals to ecosystems, pp.168–196. Cambridge: Cambridge University Press. 345 Benison, K.C., B.B. Bowen, F.E. Oboh-Ikeunobe, E.A. Jagniecki, D.A. LaClair, S.L. Story, M.R. Mormile and B-Y. Hong 2007 Sedimentology of acid saline lakes in southern Western Australia: newly described processes and products of an extreme environment. Journal of Sedimentary Research 77:366–388. Benson, J. 1991 The effect of 200 years of European settlement on the vegetation and flora of New South Wales. Cunninghamia 2(3):343–370. Bettenay, E. 1962 The salt lake systems and their associated aeolian features in the semi-arid regions of Western Australia. Journal of Soil Science 13(1):10–17. Bettenay, E., A.V. Blackmore and F.J. Hingston 1964 Aspects of the hydrologic cycle and related salinity in the Belka Valley, Western Australia. Australian Journal of Soil Research 1964(2):187–210. Bettenay, E. and F.J. Hingston 1963 The quality of groundwaters in the central wheatbelt of W.A. Journal of the Department of Agriculture 4(4):216–219. Bettinger, R.L., R. Garvey and S. Tushingham 2015 Hunter-gatherers: archaeological and evolutionary theory (2nd ed.). New York: Springer. Bickford, S. and Gell, P. 2005 Holocene vegetation change, Aboriginal wetland use and the impact of European settlement on the Fleurieu Peninsula, South Australia. The Holocene 15(2):200–215. Bindon, P.R. 1996. Useful bush plants. Perth: Western Australian Museum. Bindon, P.R. 1997 Aboriginal people and granite domes. Journal of the Royal Society of Western Australia 80:173–179. Binford, L.R.1973 Interassemblage variability – the Mousterian and the functional argument. In C. Renfrew (ed.), The explanations of culture change – models in prehistory, pp.227–254. London: Duckworth. Binford, L.R. 1977 Forty seven trips. A case study in the character of archaeological formation processes. In R.V.S. Wright (ed.), Stone tools as cultural markers, pp.24– 36. Canberra: Australian Institute of Aboriginal Studies. Binford, L.R. 1979 Organization and formation processes: looking at curated technologies. Journal of Anthropological Research 35(3):255–273. Binford, L.R. 1980 Willow smoke and dogs' tails: Hunter-gatherer settlement systems and archaeological site formation. American Antiquity 45(1):4–20. Biological Surveys Committee 1984 The biological survey of the Eastern Goldfields of Western Australia, part 1: introduction and methods. Records of the Western Australian Museum Supplement 18:1–19. 346 Bird, C.F.M. 1985. Prehistoric lithic resource utilisation: a case study from the southwest of Western Australia. Unpublished PhD thesis, Centre for Prehistory, University of Western Australia, Crawley. Bird, C.F.M. and C. Beeck 1988 Traditional plant foods in the southwest of Western Australia: the evidence from salvage ethnography. In B. Meehan and R. Jones (eds), Archaeology with ethnography: an Australian perspective, pp.113–22. Canberra: Research School of Pacific Studies, Australian National University. Bird, D.W., R.B. Bird and B.F. Codding 2009 In pursuit of mobile prey: Martu hunting strategies and archaeofaunal interpretation. American Antiquity 74(1):3–29. Bird, R.B., D.W. Bird, B.F. Codding, C.H. Parker and J.H Jones 2008 The “fire stick farming” hypothesis: Australian Aboriginal foraging strategies, biodiversity, and anthropogenic fire mosaics. Proceedings of the National Academy of Sciences of the United States of America 105(39):14,796–14,801. Bird, D.W., R.B. Bird and C.H. Parker 2005 Aboriginal burning regimes and hunting strategies in Australia’s Western Desert. Human Ecology 33(4):443–464. Bird, D.W., B.F. Codding, R.B. Bird and D.W. Zeanah 2012 Risky pursuits: Martu hunting and the effects of prey mobility: reply to Ugan and Simms. American Antiquity 77(1):186–194. Bird, M.I., D. O’Grady and S. Ulm 2016 Humans, water and the colonization of Australia. Proceedings of the National Academy of Sciences 113(41):11,477–11,482. https://dx.doi.org/10.1073/pnas.1608470113 Bird, R.B., N. Tayor, B.F. Codding and D.W. Bird 2013 Niche construction and Dreaming logic: aboriginal patch mosaic burning and varanid lizards (Varanus gouldii) in Australia. Proceedings of the Royal Society B 280:20132297. https://dx.doi.org/10.1098/rspb.2013.2297 Blatt, H. and R.J. Tracy 1996 Petrology: igneous, sedimentary, and metamorphic (2nd ed.). New York: W.H. Freeman and Company. Bleed, P. 1986 The optimal design of hunting weapons: maintainability or reliability. American Antiquity 51(4):737–747. Blewett, R.S., B.L.N. Kennett and D.L. Huston 2012 Australia in time and space. In R.S. Blewett (ed.), Shaping a nation: a geology of Australia, pp.47–120. Canberra: Geoscience Australian and ANU E Press. Bowdler, S., J. Harris, A. Murphy, G. Nayton and C. Pocock 1989. A test excavation at Mulka’s Cave (Bates Cave) near Hyden, Western Australia. Unpublished report prepared for the Department of Aboriginal Sites, Western Australian Museum, Perth. Bowen, B.B. and K.C. Benison 2009 Geochemical characteristics of naturally acid and alkaline saline lakes in southern Western Australia. Applied Geochemistry 24:268–284. 347 Bowler, J.M. 1986 Spatial variability and hydrologic evolution of Australian lake basins: analogue for Pleistocene hydrologic change and evaporite formation. Palaeogeography, Palaeoclimatology, Palaeoecology 54:21–41. Bowler, J.M., G.S. Hope, J.N. Jennings, G. Singh and D. Walker 1976 Late quaternary climates of Australia and New Guinea. Quaternary Research 6(3):359– 394. Braadbaart, F., I. Poole and A.A. van Brussel 2009 Preservation potential of charcoal in alkaline environments: an experimental approach and implications for the archaeological record. Journal of Archaeological Science 36:1672–1679. https://dx.doi.org/10.1016/j.jas.2009.03.006 Brand, J. 2005 WA sandalwood (Santalum spicatum) establishment guide for farmland in the Wheatbelt, July 2005. Retrieved 27 March 2017 from < http://avongro.com.au/Webpages/documents/sandal2005Jul.pdf > Brand-Miller, J.C. and S.H.A. Holt 1998 Australian Aboriginal plant foods: a consideration of their nutritional composition and health implications. Nutrition Research Reviews 11:5–23. Brand-Miller, J., K.W. James and P.M.A. Maggiore 2014 Tables of composition of Australian Aboriginal foods. Canberra: Aboriginal Studies Press. Briscoe, G. 2003 Counting, health and identity: a history of Aboriginal health and demography in Western Australia and Queensland, 1900–1940. Canberra: Aboriginal Studies Press. Brodie, R.S., K.C. Lawrie and D.P. Commander 2012 Groundwater – lifeblood of the continent. In R.S. Blewett (ed.), Shaping a nation: a geology of Australia, pp.333– 379. Canberra: Geoscience Australia and ANU E-press. Bronk Ramsey, C. 2009 Bayesian analysis of radiocarbon dates. Radiocarbon 51(1):337–360. Brooke, B.P., S.L. Nichol, Z. Huang and R.J. Beaman 2017 Palaeoshorelines on the Australian continental shelf: Morphology, sea-level relationship and applications to environmental management and archaeology. Continental Shelf Research 134:26– 38. Brown, A., P. Dundas, K. Dixon and S. Hopper 2008 Orchids of Western Australia. Crawley: UWA Press. Brown, A.R. 1918 Notes on the social organization of Australian tribes. The Journal of the Royal Anthropological Institute of Great Britain and Ireland 48:222–253. Cale, D.J., S.A. Halse and C.D. Walker 2004 Wetland monitoring in the Wheatbelt of south-west Western Australia: site descriptions, waterbird, aquatic invertebrate and groundwater data. Conservation Science Western Australia 5(1):20–135. 348 Campbell, E.M. 1997 Granite landforms. Journal of the Royal Society of Western Australia 80(3):101–112. Campbell, J. 2002 Invisible invaders: smallpox and other diseases in Aboriginal Australia 1780–1880. Melbourne: Melbourne University Press. Cane, S. 1987 Australian Aboriginal subsistence in the Western Desert. Human Ecology 15(4):391–434. Chapman, A. and J. Dell 1977 Reptiles and frogs of Bendering and West Bendering nature reserves. Records of the Western Australian Museum Supplement 5:47–58. Chappell, J. 1991 Late Quaternary environmental changes in eastern and central Australia, and their climatic interpretation. Quaternary Science Reviews 10(5):377– 390. Charnov, E. 1976 Optimal foraging: the marginal value theorem. Theoretical Population Biology 9(2):129–136. Chin, R.J., A.H. Hickman and R. Thom 1984 Hyden SI 50-4 Australia 1:250 000 geological series – Explanatory notes. Perth: Geological Survey of Western Australia. Churchill, D.M. 1968 The distribution and prehistory of Eucalyptus diversicolor F. Muell., E. marginata Donn ex Sm., and E. calophylla R. Br. in relation to rainfall. Australian Journal of Botany 16(1):125–151. Clarke, C.J., R.J. George, R.W. Bell and T.J. Hatton 2002 Dryland salinity in southwestern Australia: its origins, remedies and future research directions. Australian Journal of Soil Research 40(1):93–113. Clarke, C.J., R.J. George, R.W. Bell and R.J. Hobbs 1998 Major faults and the development of dryland salinity in the western wheatbelt of Western Australia. Hydrology and Earth System Sciences 2(1):77–91. Clarke, P.A. 2003 Australian ethnobotany: an overview. Australian Aboriginal Studies 2003/2:21–38. Clarke, P.A. 2009 Australian Aboriginal ethnometeorology and seasonal calendars. History and Anthropology 20(2):79–106. https://dx.doi.org/10.1080/02757200902867677 Clarke, P.A. 2011 Aboriginal people and their plants. Dural: Rosenberg Publishing. Clarkson, C. 2002 An index of invasiveness for the measurement of unifacial and bifacial retouch: a theoretical, experimental and archaeological verification. Journal of Archaeological Science 29:65–75. Clarkson, C. 2004 Technological provisioning and assemblage variation in the Eastern Victoria River region, Northern Australia: A Darwinian approach. Unpublished PhD thesis, The Australian National University, Canberra. 349 Clarkson, C. 2006 Interpreting surface assemblage variation in Wardaman country, Northern Territory: an ecological approach. In S. Ulm and I. Lilley (eds), An archaeological life: papers in honour of Jay Hall, pp.177–190. Aboriginal and Torres Strait Islander Studies Unit Research Report Series, 7. Brisbane: University of Queensland. Clarkson, C. 2007 Lithics on the land of the lightning brothers: the archaeology of Wardaman country, Northern Territory. Terra Australis 25. Canberra: ANU e-press. Clarkson, C., M. Haslam and C. Harris 2015 When to retouch, haft or discard? Modelling optimal use/maintenance schedules in lithic tool use. In N. Goodale and W. Andrefsky (eds), Lithic technological systems and evolutionary theory, pp.117–138. New York: Cambridge University Press. Clarkson, C., Z. Jacobs, B. Marwick, R. Fullagar, L. Wallis, M. Smith, R.G. Roberts, E. Hayes, K. Lowe, X. Carah, S.A. Florin, J. McNeil, D. Cox, L.J. Arnold, Q. Hua, J. Huntley, H.E.A. Brand, T. Manne, A. Fairbairn, J. Shulmeister, l. Lyle, M Salinas, M. Page, K. Connell, G. Park, K. Norman, T. Murphy and C. Pardoe 2017 Human occupation of northern Australia by 65,000 years ago. Nature 547:306–310. https://dx.doi.org/10.1038/nature22968 Clarkson, C. and S. O’Connor 2014 An introduction to stone artefact analysis. In J. Balme and A. Paterson (eds), Archaeology in practice: a student guide to archaeological analyses (2nd ed.), pp.151–206. West Sussex: Wiley-Blackwell. Cleland, J.B. 1928 Diseases amongst the Australian Aborigines. London: Bale, Sons and Danielsson Ltd. Coates, A. 1992 Flora and vegetation survey of Dragon Rocks nature reserve (no. A 36128). Unpublished report prepared for the Department of Conservation and Land Management, Western Australia. Cogger, H. 2014 Reptiles and amphibians of Australia. Collingwood: CSIRO Publishing. Collard, M., B. Buchanan, M.J. O’Brien and J. Scholnick 2013 Risk, mobility or population size? Drivers of technological richness among contact-period western North American hunter-gatherers. Philosophical Transactions of the Royal Society B 368:20120412. https://dx.doi.org/10.1098/rstb.2012.0412 Collard, L. and S. Harben 2010 Nartj Katitj Bidi Ngulluckiny Koorl? (Which knowledge path will we travel?). Studies in Western Australian History 26:75–95. Collard, M., M. Kemery and S. Banks 2005 Causes of toolkit variation among huntergatherers: a test of four competing hypotheses. Canadian Journal of Archaeology 29:1–19. 350 Commander, D.P., N. Schoknecht, W. Verboom and P. Caccetta 2001 The geology, physiography and soils of Wheatbelt valleys. In Dealing with salinity in Wheatbelt valleys: processes, prospects and practical options: proceedings of a conference held at Merredin, Western Australia – 30th July – 1st August 2001. Perth: Viv Read and Associates. Conkey, M.W. 1980 The identification of prehistoric hunter-gatherer aggregation sites: the case of Altamira. Current Anthropology 21(5):609–630. Conte, I.C. and F.G. Romero 2008 Microwear analysis of retouched glass fragments from Fortlet Miñana, Azul, Argentina, 1860–1863.International Journal of Historical Archaeology 12(3):2480262. Cooper, Z. and S. Bowdler 1988 Flaked glass tools from the Andaman Islands and Australia. Asian Perspectives 37(1):74–83. Corsini, S. 1993 Investigation of Aboriginal sites in the vicinity of Wave Rock, Hyden. Unpublished report prepared for the Mallee Aboriginal Corporation and the Western Australian Museum. Cotterell, B., and J. Kamminga 1979 The mechanics of flaking. In B. Hayden (ed.), Lithic usewear analysis, pp.97–112. New York: Academic Press. Cotterell, B. and J. Kamminga 1987 The formation of flakes. American Antiquity 52(4):675–708. Cowgill, G.L. 2015 Some things I hope you will find useful even if statistics isn’t your thing. Annual Review of Anthropology 44:1–14. Cribb, A.B. and J.W. Cribb 1975 Wild food in Australia. Sydney: William Collins Publisher. Cronin, L. 2008 Cronin’s key guide: Australian mammals. Crows Nest: Jacana Books. Cronin, L. 2014 Cronin’s key guide: Australian reptiles and frogs. Crows Nest: Jacana Books. Cullen, L.E. and P.F. Grierson 2009 Multi-decadal scale variability in autumn-winter rainfall in south-western Australia since 1655 AD as reconstructed from tree rings of Callitris columellaris. Climate Dynamics 33(2–3):433–444. Curr, E.M. 1886 The Australian race: its origin, languages, customs, place of landing in Australia and the routes by which it spread itself over that continent, Volume 1. London: Trübner and Co. Curr, E.M. 1887 Continental Australia showing the routes by which the Aboriginal race spread itself throughout the continent [map]. Victoria: Lands and Surveys Department. 351 David, B., B. Barker, F. Petchey, J-J. Delannoy, J-M. Geneste, C. Rowe, M. Eccleston, L. Lamb and R. Whear 2013 A 28,000-year-old excavated painted rock from Nawarla Gabarnmang, northern Australia. Journal of Archaeological Science 40:2493–2501. Davidson, D.S. 1938 An ethnic map of Australia. Proceedings of the American Philosophical Society 79:649–680. Davidson, D.S. 1952. Notes on the pictographs and petroglyphs of Western Australia and a discussion of their affinities with appearances elsewhere on the continent. Proceedings of the American Philosophical Society 96:76–117. Davidson, W.A. 1977 Hydrogeology and drilling results of the 1969–1970 drought relief program. Geological Survey of Western Australia Record no. 1977/3. Perth: Geological Survey of Western Australia. Daw, B., T. Walley and G. Keighery 1997 Bush tucker plants of the south-west. Como: Department of Conservation and Land Management. Day, R.B. 1951 The Humps beyond Hyden. The West Australian Saturday 27 January 1951 p. 18. Deardorff, J.W. 1977 A parameterization of ground-surface moisture content for use in atmospheric prediction models. Journal of Applied Meteorology 16:1182–1185. Deardorff, J.W. 1978 Efficient prediction of ground surface temperature and moisture, with inclusion of a layer of vegetation. Journal of Geophysical Research 83:1889–1903. De Dekker, P. 1983 Australian salt lakes: their history, chemistry, and biota – a review. Hydrobiologia 105(1):231–244. De La Pena, P. 2015 A qualitative guide to recognize bipolar knapping for flint and quartz. Lithic Technology 40(4):1–16. Dell, J. 1977 Birds of Bendering and West Bendering nature reserves. Records of the Western Australian Museum Supplement 5:31–46. de Lombera Hermida, A. 2008 Quartz morphostructural groups and their mechanical implications. Annali dell’Università degli Studi di Ferrara 2008:101–104. de Lombera Hermida, A. 2009 The scar identification of lithic quartz industries. In F. Sternke, L. Eigeland and L.J. Costa (eds), Non-flint raw material use in prehistory, pp.5–11. Oxford: Archaeopress. Department of Environment 2005a The ecology of wheatbelt lakes. Water notes 33. Perth: Department of Environment Department of Environment 2005b The Wheatbelt's ancient rivers. Water notes 34. Perth: Department of Environment. 352 Department of Primary Industries and Regional Development (DPIRD) 2016a Soil Water Tool [Computer Software]. Retrieved from < https://www.agric.wa.gov.au/climate-weather/soil-water-tool > Department of Primary Industries and Regional Development (DPIRD) 2016b NRInfo [Computer Software]. Retrieved from < https://maps.agric.wa.gov.au/nrinfo/ > Department of Water 2009 Waterway assessment of the Camm River: Lockhart River confluence to Hyden. Water resource management series report number WRM 57. Perth: Department of Water. Dickman, C.R. 1996 Impact of exotic generalist predators on the native fauna of Australia. Wildlife Biology 2(3):185–195. Dickson, F.P. 1977 Quartz flaking. In R.V.S. Wright (ed.) Stone tools as cultural markers: change, evolution and complexity, pp.97–103. Canberra: Australian Institute of Aboriginal Studies. Diez-Martin, F., P.S. Yustos, M. Dominguez-Rodrigo and M.E. Prendergast 2011 An experimental study of bipolar and freehand knapping of Naibor Soit quartz from Olduvai Gorge (Tanzania). American Antiquity 76(4):690–708. Dingman, S.L. 1994 Physical hydrology. New York: Macmillan. Dirnböck, T., R.J. Hobbs, R.J. Lambeck and P.A. Caccetta 2002 Vegetation distribution in relation to topographically driven processes in southwestern Australia. Applied Vegetation Science 5(1):147–158. Dodson, J.R. 2001 Holocene vegetation change in the Mediterranean-type regions of Australia. The Holocene 11(6):673–680. Dodson, J.R. and J.J. Lu 2000 A Late Holocene vegetation and environment record from Byenup Lagoon, south-western Australia. Australian Geographer 31(1):41–54. Dominguez-Villar, D., C. Arteaga, R. Garcia-Gimenez, E.A. Smith and J. Pedraza 2008 Diurnal and seasonal water variations of temperature, pH, redox potential and conductivity in gnammas (weathering pits): implications for chemical weathering. Catena 72:37–48. Dortch, C.E. 2000 Past Aboriginal hunter-gatherer economy and territorial organisation in coastal districts of Western Australia's lower Southwest. Unpublished PhD thesis, Centre for Archaeology, Department of Anthropology, University of Western Australia, Crawley. Dortch, J. 1996 Late Pleistocene and recent Aboriginal occupation of Tunnel Cave and Witchcliffe rock shelter, Southwestern Australia. Australian Aboriginal Studies 2:51–60. 353 Dortch, J. 2000 Palaeo-environmental change and the persistence of human occupation in south-western Australian forests. Unpublished PhD thesis, Centre for Archaeology, Department of Anthropology, University of Western Australia, Crawley. Dortch, J. 2004 Palaeoenvironmental change and the persistence of human occupation in south-western Australian forests. BAR International Series 1288. Oxford: Archaeopress. Dowling, P. 1997 “A great deal of sickness”: introduced diseases among the Aboriginal people of colonial southeast Australia, 1788–1900. Unpublished PhD thesis, Australian National University, Canberra. Driscoll, K. 2010 Understanding quartz technology in early prehistoric Ireland. Unpublished PhD thesis, College of Art and Celtic Studies, School of Archaeology, University College Dublin, Belfield. Driscoll, K. 2011a Vein quartz in lithic traditions: an analysis based on experimental archaeology. Journal of Archaeological Science 38:734–745. Driscoll, K. 2011b Identifying and classifying vein quartz artefacts: an experiment conducted at the World Archaeological Congress, 2008. Archaeometry 53(6):1280– 1296. https://dx.doi.org/10.1111/j.1475-4754.2011.00600.x Dyson-Hudson, R. and E.A. Smith 1978 Human territoriality: an ecological reassessment. American Anthropologist 80:21–41. Emlen, J.M. 1966 The role of time and energy in food preference. The American Naturalist 100(916):611–617. Erickson, R., A.S. George, N.G. Marchant and M.K. Morcombe 1979 Flowers and plants of Western Australia (revised ed.). Sydney: Reed. Farrington, P. and R.B. Salama 1996 Controlling dryland salinity by planting trees in the best hydrogeological setting. Land Degradation and Development 7:183–204. Ferguson, W.C. 1985 A mid-Holocene depopulation of the Australian Southwest. Unpublished PhD thesis, Australian National University, Canberra. Fernie, N. 1930 Water supplies from rock catchments in the Western Australian wheat belt. The Journal of the Institution of Engineers Australia 2(6):198–208. Ferris, J.M. 2015 Procurement costs and tool performance requirements: determining constraints on lithic toolstone selection in Baja Calafornia Sur. In N. Goodale and W. Andrefsky (eds), Lithic technological systems and evolutionary theory, pp.139–155. New York: Cambridge University Press. Fetherston, J.M., S.M. Stocklmayer and V.C. Stocklmayer 2017 Gemstones of Western Australia second edition. Geological Survey of Western Australia, Mineral Resources Bulletin 25. Perth: Geological Survey of Western Australia. 354 Florabank (n.d.) Fact Sheets. Retrieved 27 March 2017 from < http://www.florabank.org.au/lucid/key/species%20navigator/media/html/index.htm > Fullagar, R., J. McDonald, J. Field and D. Donlon 2009 Deadly weapons: backed microliths from Narrabeen, New South Wales. In M. Haslam, G. Robertson, A. Crowther, S. Nugent and L. Kirkwood (eds), Archaeological science under a microscope: Studies in residue and ancient DNA analysis in honour of Thomas H. Loy, pp.258–270. Terra Australis 30. Canberra: ANU E Press. Galloway, P.D. 2004 Agricultural sub-regions of the Avon River basin. Resource Management Technical Report 284. Western Australia: Department of Agriculture. Galloway, R.W. 1965 Late Quaternary climates in Australia. The Journal of Geology 73(4):603–618. Gammage, B. 2011 The biggest estate on earth: how Aborigines made Australia. Crow's Nest: Allen and Unwin. Garvey, R. 2015 A model of lithic raw material procurement. In N. Goodale and W. Andrefsky (eds), Lithic technological systems and evolutionary theory, pp.156–171. New York: Cambridge University Press. Geological Survey of Western Australia 1984 Hyden SI 50-4 [map] Australia 1:250 000 geological series. Perth: Geological Survey of Western Australia. George, R.J. 1992 Hydraulic properties of groundwater systems in the saprolite and sediments of the wheatbelt, Western Australia. Journal of Hydrology 130:251–278. George, R., J. Clarke and P. English 2008 Modern and palaeogeographic trends in the salinisation of the Western Australian wheatbelt: a review. Australian Journal of Soil Research 46:751–767. George, R.J. and M. Coleman 2001 Hidden menace or opportunity – groundwater hydrology, playas and commercial options for salinity in wheatbelt valleys. In Dealing with salinity in Wheatbelt valleys: processes, prospects and practical options: proceedings of a conference held at Merredin, Western Australia – 30th July – 1st August 2001. Perth: Viv Read and Associates. George, R., D. McFarlane and B. Nulsen 1997 Salinity threatens the viability of agriculture and ecosystems in Western Australia. Hydrogeology Journal 5(1):6–21. Ghauri, S. 2004 Groundwater trends in the central agricultural region. Resource Management Technical Report 269. Perth: Department of Agriculture and Food. Gibbs, M. and P. Veth 2002 Ritual Engines and the Archaeology of Territorial Ascendancy. In S. Ulm, C. Westcott, J. Reid, A. Ross, I. Lilley, J. Pragnell and L. Kirkwood (eds), Barriers, Borders, Boundaries: Proceedings of the 2001 Australian Archaeological Association Annual Conference Tempus 7, pp.11–19. Brisbane: University of Queensland. 355 Gibson, N., G.J. Keighery, N.M. Lyons and A. Webb 2004 Terrestrial flora and vegetation of the Western Australian wheatbelt. Records of the Western Australian Museum Supplement 67:139–189. Glendenning, W. 2004 A report of an Aboriginal heritage survey of the area of proposed water treatment facilities and associated pipework upgrades at Karlgarin and Wave Rock, near Hyden. Unpublished report prepared for the Water Corporation. Glover, J.E. 1984 The geological sources of stone for artefacts in the Perth Basin and nearby areas. Australian Aboriginal Studies 1984/1:17–25. Goode, B. 2011 An ethnographic Aboriginal heritage survey of the proposed development of Reserve 23128 in Hyden, Western Australia. Unpublished report prepared for South West Aboriginal Land and Sea Council (SWALSC). Goodyear, A.C. 1989 A hypothesis for the use of crypto-crystalline raw materials among Paleoindian groups of North America. In C.G. Ellis and J.C. Lothrop (eds), Eastern Paleoindian lithic resource use, pp.1–9. Boulder: Westview Press. Gordon, C.C. and J.E. Buikstra 1981 Soil pH, bone preservation, and sampling bias at mortuary sites. American Antiquity 46(3):566–571. Gould, R.A. 1968 Living archaeology: the Ngatatjara of Western Australia. Southwestern Journal of Anthropology 24(2):101–122. Gould, R.A. 1969a Yiwara: Foragers of the Australian Desert. London: Collins. Gould, R.A. 1969b Subsistence behaviour among the Western Desert Aborigines of Australia. Oceania 39(4):253–274. Gould, R.A. 1971a The archaeologist as ethnographer: A case study from the Western Desert of Australia. World Archaeology 3(2):143–177. Gould, R.A. 1971b Uses and effects of fire among the Western Desert Aborigines of Australia. Mankind 8:14–24. Gould, R.A. 1984 Rock pools and desert dances. Natural History 93(3):61–71. Gould, R.A. 1991 Arid-land foraging as seen from Australia: Adaptive models and behavioral realities. Oceania 62(1):12–33. Gouramanis, C., J. Dodson, D. Wilkins, P. De Deckker and B.M. Chase 2012 Holocene palaeoclimate and sea-level fluctuation recorded from the coastal Barker Swamp, Rottnest Island, south-western Western Australia. Quaternary Science Reviews 54:40–57. Goward, T. 2011 Aboriginal glass artefacts of the Sydney region. Unpublished Hons thesis, Department of Archaeology, University of Sydney, Camberdown/Darlington. 356 Graf, K.E. 2010 Hunter-gatherer dispersals in the mammoth-steppe: technological provisioning and land-use in the Enisei River valley, south-central Siberia. Journal of Archaeological Science 37:210–223. Green, N. 2011 Explorers and the Aborigines. In M. Herlock and S. Milentis (eds), Western Australian exploration 1836–1845, pp.419–437. Perth: Hesperian Press. Green, N.J. 1981 Aborigines and white settlers in the nineteenth century. In C.T. Stannage, C.T. (ed.), A new history of Western Australia, pp.72–123. Crawley: University of Western Australia Press. Green, R.L. and R.J. Watling 2007 Trace element fingerprinting of Australian ocher using Laser Ablation Inductively Coupled Plasma-Mass Spectrometry (LA-ICP-MS) for the provenance establishment and authentication of Indigenous art. Journal of Forensic Sciences 52(4):851–859. https://doi.org/10.1111/j.1556-4029.2007.00488.x Grey, G. 1841 Journals of two expeditions of discovery in north-west and western Australia, during the years 1837, 38 and 39. London: Boone. Grøn, O. 2012 Our grandfather sent the elk – some problems for hunter-gatherer predictive modelling. Quartär 59:175–188. Gunn, R.G. 2003 Mulka’s Cave Aboriginal Rock Art Site: a discussion of management options. Unpublished report prepared for the Department of Indigenous Affairs. Gunn, R.G. 2004 Mulka’s Cave Aboriginal rock art site: an archaeological recording. Unpublished report prepared for the Department of Indigenous Affairs and HydenKarlgarin Landcare Group. Gunn, R.G. 2006 Mulka’s Cave Aboriginal rock art site: its context and content. Records of the Western Australian Museum 23:19–41. Gunness, A. 2002 A survey of vegetation and flora: Wogarl Reserve, in the Graball Catchment, Shire of Narembeen. Nedlands: Wildflower Society of Western Australia. Haebich, A. and L. Tilbrook 1981 South west Aboriginal studies bibliography: with annotations and appendices. Mount Lawley: Mount Lawley College. Hall, J. and W.R.F. Love 1985 Prickly Bush, a site with backed blades on the Brisbane River: a pilot study towards the measurement of site “disturbance”. Queensland Archaeological Research 2:71–81. Hallam, S.J. 1972 An archaeological survey of the Perth area, Western Australia: a progress report on art and artefacts, dates and demography. Australian Institute of Aboriginal Studies Newsletter 3(5):11–19. Hallam, S.J. 1975. Fire and hearth: a study of European usage and European usurpation in southwestern Australia. Canberrra: Australian Institute of Aboriginal Studies. 357 Hallam, S.J. 1977 Topographic archaeology and artifactual evidence. In R.V.S. Wright (ed.), Stone tools as cultural markers: change, evolution and complexity, pp.169–177. Canberra: Australian Institute of Aboriginal Studies. Hallam, S.J. 1989 Plant usage and management in southwest Australian Aboriginal societies. In D.R. Harris and G.C. Hillman (eds), Foraging and farming: the evolution of plant exploitation, pp. 136–147. London: Unwin Hyman Ltd. Halse, S.A., M.N. Lyons, A.M. Pinder and R.J. Shiel 2004 Biodiversity patterns and their conservation in wetlands of the Western Australian wheatbelt. Records of the Western Australian Museum Supplement 67:337–364. Halse, S.A., G.B. Pearson and S. Patrick 1993 Vegetation of depth-gauged wetlands in nature reserves of south-west Western Australia. Department of Conservation and Land Management Technical Report 30. Como: Department of Conservation and Land Management. Halse, S.A., J.K. Ruprecht and A.M. Pinder 2003 Salinisation and prospects for biodiversity in rivers and wetlands of south-west Western Australia. Australian Journal of Botany 51(6):673–688. Hammond, J.E. 1933. Winjan’s people. Perth: Imperial Printing Company. Harpending, H. and H. Davis 1977 Some implications for hunter-gatherer ecology derived from the spatial structure of resources. World Archaeology 8(3):275–286. Harrison, R. 2000 ‘Nowadays with glass’: regional variation in Aboriginal bottle glass artefacts from Western Australia. Archaeology in Oceania 35:34–47. Harrison, S.P. 1993 Late Quaternary lake-level changes and climates of Australia. Quaternary Science Reviews 12(4):211–231. Hassell, E. 1975 My dusky friends. Dalkeith: C.W. Hassell. Hatton, T.J., J. Ruprecht and R.J. George 2003 Preclearing hydrology of the Western Australian wheatbelt: target for the future? Plant and Soil 257:341–356. Hawkes, K. and J. O’Connell 1992 On optimal foraging models and subsistence transitions. Current Anthropology 33(1):63–66. Hays, J.D., J. Imbrie and N.J. Shackleton 1976 Variations in the earth’s orbit: pacemaker of the Ice Ages. Science 194(4270):1121–1132. Hefferan, K. and J. O’Brien 2010 Earth minerals. West-Sussex: Wiley-Blackwell. Henrich, J. 2004 Demography and cultural evolution: why adaptive cultural processes produced maladaptive losses in Tasmania. American Antiquity 69:197– 218. https://dx.doi.org/10.2307/4128416 358 Hercock, M. (ed.) 2014 The Western Australian explorations of John Septimus Roe 1829–1849. Carlisle: Hesperian Press. Hesse, P.P., J.W. Magee and S. van der Kaars 2004 Late Quaternary climates of the Australian arid zone: a review. Quaternary International 118:87–102. Hillel, D. 2004 Introduction to environmental soil physics. San Diego: Academic Press. Hiscock, P. 1988 Prehistoric settlement patterns and artefact manufacture at Lawn Hill, northwest Queensland. Unpublished PhD thesis, Department of Anthropology and Sociology, University of Queensland, St Lucia. Hiscock, P. 1993 Bondaian technology in the Hunter Valley, New South Wales. Archaeology in Oceania 28:65–76. Hiscock, P. 1994 Technological responses to risk in Holocene Australia. Journal of World Prehistory 8(3):267–292. Hiscock, P. 1996 Mobility and technology in the Kakadu coastal wetlands. IndoPacific Prehistory Association Bulletin 15:151–157. Hiscock, P. 2002a Pattern and context in the Holocene proliferation of backed artifacts in Australia. In R.G. Elston and S.L. Kuhn (eds), Thinking small: global perspectives on microlithization, pp.163–177. Archaeological Papers of the American Anthropological Association 12. Washington: American Anthropological Association. Hiscock, P. 2002b Quantifying the size of artefact assemblages. Journal of Archaeological Science 29(3):251–258. Hiscock, P. 2006 Blunt and to the point: changing technological strategies in Holocene Australia. In I. Lilley (ed.), Archaeology of Oceania: Australia and the Pacific Islands, pp.69–95. Malden: Blackwell Publishing. Hiscock, P. 2007 Looking the other way. A materialist/technological approach to classifying tools and implements, cores and retouched flakes. In S. McPherron and J. Lindley (eds), Tools or cores? The identification and study of alternative core technology in lithic assemblages, pp.198–222. Newcastle: Cambridge Scholars Publishing. Hiscock, P. 2013 Occupying new lands: global migrations and cultural diversification with particular reference to Australia. In K.E. Graf, C.V. Ketron and M.R. Waters (eds), Paleoamerican odyssey, pp.3–11. College Station Texas: Department of Anthropology, Texas A&M University. Hiscock, P. 2015 Dynamics of knapping with bipolar techniques: modeling transitions and the implications of variability. Lithic Technology 40(4):342–348. https://doi.org/10.1179/2051618515Y.0000000021 359 Hiscock, P. and C. Clarkson 2000 Analysing Australian stone artefacts: an agenda for the twenty first century. Australian Archaeology 50:98–108. Hobbs, R.J. 2003 The wheatbelt of Western Australia. Pacific Conservation Biology 9(1):9–11. Hogg, A., Hua, Q., Blackwell, P., Niu, M., Buck, C., Guilderson, T., Heaton, T., Palmer, J., Reimer, P., Reimer, R., Turney, C. and S. Zimmerman 2013 SHCal13 Southern Hemisphere Calibration, 0–50,000 Years cal BP. Radiocarbon 55(4):1889– 1903. https://doi.org/10.2458/azu_js_rc.55.16783 Holdaway, S. and N. Stern 2004. A record in stone: the study of Australia's flaked stone artefacts. Canberra: Aboriginal Studies Press. Hollick, A. 2004 South East Humps focus catchment report. Research Management Technical Report 291. Perth: Department of Agriculture and Food. Hopper, S.D. 2000 Creation of conservation reserves and managing fire on granite outcrops – a case study of Chiddarcooping Nature Reserve in the Western Australian wheatbelt. Journal of the Royal Society of Western Australia 83:173–186. Hopper, S.D. 2001 Plant list for Hyden Rock. In C.R. Twidale and J.A. Bourne (eds), Field Guide to Hyden Rock, including Wave Rock, pp.64–68. Hyden: Wave Rock Management Pty Ltd. Hopper, S.D., A.P. Brown and N.G. Marchant 1997 Plants of Western Australian granite outcrops. Journal of the Royal Society of Western Australia 80(3):141–158. Hopper, S.D. and H. Lambers 2014 Human relationships with and use of Kwongkan plants and lands. In H. Lambers (ed.), Plant life on the sandplains in southwest Australia: a global biodiversity hotspot, pp.285–306. Crawley: UWA Publishing. Hopper, S.D., S. van Leeuwen, A.P. Brown and S.J. Patrick 1990 Western Australia's endangered flora and other plants under consideration for declaration. Perth: Department of Conservation and Land Management. Horn, H.S. 1968 The adaptive significance of colonial nesting in the Brewer’s Blackbird (Euphagus cyanocephalus). Ecology 49:682–694. Horne, G. and G. Aiston 1928 Savage life in Central Australia. London: Macmillan. Host, J. and C. Owen 2009 “It’s still in my heart, this is my country”: the single Noongar claim history. Perth: University of Western Australia Press. Houser, P.R. 2003 Infiltration and soil moisture processes. In T.D. Potter and B.R. Colman (eds), Handbook of weather, climate and water: atmospheric, chemistry, hydrology and societal impacts, pp.493–506. Hoboken: John Wiley and Sons. Itzstein-Davey, F. 2004 An Early Holocene palaeoenvironmental record from Two Mile Lake, South-Western Australia. Australian Geographer 35(3):317–332. 360 Jeske, R.J. and R. Lurie 1993 The archaeological visibility of bipolar technology: an example from the Koster site. Midcontinental Journal of Archaeology 18(2):131–160. Johnson, D. 2009 The geology of Australia. Melbourne: Cambridge University Press. Jones, R. 1969 Fire-stick farming. Australian Natural History 16:224–228. Kamminga, J. 1980 A functional investigation of Australian microliths. The Artefact 5:1–18. Kelleher, M.H. 2002 Archaeology of sacred space: the spatial nature of religious behaviour in the Blue Mountains National Park, Australia. Unpublished PhD thesis, University of Sydney, Sydney. Kelly, R.L. 2013 The lifeways of hunter-gatherers: the foraging spectrum (2nd ed.). New York: Cambridge University Press. Kendrick, G.W. 1977 Middle Holocene marine molluscs from near Guildford, Western Australia, and evidence for climate change. Journal of the Royal Society of Western Australia 59(4):97–104. Kershaw, P., P. Moss and S. van der Kaars 2003 Causes and consequences of long-term climatic variability on the Australian continent. Freshwater Biology 48(7):1274–1283. Kershaw, A.P. and G.C. Nanson 1993 The last full glacial cycle in the Australian region. Global and Planetary Change 7(1):1–9. Kitchener, D.J. and A. Chapman 1977 Mammals of Bendering and West Bendering nature reserves. Records of the Western Australian Museum Supplement 5:17–29. Kitchener, D.J., A. Chapman and B.G. Muir 1980 The conservation value for mammals of reserves in the Western Australian wheatbelt. Biological Conservation 18(3):179–207. Knutsson, H., K. Knutsson, N. Taipale, M. Tallavaara and K. Darmark 2015 How shattered flakes were used: micro-wear analysis of quartz flake fragments. Journal of Archaeological Science: Reports 2:517–531. Kuhn, S.L. 1990 A geometric index of reduction for unifacial stone tools. Journal of Archaeological Science 17:585–593. Kuhn, S.L. 1995 Mousterian lithic technology: an ecological perspective. New Jersey: Princeton University Press. Kuhn, S.L. and D.S. Miller 2015 Artifacts as patches: the marginal value theorem and stone tool life histories. In N. Goodale and W. Andrefsky (eds), Lithic technological systems and evolutionary theory, pp.172–197. New York: Cambridge University Press. 361 Laing, I.A.F. and E.J. Hauck 1997 Water harvesting from granite outcrops in Western Australia. Journal of the Royal Society of Western Australia 80(3):181–184. Landgate 2015 Historical town names. Retrieved 20 February 2016 from < http://www0.landgate.wa.gov.au/maps-and-imagery/wa-geographic-names/namehistory/historical-town-names#H > Landor, H. and H.M. Lefroy 1843 Landor and Lefroy’s expedition to the south-east of York, January 1843 (transcript). In M. Herlock and S. Milentis (eds), Western Australian exploration 1836–1845, pp.327–333. Perth: Hesperian Press. Langford-Smith, T. 1978 A select review of silcrete research in Australia. In T. Langford-Smith (ed.), Silcrete in Australia, pp.1–11. Armidale: University of New England. Laskar, J., P. Robutel, F. Joutel, M. Gastineau, A.C.M. Correia and B. Levard 2004 A long term numerical solution for the insolation quantities of the Earth. Astronomy and Astrophysics 428(1):261–285. Leake, B.W. 1950. Countryman. The Western Mail Supplement, 17 August 1950. Leigh, J., R. Boden and J. Briggs 1984 Extinct and endangered plants of Australia. Melbourne: MacMillan. Lewis, S.E., C.R. Sloss, C.V. Murray-Wallace, C.D. Woodroffe and S.G. Smithers 2013 Post-glacial sea-level changes around the Australian margin: a review. Quaternary Science Reviews 74:115–138. Lin, S.C., C.M. Pop, H.L. Dibble, W. Archer, D. Desta, M. Weiss and S.P. McPherron 2016 A core reduction experiment finds no effect of original stone size and reduction intensity on flake debris size distribution. American Antiquity 81(3):562–575. Low, T. 1991 Wild food plants of Australia. Sydney: HarperCollins. Lowe, L.D., J.A. Webb, R.J. Nathan, T. Etchells, H.M. Malano 2009 Evaporation from water supply reservoirs: an assessment of uncertainty. Journal of Hydrology 376:261–274. Luke, G.J., K.L. Burke and T.M. O’Brien 1987 Evaporation data for Western Australia. Resource Management Technical Report 65. Perth: Department of Agriculture and Food. Lullfitz, A., J. Dortch, S.D. Hopper, C. Pettersen, R. Reynolds and D. Guilfoyle 2017 Human niche construction: Noongar evidence in pre-colonial southwestern Australia. Conservation and Society 15(2):201–216. MacArthur, R.H. and E.R. Pianka 1966 On optimal use of a patchy environment. The American Naturalist 100(916):603–609. 362 Macintyre, K., B. Dobson, J. Harris and G. Quartermaine 1992 Report on an ethnographic and archaeological survey of the proposed KondininBounty/Forrestania mines power line. Unpublished report prepared for the State Energy Commission of Western Australia (SECWA). Mackay, A. 2005 Informal movements: changing mobility patterns at Ngarrabullgan, Cape York Australia. In C. Clarkson and L. Lamb (eds), Lithics' down under': Australian perspectives on lithic reduction, use and classification, pp.95–107. BAR International Series 1408. Oxford: Archaeopress. Magee, J.W. 2009 Palaeovalley groundwater resources in arid and semi-arid Australia: a literature review. Geoscience Australia Record 2009/03. Canberra: National Water Commission and Geoscience Australia. Manninen, M.A. and K. Knutsson 2014 Lithic raw material diversification as an adaptive strategy – technology, mobility, and site structure in Late Mesolithic northernmost Europe. Journal of Anthropological Archaeology 33:84–98. Martin, H.A. 2006 Cenozoic climatic change and the development of the arid vegetation in Australia. Journal of Arid Environments 66(3):533–563. Martindale, A. and I. Jurakic 2006 Identifying expedient glass tools from a postcontact Tsimshian village using low power (10–100×) magnification. Journal of Archaeological Science 33(3):414–427. Martinez Cortizas, A. and C. Llana 1996 Morphostructural variables and the analysis of their effects on quartz blank characteristics. In N. Moloney, L. Raposo and M. Santonja (eds), Non-flint stone tools and the Palaeolithic occupation of the Iberian Peninsula, pp.49–53. BAR International Series 649. Oxford: Archaeopress. Mayewski, P.A., E.E. Rohling, J.C. Stager, W. Karlen, K.A. Maasch, L.D. Meeker, E.A. Meyerson, F. Gasse, S. van Kreveld, K. Holmgren, J. Lee-Thorp, G. Rosqvist, F. Rack, M. Staubwasser, R.R. Schneider and E.J. Steig 2004. Holocene climate variability. Quaternary Research 62:243–255. McBryde, I. 1974. Aboriginal Prehistory in New England: an archaeological survey of Northeastern New South Wales. Sydney: Sydney University Press. McBryde, I. 1979 Ethnohistory in an Australian context: independent discipline or convenient data quarry? Aboriginal History 3(1/2):128–151. McBryde, I. 1986 Australia’s Once and Future Archaeology. Archaeology in Oceania 21(1):13–28. McDonald, J.L., D. Donlon, J.H. Field, R.L.K. Fullagar, J. Brenner Coltrain, P. Mitchell and M. Rawson 2007 The first archaeological evidence for death by spearing in Australia. Antiquity 81:877–885. 363 McDonald, J., W. Renyen, F. Petchey, K. Ditchfield, C. Byrne. D. Vannieuwenhuyse, M. Leopold and P. Veth 2018 Karnatukul (Serpent’s Glen): a new chronology for the oldest site in Australia’s Western Desert. PLOS One 13(9):e0202511. https://doi.org/10.1371/journal.pone.0202511 McKenzie, N.L., A.A. Burbidge and N.G. Marchant 1973 Results of a biological survey of a proposed wildlife sanctuary at Dragon Rocks near Hyden, Western Australia. Department of Fisheries and Fauna Western Australia Report no. 12. Perth: Department of Fisheries and Fauna. McKenzie, N.L., N. Gibson, G.J. Keighery and J.K. Rolfe 2004 Patterns in the biodiversity of terrestrial environments in the Western Australian wheatbelt. Records of the Western Australian Museum Supplement 67:293–335. McMahon, T.A., M.C. Peel, L. Lowe, R. Srikanthan and T.R. McVicar 2013 Estimating actual, potential, reference crop and pan evaporation using standard meteorological data: a pragmatic synthesis. Hydrology and Earth System Sciences 17:1331–1363. Meagher, S.J. 1974 The food resources of the Aborigines of the south-west of Western Australia. Records of the Western Australian Museum 3:14–65. Meeking, J. 1979. The history of Hyden. Northam: Mid-West Print. Mitchell, M.B. 2016 The Esperance Nyungars, at the Frontier: an archaeological investigation of mobility, aggregation and identity in late-Holocene Aboriginal society, Western Australia. Unpublished PhD thesis, The Australian National University, Canberra. Mitchell, P.J., E. Veneklaas, H. Lambers, S.S.O. Burgess 2009 Partitioning of evapotranspiration in a semi-arid eucalupt woodland in south-western Australia. Agricultural and Forest Meteorology 149:25–37. Mithen, S.J. 1989 Modeling hunter-gatherer decision making: complementing Optimal Foraging Theory. Human Ecology 17(1):59–83. Moore, G. (ed.) 2001 Soil guide: a handbook for understanding and managing agricultural soils. Agriculture Western Australia Bulletin No. 4343. Perth: National Landcare and Agriculture Western Australia. Moore, G.F. 1884 Diary of ten years eventful life of an early settler in Western Australia, and also a descriptive vocabulary of the language of the Aborigines. London: Walbrook. Moros, M., P. De Dekker, E. Jansen, K. Perner and R.J. Telford 2009 Holocene climate variability in the Southern Ocean recorded in a deep-sea sediment core off South Australia. Quaternary Science Reviews 28(19):1932–1940. Morse, K., R. Cameron and W. Reynen 2014 A tale of three caves: new dates for Pleistocene occupation in the inland Pilbara. Australian Archaeology 79:167–178. 364 Mouritz, M. 1986. Only the wind was free: 60 years in Wave Rock country. Canberra: National Library of Australia. Muir, B.G. 1977 Biological survey of the Western Australian wheatbelt, part 2: vegetation and habitat of Bendering Reserve. Records of the Western Australian Museum Supplement 3:1–142. Muir, B.G. 1978 Some nature reserves of the Western Australian wheatbelt, part 4: Narembeen Shire. Unpublished report prepared for the Department of Fisheries and Wildlife. Muir, B.G. 1979 Some nature reserves of the Western Australian wheatbelt, part 20: Kondinin Shire. Unpublished report prepared for the Department of Fisheries and Wildlife. Murphy-White, S. 2007 Farm water for the north-eastern and eastern wheatbelt of Western Australia (zones 2 and 5). Resource Management Technical Report 318. Perth: Department of Agriculture and Food. Murphy-White, S. and E. Hauck 1999 Agricultural water supply areas of the northeastern and eastern wheatbelt, Western Australia. In Water 99 joint congress – Brisbane, Australia 6–8 July 1999. Barton: Institute of Engineers, Australia. Murphy-White, S. and P. Leoni 2006 Lockhart catchment appraisal 2005. Resource Management Technical Report 310. Perth: Department of Agriculture and Food. Myers, F.R. 1982 Always ask: resource use and land ownership among Pintupi Aborigines of the Australian Western Desert. In N.M. Williams and E.S. Hunn (eds), Resource managers, pp.173–196. Canberra: Australian Institute of Aboriginal Studies. Nanson, G.C., D.M. Price and S.A. Short 1992 Wetting and drying of Australia over the past 300 ka. Geology 20(9):791–794. Native Grass Resources Group (n.d.) Native grass propagation. Retrieved 27 March 2017 from < https://nativegrassresourcesgroup.files.wordpress.com/2016/08/nativegrass-propagation.pdf > Nelson, M.C. 1991 The study of technological organization. Archaeological Method and Theory 3:57–100. Nettle, D., M.A. Gibson, D.W. Lawson and R. Sear 2013 Human behavioural ecology: current research and future prospects. Behavioral Ecology 24(5):1031–1040. https://doi.org/10.1093/beheco/ars222 Nevill, S. 2014 Guide to the wildlife of Perth and the south west. Perth: Simon Nevill Publications. Neville, A.O. 1948 Contributory causes of Aboriginal depopulation in Western Australia. Mankind 4:3–13. 365 Newsome, J.C. and E.J. Pickett 1993 Palynology and palaeoclimatic implications of two Holocene sequences from southwestern Australia. Palaeogeography, Palaeoclimatology, Palaeoecology 101(3):245–261. NHMRC, NRMMC 2011 Australian drinking water guidelines paper 6 national water quality management strategy. Canberra: National Health and Medical Research Council and National Resource Management Ministerial Council. Nussinovitich, A. 2010 Plant gum exudates of the world: sources, distribution, properties, and applications. Boca Raton: CRC Press. O’Connell, J.F. and K. Hawkes 1981 Alyawara plant use and optimal foraging theory. In B. Winterhalder and E.A. Smith (eds), Hunter-gatherer foraging strategies: ethnographic and archaeological analyses, pp.99–125. Chicago: University of Chicago Press. O’Connell, J.F., P.K. Latz and P. Barnett 1983 Traditional and modern plant use among the Alyawara of Central Australia. Economic Botany 37(1):80–109. O’Connor, M.H. and S.M. Prober, S.M. 2010 A calendar of Ngadju seasonal knowledge. Floreat: CSIRO Sustainable Ecosystems. O'Connor, R., G. Quartermaine and C. Bodney 1989 Report on an investigation into the Aboriginal significance of wetlands and rivers in the Perth-Bunbury region. Perth: Western Australian Water Resources Council. Oliver, Y.M. and M.J. Robertson 2009 Quantifying the benefits of accounting for yield potential in spatially and seasonally responsive nutrient management in a Mediterranean climate. Australian Journal of Soil Research 47:114–126. Orians, G.H. and N.E. Pearson 1979 On the theory of central place foraging. In D.J. Horn, R.D. Mitchell, and G.R. Stairs (eds), Analysis of ecological systems, pp.154– 177. Columbus: Ohio State University. Paczkowska, G. and A.R. Chapman 2000 The Western Australian flora: a descriptive catalogue. Perth: Wildflower Society of Western Australia, Western Australian Herbarium, CALM and the Botanic Gardens and Parks Authority. Pargeter, J. and P. de la Pena 2017 Milky quartz bipolar reduction and lithic miniaturization: experimental results and archaeological implications. Journal of Field Archaeology 42(6):551–565. https://doi.org/10.1080/00934690.2017.1391649 Pargeter, J. and M.I. Eren 2017 Quantifying and comparing bipolar versus freehand flake morphologies, production currencies, and reduction energetics during lithic miniaturization. Lithic Technology 2017. https://doi.org/10.1080/01977261.2017.1345442 Parry, W. and R.J. Kelly 1987 Expedient core technology and sedentism. In J.K. Johnson and C.A. Morrow (eds), The organisation of core technology, pp.285–304. Boulder: Westview Press. 366 Pascoe, B. 2013 Dark emu, black seeds: agriculture or accident? Broome: Magabala Books. Pate, J.S. and K.W. Dixon 1982 Tuberous, cormous and bulbous plants: biology of an adaptive strategy in Western Australia. Crawley: University of Western Australia Press. Paterson, A. 1999 Confronting the sources: the archaeology of culture-contact in the south-western Lake Eyre Basin, Central Australia. Unpublished PhD Thesis, Department of Archaeology, University of Sydney, Camberdown/Darlington. Pearce, R.H. and M. Barbetti 1981 A 38,000-year-old archaeological site at Upper Swan, Western Australia. Archaeology in Oceania 16:173–178. Peck, C. 1982 Monitoring and conservation work on sites in south-west – September 1982. Unpublished report prepared for the Western Australian Museum. Pereira, L.S., R.G. Allen, M. Smith and D. Raes 2015 Crop evapotranspiration estimation with FAO56: Past and future. Agricultural Water Management 147:4–20. Perkins, D. 2011 Mineralogy (3rd ed.). Essex: Pearson. Petherick, L., H. Bostock, T.J. Cohen, K. Fitzsimmons, J. Tibby, M.-S. Fletcher, P. Moss, J. Reeves, S. Mooney, T. Barrows, J. Kemp, J. Jansen, G. Nanson, A. Dosseto 2013 Climatic records over the past 30 ka from temperate Australia – a synthesis from the Oz-INTIMATE workgroup. Quaternary Science Reviews 74:58– 77. Pickett, E.J. 1997 The late Pleistocene and Holocene vegetation history of three lacustrine sequences from the Swan Coastal Plain, southwestern Australia. Unpublished PhD thesis, Department of Geography, University of Western Australia, Crawley. Pickett, E.J., S.P. Harrison, G. Hope, K. Harle, J.R. Dodson, A.P. Kershaw, I.C. Prentice, J. Backhouse, E.A. Colhoun, D. D’Costa, J. Flenley, J. Grindrod, S. Haberle, C. Hassell, C. Kenyon, M. Macphail, H. Martin, A.H. Martin, M. McKenzie, J.C. Newsome, D. Penny, J. Powell, J.I. Raine, W. Southern, J. Stevenson, J-P. Sutra, I. Thomas, S. van der Kaars, J. Ward 2004 Pollen-based reconstructions of biome distributions for Australia, Southeast Asia and the Pacific (SEAPAC region) at 0, 6000 and 18,000 14C yr BP. Journal of Biogeography 31(9):1381–1444. Pinder. A.M., S.A. Halse, J.M. McRae and R.J. Shiel 2004 Aquatic invertebrate assemblages of wetlands and rivers in the Wheatbelt region of Western Australia. Records of the Western Australian Museum Supplement 67:7–37. Prober, S.M., E. Yuen, M.H. O'Connor and L. Schultz 2016 Ngadju kala: Australian Aboriginal fire knowledge in the Great Western Woodlands. Austral Ecology 41:716– 732. 367 Quartermaine, G. 2000 Report on an archaeological investigation of Aboriginal sites Corrigin to Hyden section of the Brookton Highway. Unpublished report prepared for Halpern Glick Maunsell. Quigley, M.C., T. Horton, J.C. Hellstrom, M.L. Cupper and M. Sandiford 2010 Holocene climate change in arid Australian from speleothem and alluvial records. The Holocene 20(7):1093–1104. Radcliffe-Brown, A.R. 1930 The social organization of Australian Tribes (part 1). Oceania 1(1):34–63. Rafferty. J.P. (ed.) 2012 Minerals. New York: Brittanica Educational Publishing. Ranatunga, K., E.R. Nation and D.G. Barratt 2008 Review of soil water models and their applications in Australia. Environmental Modelling and Software 23(9):1182– 1206. Randolph, P. 1973. Bates Cave, Hyden. Unpublished report prepared for the Department of Aboriginal Sites, Western Australian Museum, Perth. Reeves, J.M., T.D. Barrows, T.J. Cohen, A.S. Kiem, H.C. Bostock, K.E. Fitzsimmons, J.D. Jansen, J. Kemp, C. Krause, L. Petherick, S.J. Phipps 2013 Climate variability over the last 35,000 years recorded in marine and terrestrial archives in the Australian region: an OZ-INTIMATE compilation. Quaternary Science Reviews 74:21–34. Rick, A. 2011 Survey and analysis of plant communities growing on gypsum in the Western Australian wheatbelt. Unpublished report prepared for the Wheatbelt NRM Region and the Department of Environment and Conservation, Western Australia. Ritchie, J.T. 1972 Model for predicting evaporation from a row crop with incomplete cover. Water Resources Research 8(5):1204–1213. Robertson, G. 2005. Backed artefact use in eastern Australia: a residue and usewear analysis. Unpublished PhD thesis, School of Social Science, University of Queensland, St Lucia. Robertson, G. 2009 Aboriginal craft and subsistence activities at Native Well I and Native Well II, Central Western Highlands, Queensland: results of a residue and usewear analysis of backed artefacts. In M. Haslam, G. Robertson, A. Crowther, S. Nugent and L. Kirkwood (eds), Archaeological science under a microscope: studies in residue and ancient DNA analysis in honour of Thomas H. Loy, pp.239–257. Terra Australis 30. Canberra: ANU E Press. Robertson, G. and V. Attenbrow 2008 Skin-working at Emu Tracks 2, New South Wales, Australia: an integrated residue and use-wear analysis of backed artefacts. Lithic Technology 33:31–49. https://doi.org/10.1080/01977261.2008.11721059 Robertson, G., V. Attenbrow and P. Hiscock 2009 Multiple uses for Australian backed artefacts. Antiquity 83:296–308. 368 Roe, J.S. 1836. Explorations east of York to Wongan Hills, October to November 1836, field books (transcript). In M. Hercock (ed.), The Western Australian explorations of John Septiumus Roe 1829–1849, pp.236–297. Perth: Hesperian Press. Roe, J.S. 1852 Report of an expedition under the Surveyor-General, Mr J.S. Roe, to the south-eastward of Perth, in Western Australia, between the months of September, 1848, and February, 1849, to the Hon. The Colonial Secretary. Journal of the Royal Geographical Society of London 22:1–57. Rolland, N. and H.L. Dibble 1990 A new synthesis of Middle Palaeolithic variability. American Antiquity 55(3):480–499. Rossi, A.M. 2010 An archaeological re-investigation of the Mulka’s Cave Aboriginal rock art site, near Hyden, southwestern Australia. Unpublished MA thesis, School of Arts, Edith Cowan University, Joondalup. Rossi, A.M. 2014a Re-evaluating the antiquity of Aboriginal occupation at Mulka’s Cave. Australian Archaeology 78:39–46. Rossi, A.M. 2014b Unconventional proxies: using archaeological measures of occupation intensity to infer Holocene climatic variability in inland southwestern Australia. Paper presented at the Australian Archaeological Association conference, Cairns. Rossi, A.M. and R.E. Webb 2007 The consequences of allowing unfettered tourist access at an Aboriginal site in a fragile environment: the erosive effect of trampling. Conservation and Management of Archaeological Sites 9:219–236. Rossi, A.M. and Webb, R.E. 2008 The erosive effect of tourism at an Aboriginal rock art site on the western edge of the arid zone in south-western Australia. Antiquity 82 Project Gallery. Retrieved 10 January 2014 from < http://www.antiquity.ac.uk/projgall/rossi315/ > Runnels, C. 1976 More on glass implements from Greece. Newsletter of Lithic Technology 5(3):27–31. Salama, R.B. 1994 The evolution of saline lakes in the relict drainage of the Yilgarn River, Western Australia. In R.W. Renaut and W.M. Last (eds), Sedimentology and geochemistry of modern and ancient saline lakes, pp.189–199. SEPM Special Publication no. 50. Tulsa: Society for Sedimentary Geology. Saunders, D.A. 1989 Changes in the avifauna of a region, district and remnant as a result of fragmentation of native vegetation: the wheatbelt of Western Australia. A case study. Biological Conservation 50:99–135. Saunders, D.A. and J.A. Ingram 1995 Birds of southwestern Australia: an atlas of changes in distribution and abundance of wheatbelt avifauna. Chipping Norton: Surrey Beatty and Sons. 369 Scadding, R., V. Winton and V. Brown 2015 An LA-ICP-MS trace element classification of ochres in the Weld Range environ, Mid West region, Western Australia. Journal of Archaeological Science 54:300–312. https://doi.org/10.1016/j.jas.2014.11.017 Schoknecht, N.R. and S. Pathan 2013 Soil groups of Western Australia: a simple guide to the main soils of Western Australia (4th ed.). Resource Management Technical Report 380. Perth: Department of Agriculture and Food. Schut, A.G.T., G.W. Wardell-Johnson, C.J. Yates, G. Keppel. I. Barab, S.E. Franklin, S.D. Hopper, K.P. Van Niel, L. Mucina and M. Byrne 2014 Rapid characterisation of vegetation structure to predict refugia and climate change impacts across a global biodiversity hotspot. PLoS ONE 9(1):e82778. https://dx.doi.org/10.1371/journal.pone.0082778 Semeniuk, V. 1986 Holocene climate history of coastal southwestern Australia using calcrete as an indicator. Palaeogeography, Palaeoclimatology, Palaeoecology 53(2):289–308. Serventy, V.L. 1952 Cave paintings near York and Hyden. The Western Australian Naturalist 3:121–131. Shaw, E. 2014 Hydrology in Practice (3rd ed.). London: CRC Press. Shennan, S.J. 2001 Demography and cultural innovation: a model and some implications for the emergence of modern human culture. Cambridge Archaeological Journal 11:5–16. Shott, M. 1986 Technological organization and settlement mobility: an ethnographic examination. Journal of Anthropological Research 42(1):15–51. Simpson, K. and N. Day 2004 Field guide to the birds of Australia (7th ed.). Campberwell: Viking. Slack, M.J., R.L.K. Fullagar, J.M. Field and A. Border 2004 New Pleistocene ages for backed artefact technology in Australia. Archaeology in Oceania 39:131–137. Smith, E.A. 1979 Human adaptation and energetic efficiency. Human Ecology 7(1):53–74. Smith, E.A. 1983 Anthropological applications of optimal foraging theory: a critical review (including comments and reply). Current Anthropology 24(5):625–651. Smith, M.A. 2013 The archaeology of Australia's deserts. Cambridge: Cambridge University Press. Smith, M.V. 1993 Recherche à l’Espérance: a prehistory of the Esperance region of South-Western Australia. Unpublished PhD thesis, Department of Anthropology, University of Western Australia, Crawley. 370 Smith, M.V. 2011 Moving on: an archaeological record of mobility in the Esperance area of south-western Australia. Records of the Western Australian Museum Supplement 79:16–29. Speal, C.S. 2009 The economic geography of chert lithic production in the southern Maya lowlands: a comparative examination of early-stage reduction debris. Latin American Antiquity 20(1):91–119. Stannage, C.T. (ed.) 1981. A new history of Western Australia. Crawley: University of Western Australia Press. Stanner, W.E.H. 1965 Aboriginal territorial organization: estate, range, domain and regime. Oceania 36(1):1–26. Staub, A. and J. Taylor 1999 Report on an Aboriginal heritage survey: proposed Hyden to Kondinin water pipeline. Unpublished report prepared for WML consultants. Storr, G.M. 1991 Birds of the south-west division of Western Australia. Records of the Western Australian Museum Supplement 35:1–150. Sullivan, A.P. and K.C. Rozen 1985. Debitage analysis and archaeological interpretation. American Antiquity 50(4):755–779. Suzuki, H., Y. Uesugi, K. Endo, H. Ohmori, K. Takeuchi and K. Twasaki 1982. Studies in the Holocene and recent climatic fluctuations in Australia and New Zealand. Tokyo: University of Tokyo Press. Sweedman, L. and D. Merritt 2006 Australian seeds: a guide to their collection, identification and biology. Collingwood: CSIRO Publishing. Taipale, N., K. Knutsson and H. Knutsson 2014 Unmodified quartz flake fragments as cognitive tool categories: testing the wear preservation, previous low magnification use-wear results and criteria for tool blank selection in two Late Mesolithic quartz assemblages from Finland. In J. Marreiros, N. Bicho and J.F. Gibaja (eds), International conference on use-wear analysis: use-wear 2012, pp.352–361. Newcastle upon Tyne: Cambridge Scholars Publishing. Tallavaara, M., M.A. Manninen, E. Hertell and T. Rankama 2010 How flakes shatter: a critical evaluation of quartz fracture analysis. Journal of Archaeological Science 37:2442–2448. Tanny, J., S. Cohen, S. Assouline, F. Lange, A. Grava, D. Berger, BB. Teltch and M.B. Parlange 2008 Evaporation from a small water reservoir: direct measurements and estimates. Journal of Hydrology 351:218–229. Taylor, G. and R.A. Eggleton 2017 Silcrete: an Australian perspective. Australian Journal of Earth Sciences 2017. https://doi.org/10.1080/08120099.2017.1318167 Timms, B.V. 2013 Geomorphology of pit gnammas in southwestern Australia. Journal of the Royal Society of Western Australia 96:7–16. 371 Tindale, N. 1941 A Tasmanian stone implement made from bottle glass. Papers and Proceedings of the Royal Society of Tasmania 1941:1–2. Tindale, N.B. 1974 Aboriginal tribes of Australia: their terrain, environmental controls, distribution, limits and proper names. Berkeley: University of California Press. Tonkinson, R. 1978 The Mardudjara Aborigines: living the dream in Australia's Western Desert. New York: Holt, Rinehart and Winston. Torrence, R. 1983 Time budgeting and hunter-gatherer technology. In G. Bailey (ed.), Hunter-gatherer economy in prehistory, pp.11–22. Cambridge: Cambridge University Press. Torrence, R. 1989 Retooling: towards a behavioral theory of stone tools. In R. Torrence (ed.), Time, energy and stone tools, pp.57–66. Cambridge: Cambridge University Press. Torrence, R. 2001 Hunter-gatherer technology: macro- and microscale approaches. In C. Panter-Brick, R.H. Layton and P. Rowley-Conwy (eds), Hunter-gatherers: an interdisciplinary perspective, pp.73–98. Cambridge: Cambridge University Press. Treble, P.C., J.M.G. Shelley and J. Chappell 2003 Comparison of high-resolution sub-annual records of trace elements in a modern (1911–1992) speleothem with instrumental climate data from southwest Australia. Earth and Planetary Science Letters 216:141–153. Trigger, B.G. 1982 Ethnohistory: problems and prospects. Ethnohistory 29(1):1–19. Turney, C.S.M., M. Bird, L.K. Fifield, R.G. Roberts, M. Smith, C.E. Dortch, R. Grun, E. Lawson, L.K. Ayliffe, G.H. Miller, J. Dortch and R.G. Cresswell 2001 Early human occupation at Devils Lair, southwestern Australia, 50,000 years ago. Quaternary Research 55:3–13. Turney, C.S.M., S. Haberle, D. Fink, A.P. Kershaw, M. Barbetti, T.T. Barrows, M. Black, T.J. Cohen, T. Corrège, P.P. Hesse, Q. Hua, R. Johnston, V. Morgan, P. Moss, G. Nanson, T. Van Ommen, S. Rule, N.J. Williams, J-X. Zhao, D. D’Costa, YX. Feng, M. Gagan, S. Mooney and Q. Xia 2006 Integration of ice-core, marine and terrestrial records for the Australian Last Glacial Maximum and Termination: a contribution from the OZ INTIMATE group. Journal of Quaternary Science 21(7):751–761. Veth, P. 1987 Martutjarra prehistory: variation in arid zone adaptations. Australian Archaeology 25:102–111. Veth, P. and S. O’Connor 2005 Archaeology, claimant connection to sites, and Native Title: employment of successful categories of data with specific comments on glass artifacts. Australian Aboriginal Studies 2005(1):2–15. 372 Wallis, L., J. Huntley, M. Marsh, A. Watchman and A. Ewen 2016 PXRF analysis of a yellow ochre quarry and rock art motifs in the Central Pilbara. Journal of the Anthropological Society of South Australia 40:134–155. Wanner, H., J. Beer, J. Bütikofer, T.J. Crowley, U. Cubasch, J. Flückiger, H. Goosse, M. Grosjean, F. Joos, J.O. Kaplan, M. Küttel, S.A. Müller, I.C. Prentice, O. Solomina, T.F. Stocker, P. Tarasov, M. Wagner and M. Widmann 2008 Mid- to Late Holocene climate change: an overview. Quaternary Science Reviews 27(19):1,791–1,828. Watts, I. 2010 The pigments from Pinnacle Point Cave 13B, Western Cape, South Africa. Journal of Human Evolution 59:392–411. https://dx.doi.org/10.1016/j.jhevol.2010.07.006 Webb, R.E. 2000 Reconstructing the ethnobotany of a depleted flora: food plant availability in the Murchison Basin, Western Australia, prior to European arrival. In R.A. Nicholson and T.P. O’Connor (eds), People as an agent of environmental change, pp.92–116. Oxford: Oxbow Books. Webb, R.E. and R.G. Gunn 2004 Re-recording culturally significant Aboriginal sites in south-western Australia as a guide to Noongar usage of the region in the past. Rock Art Research 21:93–97. Webb, R.E. and A.M. Rossi 2008 How was Mulka’s Cave, an Aboriginal rock art site near Hyden, in south-central Western Australia, used by the people who decorated its walls, when the present entrance was much smaller? Records of the Western Australian Museum 24:307–318. Western Australian Herbarium 1998- Florabase – the Western Australian Flora [Computer Software]. Retrieved from < https://florabase.dpaw.wa.gov.au/ > Western Australian Museum n.d. Frogwatch – Frogs – Retrieved 27 March 2017 from < http://museum.wa.gov.au/explore/frogwatch/frogs/ > White, I. (ed.) 1985 Daisy Bates: the native tribes of Western Australia. Canberra: National Library of Australia. White, P. 2011 Changing perspectives in Australian archaeology, part I. Regional archaeology in Australia. Technical Reports of the Australian Museum Online 23(1):3–5. Whittaker, J. C. 1994 Flintknapping: making and understanding stone tools. Austin: University of Texas Press. Wildlife Society of Western Australia 2013 Seed notes for Western Australia. Retrieved 27 March 2017 from < https://www.dpaw.wa.gov.au/about-us/science-andresearch/publications-resources/114-seed-notes-for-western-australia > Wildy, D.T., J.S. Pate and J.R. Bartle 2004 Budgets of water use by Eucalyptus kochii tree belts in the semi-arid wheatbelt of Western Australia. Plant and Soil 262:129–149. 373 Willey, O. 1950 In search of the wheatbelt cave of Mulka the terrible. The West Australian Saturday 28 October 1950 p. 18. Williams, A.N. and M.A. Smith 2012 AustArch 2: A database of 14C and luminescence ages from archaeological sites in the Top End. Australian Archaeology 74:146. Williams, A.N. and M.A. Smith 2013 AustArch 3: A database of 14C and luminescence ages from archaeological sites in southern Australia. Australian Archaeology 76:102. Williams, A.N., M.A. Smith, C.S.M Turney and M. Cupper 2008 AustArch1: A database of 14C and luminescence ages from archaeological sites in the Australian arid zone. Australian Archaeology 66:99. Williams, A.N., S. Ulm, M. Smith and J. Reid 2014 AustArch: A database of 14C and non-14C ages from archaeological sites in Australia – composition, compilation and review (data paper). Internet Archaeology 36. https://dx.doi.org/10.11141/ia.36.6 Williams, M., E. Cook, S. van der Kaars, T. Barrows, J. Shulmeister and P. Kershaw 2009 Glacial and deglacial climatic patterns in Australia and surrounding regions from 35,000 to 10,000 years ago reconstructed from terrestrial and near-shore proxy data. Quaternary Science Reviews 28(23–24):2398–2419. Williamson, C. 2004 Finding meaning in the patterns: the analysis of material culture from a contact site in Tasmania. In R. Harrison and C. Williamson (eds), After Captain Cook: the archaeology of the recent Indigenous past in Australia, pp.75–102. Sydney: Archaeological Computing Laboratory. Wilmsen, E.N. 1973 Interaction, spacing behavior, and the organization of hunting bands. Journal of Anthropological Research 29(1):1–31. Winmar, R. 1996 Walwalinj: the hill that cries. Nyungar language and culture. Perth: Dorothy Winmar. Winter, T.C. 1981 Uncertainties in estimating the water balance of lakes. Journal of the American Water Resources Association 17(1):82–115. https://dx.doi.org/10.1111/j.1752-1688.1981.tb02593.x Winterhalder, B. 1981 Optimal foraging strategies and hunter-gatherer research in anthropology: theory and models. In B. Winterhalder and E.A. Smith (eds), Huntergatherer foraging strategies: ethnographic and archaeological analyses, pp.13–35. Chicago: University Press. Winterhalder, B. and E.A. Smith 2000 Analyzing adaptive strategies: human behavioural ecology at twenty-five. Evolutionary Anthropology 9(2):51–72. Witt, G.B., J. Luly and R.J. Fairfax 2006 How the west was once: vegetation change in southwest Queensland from 1930–1995. Journal of Biogeography 33(9):1,585– 1,596. 374 Witter, D. 1990. Regions and resources. Unpublished PhD thesis, Prehistory Department, Research School of Pacific Studies, Australian National University, Canberra. Woinarski, J.C.Z., A.A. Burbidge and P.L. Harrison 2015 Ongoing unravelling of a continental fauna: decline and extinction of Australian mammals since European settlement. Proceedings of the National Academy of Sciences 112:4531–4540. https://dx.doi.org/10.1073/pnas.1417301112 Wolski, N. and T.H. Loy 1999 On the invisibility of contact: residue analyses on Aboriginal glass artefacts from western Victoria. The Artefact 22:65–73. Yassini, I. and G.W. Kendrick 1988 Middle Holocene ostracodes, foraminifers and environments of beds at Point Waylen, Swan River Estuary, southwestern Australia. Alcheringa 12(2):107–121. Zheng, H., C.M. Powell and H. Zhao 2002 Eolian and lacustrine evidence of late Quaternary palaeoenvironmental changes in southwestern Australia. Global and Planetary Change 35(1):75–92. 375 APPENDIX A FLORAL SPECIES PRESENT IN THE STUDY AREA The floral list was compiled using literature derived from surveys and analyses performed within the study area (Beard 1972, 1979; Coates 1992; Department of Water 2009; Gunness 2002; Halse et al. 1993; Hopper 2001; Muir 1977, 1978, 1979; Rick 2011). Species are listed alphabetically, with taxonomy and naming conventions as per Florabase (Western Australian Herbarium (1998-). Any non-current names encountered in the references above have been updated; taxonomic changes made after 1 April 2017 will not be reflected. A cross in the relevant column indicates that a plant species was found in that particular Landscape Division. Genus-level identifications (e.g. Acacia sp., Eucalyptus spp. – assumed to represent two distinct species) have generally been omitted, based on the assumption that the relevant species was likely identified in another source; they were retained or amended in certain circumstances, as below: o Entries were retained where the genus-level identification/s was the only representation of a genus in a particular Landscape Division; o Spp. was converted to sp. where the floral list included just one named species of the same genus in a particular Landscape Division, i.e. the named species could not possibly incorporate all of the genus-level identifications. Landscape Divisions have been abbreviated, as below: o GR = Granite o HEA = Heath o MAL = Mallee o SAL = Saline o TH = Thicket o WL = Woodland 376 Species Acacia acanthoclada GR HEA MAL SAL X X X X Acacia acuaria X Acacia acutata X X Acacia ancistrophylla X Acacia anfractuosa Acacia assimilis X X X X X X X X X X X X X X X Acacia bidentata X X Acacia brachyclada X Acacia brachyphylla X Acacia celastrifolia X X X Acacia chrysocephala X Acacia colletioides X Acacia consanguinea X X Acacia coolgardiensis X Acacia cupularis X Acacia densiflora Acacia dentifera X X Acacia dermatophylla X Acacia desertorum X X Acacia dielsii X X Acacia ephedroides X Acacia eremophila X X Acacia erinacea X X X X X Acacia filifolia X Acacia fragilis Acacia hemiteles X X Acacia ericifolia Acacia fauntleroyi X X X X X X X Acacia inceana X X Acacia lasiocarpa X X X X X Acacia leptopetala X Acacia leptospermoides X Acacia lineolata X Acacia lirellata X Acacia mackeyana X Acacia merinthophora X Acacia merrallii Acacia microbotrya X X Acacia intricata Acacia lasiocalyx X X X Acacia beauverdiana Acacia chrysella WL X Acacia acuminata Acacia andrewsii TH X X X X X X X 377 Species Acacia mimica GR HEA MAL X X X Acacia moirii X Acacia multispicata X SAL TH X X X Acacia myrtifolia X Acacia neurophylla X X Acacia newbeyi X Acacia pulchella X Acacia pycnocephala X Acacia rigens X Acacia rostellata X Acacia saligna X X Acacia sessilispica X X X Acacia shuttleworthii X Acacia sphacelata Acacia spinosissima X X X X X X X X X Acacia tetanophylla X Acacia tratmaniana X X X X Acacia uncinella X X Acacia verriculum X X X X Acacia viscifolia X Acacia yorkrakinensis X Acrotriche patula X X X X Actinobole uliginosum X Actinotus superbus X Adenanthos argyreus Adenanthos flavidiflorus X X X Allocasuarina acutivalvis X X X Allocasuarina campestris X X X X X Allocasuarina corniculata Allocasuarina huegeliana X Allocasuarina humilis Allocasuarina microstachya X X X X X Allocasuarina pinaster X Allocasuarina spinosissima X Allocasuarina thuyoides X Alyxia buxifolia X Amphibromus nervosus X Amphipogon caricinus Amphipogon debilis X X Acacia signata Acacia trigonophylla X X Acacia sedifolia X X X Acacia sclerophylla Acacia sulcata WL X X X X X X X X X X X X X X X X X X X X X 378 Species GR HEA MAL Amphipogon strictus SAL TH X Amyema miquelii X Amyema miraculosa X Andersonia lehmanniana X Andersonia parvifolia X Andersonia sp. X X Angianthus tomentosus X Anigozanthos humilis X Anthotium rubriflorum X Aphelia brizula X Aphelia nutans X X X X Aristida contorta X Astroloma epacridis X X X Astroloma serratifolium X X X Astroloma sp. Dumbleyung X X X X Atriplex bunburyana X Atriplex hymenotheca X Atriplex paludosa X Atriplex semibaccata X X X X X X X Austrostipa drummondii X X Austrostipa hemipogon X X X X X X Austrostipa juncifolia X Austrostipa pycnostachya X X X Austrostipa trichophylla X Austrostipa variabilis X X Baeckea grandibracteata X X Baeckea muricata X Baeckea sp. Baeckea spp. X X Atriplex vesicaria Austrostipa elegantissima X X Atriplex amnicola Austrostipa compressa X X Argyroglottis turbinata Atriplex sp. WL X X X X X X Baeckea sp. Koonadgin X Banksia audax X X X X X X X X X X Banksia cirsioides X Banksia densa X Banksia drummondii X Banksia elderiana X Banksia erythrocephala X Banksia laevigata X X X X X 379 Species GR TH WL X X X X X X Banksia violacea X X X Beaufortia bracteosa X X X Beaufortia micrantha X X X Beaufortia orbifolia X X X Beaufortia puberula X X X Beaufortia schaueri X X X X Bertya sp. X X X X Beyeria brevifolia X X X X X X X Banksia rufa Banksia sphaerocarpa X HEA MAL X Beyeria lechenaultii Billardiera lehmanniana X Blennospora drummondii X X X Blennospora phlegmatocarpa X X Boronia coerulescens X X X Boronia crassifolia X Boronia subsessilis X Boronia ternata X Borya constricta X Borya laciniata X Borya nitida X X Borya sphaerocephala X X Bossiaea concinna X X X X X X X X X X X X Bossiaea walkeri X X Brachyloma geissoloma X Brachyloma mogin X Brachyscome bellidioides X Brachyscome iberidifolia X Brachyscome perpusilla X Bromus arenarius X Bulbine semibarbata X Caesia micrantha X Caladenia denticulata X Caladenia dimidia X Caladenia doutchiae X Caladenia falcata X Caladenia flava X Caladenia hirta X Caladenia hoffmanii X Caladenia longicauda X Caladenia mesocera X Caladenia paradoxa X X Billardiera variifolia Boronia capitata SAL X X X X X X X X 380 Species GR HEA MAL SAL TH Caladenia pendens X Caladenia pulchra X Caladenia reptans X Caladenia roei X Caladenia vulgata X Calandrinia calyptrata X Calandrinia eremaea X Calandrinia granulifera X X X Calandrinia polyandra Calectasia grandiflora X X X Callistemon phoeniceus X X X X Callitris arenaria X Callitris canescens X Callitris columellaris Callitris preissii WL X X X X X X X X Callitris roei X X X X Callitris verrucosa X X X Calothamnus gilesii X X X Calothamnus huegelii X X Calothamnus lateralis X Calothamnus quadrifidus X X X X Calotis hispidula X Calytrix sapphirina X X X X X X X X X X X Calytrix simplex X X Calytrix strigosa X Calytrix violacea X X X Carpobrotus edulis X Carpobrotus modestus X Carpobrotus virescens X Cassytha aurea Cassytha glabella X X Calytrix fraseri Calytrix leschenaultii X X Calytrix breviseta Calytrix desolata X X X X X X Cassytha melantha X Cassytha nodiflora X Cassytha pomiformis X Cassytha racemosa X X X X Caustis dioica X Centrolepis aristata X Centrolepis drummondiana X Centrolepis pilosa X Centrolepis polygyna X X X X 381 Species Chamaescilla corymbosa GR HEA MAL X SAL TH X Chamaescilla spiralis X Chamaexeros fimbriata X Chamelaucium ciliatum X X Chamelaucium megalopetalum X X Chamelaucium pauciflorum X X X X Chamelaucium sp. Merredin Cheilanthes austrotenuifolia X Cheilanthes distans X Cheilanthes sieberi X Cheilanthes tenuifolia X Cheiranthera filifolia X Choretrum glomeratum X X Choretrum pritzelii X X Chorizema aciculare X X X X X X Clematis delicata Clematis spp. X X Chloris truncata Chthonocephalus pseudevax WL X X Coleanthera myrtoides X Comesperma drummondii X X X Comesperma integerrimum X Comesperma scoparium X Comesperma volubile X X X X X Conospermum amoenum X Conospermum bracteosum X Conospermum brownii X Conospermum croniniae X X Conospermum filifolium X X X X Conospermum stoechadis X Conostephium preissii X X X Conostylis aculeata X Conostylis albescens X Conostylis argentea X Conostylis petrophiloides X Coopernookia strophiolata X X X Corynotheca micrantha X Crassula colorata X Crassula decumbens X Crassula exserta X X X Cratystylis conocephala X X X X X Cryptandra apetala X Cryptandra dielsii X 382 Species GR Cryptandra leucopogon HEA MAL X X Cryptandra minutifolia X Cryptandra nutans X X X X X Cryptandra pungens Cryptandra sp. TH X Cryptandra myriantha Cryptandra polyclada SAL WL X X X X X X X X X Cyanicula amplexans X Cyanicula aperta X Cyanicula ashbyae X Cyanostegia angustifolia X Cyanostegia lanceolata X Cyanostegia microphylla X Cyathostemon heterantherus X X Cyphanthera microphylla X Dampiera angulata X X Dampiera glabrescens Dampiera haematotricha X X X Dampiera heteroptera X Dampiera juncea X Dampiera lavandulacea X X X X Dampiera luteiflora X Dampiera oligophylla X Dampiera sacculata X X X Dampiera stenostachya X Dampiera wellsiana X Darwinia diosmoides Darwinia sp. Karonie X X X X Dasymalla axillaris X X Dasymalla terminalis X X Daucus glochidiatus X Daviesia abnormis X X Daviesia audax Daviesia benthamii X X X X X X Daviesia cardiophylla X Daviesia daphnoides X Daviesia hakeoides X X X X X Daviesia incrassata X Daviesia inflata X X X X Daviesia intricata X Daviesia lancifolia X Daviesia nudiflora X Daviesia preissii X X Daviesia rhombifolia X X X X 383 Species GR HEA MAL Daviesia spiralis X X Daviesia teretifolia X Daviesia uncinata X Daviesia uniflora X Desmocladus asper Desmocladus lateriflorus X TH WL X X X X SAL X X X X Desmocladus myriocladus X Desmocladus parthenicus X Dianella revoluta X Dichopogon fimbriatus X Dicrastylis parvifolia X X X X Didymanthus roei X Disphyma crassifolium X Diuris corymbosa X Diuris picta X Diuris setacea X X X Diuris sp. X Dodonaea amblyophylla X Dodonaea bursariifolia Dodonaea caespitosa X X X X Dodonaea inaequifolia X Dodonaea microzyga X Dodonaea ptarmicaefolia X X X Dodonaea stenozyga X X Dodonaea viscosa X Drosera andersoniana X Drosera androsacea X Drosera bulbosa X Drosera glanduligera X Drosera macrantha X X X X X X Drosera macrophylla X X X X Drosera sp. X Drosera stricticaulis X Drosera subhirtella X X Drummondita hassellii X X Ecdeiocolea monostachya X X Einadia nutans X X X X X X Enekbatus stowardii X X Enneapogon caerulescens Eragrostis dielsii X X Enchylaena lanata Enchylaena tomentosa X X Drosera pycnoblasta Elythranthera brunonis X X X 384 Species GR HEA Eremaea beaufortioides X Eremaea pauciflora X MAL SAL X TH WL X X Eremophila brevifolia X Eremophila decipiens Eremophila drummondii X X X X X Eremophila lehmanniana X Eremophila scoparia X X X Erempohila glabra Eriachne ovata X X X Erichsenia uncinata X Ericksonella saccharata X X Ericomyrtus serpyllifolia X X Eriochilus dilatatus X Erodium cygnorum X X X X X Erymophyllum tenellum X X Eucalyptus albida X X Eucalyptus alipes X X X Eucalyptus annulata X Eucalyptus astringens X Eucalyptus burracoppinensis X X Eucalyptus calycogona X X Eucalyptus capillosa X Eucalyptus celastroides X Eucalyptus comitae-vallis X Eucalyptus cylindriflora X X X X X X X Eucalyptus dissimulata X X X Eucalyptus falcata X X X X X X X X X X X Eucalyptus grossa X Eucalyptus incrassata X X Eucalyptus kondininensis X X Eucalyptus leptopoda X X X X X X X Eucalyptus longicornis X X X X X X Eucalyptus moderata Eucalyptus occidentalis X X Eucalyptus horistes Eucalyptus myriadena X X Eucalyptus gracilis Eucalyptus merrickiae X X Eucalyptus gardneri Eucalyptus loxophleba X X Eucalyptus flocktoniae Eucalyptus foecunda X X Eucalyptus densa Eucalyptus eremophila X X X X X 385 Species GR HEA Eucalyptus oleosa MAL SAL X TH WL X Eucalyptus ornata X Eucalyptus ovularis X Eucalyptus phaenophylla X Eucalyptus phenax X Eucalyptus pileata X Eucalyptus platycorys X Eucalyptus redunca X X X X Eucalyptus rigidula X Eucalyptus salicola X X X X X X X Eucalyptus salmonophloia X X X X Eucalyptus salubris X X X X Eucalyptus sargentii X Eucalyptus sheathiana X Eucalyptus spathulata Eucalyptus sporadica X X Eucalyptus subangusta X X X X X X Eucalyptus transcontinentalis X X X X X X X X Eucalyptus wandoo X Eucalyptus yilgarnensis X X Euryomyrtus leptospermoides X Euryomyrtus maidenii X Eutaxia neurocalyx X Exocarpos aphyllus X X X Exocarpos sparteus X X X X X Frankenia desertorum X Frankenia drummondii X Frankenia punctata X Frankenia tetrapetala X Gahnia ancistrophylla X X X Gahnia drummondii X X X Gahnia trifida X X Gastrolobium densifolium X Gastrolobium hookeri X Gastrolobium parviflorum X Gastrolobium punctatum X X X X Gastrolobium crassifolium X X Frankenia cinerea X X X X X X X Gastrolobium reticulatum Gastrolobium spinosum X X X X Gastrolobium trilobum X X Glischrocaryon aureum X Glischrocaryon flavescens X Glischrocaryon roei X X X X X X 386 Species GR HEA MAL Gnephosis brevifolia SAL X Gnephosis tridens X Gompholobium hendersonii X X Gonocarpus confertifolius X X X Goodenia coerulea X Goodenia convexa X Goodenia pinifolia X X Goodenia salina X X X X X Goodenia scapigera X Grevillea acacioides X Grevillea acuaria X X Grevillea anethifolia X X X Grevillea asteriscosa X Grevillea cagiana X X Grevillea didymobotrya X X X Grevillea eriostachya X X X Grevillea eryngioides X X X Grevillea excelsior X Grevillea hookeriana X X Grevillea huegelii X X X X X X X Grevillea integrifolia X X Grevillea involucrata X Grevillea oligantha X X Grevillea paradoxa Grevillea petrophiloides X X X X Grevillea pilosa X X X X X Grevillea shuttleworthiana Grevillea teretifolia X Grevillea umbellulata X X X X X Grevillea wittweri X Grevillea yorkrakinensis X X X X Gunniopsis septifraga X Hakea adnata X X Hakea ambigua X X Hakea baxteri X Hakea corymbosa X X X Hakea cygna X X X Hakea falcata X X X Grevillea prostrata Hakea erecta X X X Grevillea incrassata Grevillea paniculata WL X Gnephosis tenuissima Gonocarpus nodulosus TH X X X X X X X 387 Species GR HEA MAL Hakea francisiana X X Hakea gilbertii X X Hakea horrida X X Hakea incrassata X X Hakea kippistiana SAL X X X Hakea lissocarpha X X Hakea marginata X X X Hakea meisneriana X X X X Hakea minyma X X X X X Hakea obliqua X X Hakea pandanicarpa X X X X Hakea platysperma X X Hakea preissii X Hakea prostrata X X X Hakea scoparia X X X Hakea subsulcata X X X X X X X X X X Hakea trifurcata X X X Halgania lavandulacea X Halgania sp. Halgania spp. X X Helichrysum leucopsideum X Hemigenia dielsii X Hemigenia diplanthera X Hemiphora uncinata X Hibbertia eatoniae Hibbertia exasperata X X X X X X Hibbertia glomerosa X X Hibbertia gracilipes X Hibbertia hemignosta X Hibbertia pungens X X X X Hibbertia recurvifolia X X X Hibbertia stowardii X Hibbertia uncinata X Hibbertia verrucosa X X X Homalocalyx thryptomenoides Hyalochlamys globifera X X Hibbertia lineata Hibbertia rupicola X X Hakea oldfieldii Hakea sulcata X X Hakea newbeyana Hakea petiolaris WL X Hakea lehmanniana Hakea multilineata TH X X X 388 Species GR Hyalosperma demissum X Hyalosperma glutinosum X Hybanthus floribundus X Hydrocotyle alata X Hydrocotyle callicarpa X Hydrocotyle diantha X HEA SAL TH X X X X X X Hypocalymma puniceum X Hypolaena pubescens X Isoetes australis X Isolepis stellata X Isopogon attenuatus X Isopogon buxifolius X Isopogon divergens X X X X X X Isopogon polycephalus X X X Isopogon scabriusculus X X X Isopogon teretifolius X X X Isopogon villosus X X Isotoma petraea X Jacksonia condensata X Jacksonia nematoclada X Jacksonia racemosa X X Juncus sp. X Kennedia prostrata X Kippistia suaedifolia X X Kunzea jucunda X X Kunzea preissiana Kunzea pulchella X X Kennedia prorepens Kunzea micromera WL X Hydrocotyle rugulosa Hypocalymma angustifolium MAL X X X X Labichea stellata X Lasiopetalum microcardium X Lawrencella davenportii X Lawrencella rosea X X Lawrencia squamata X X Laxmannia grandiflora X Laxmannia paleacea X Laxmannia squarrosa X X Lepidium rotundum X Lepidobolus chaetocephalus X X Lepidobolus preissianus X X Lepidosperma angustatum X X X Lepidosperma drummondii X X X X X X X X X X 389 Species GR Lepidosperma effusum Lepidosperma gracile HEA MAL X X X X X X X X X X X X Lepidosperma sanguinolentum X X X Lepidosperma scabrum X Lepidosperma tenue X X X Lepidosperma tuberculatum X X X Leptomeria pauciflora X Leptomeria preissiana X Leptosema daviesioides X Leptospermum erubescens X X Leptospermum incanum X X Leptospermum inelegans X Leptospermum nitens X X X X X X X X X X X Leptospermum oligandrum X Leptospermum spinescens X X Leucopogon blepharolepis X X Leucopogon conostephioides X X Leucopogon crassifolius X X X X Leucopogon cuneifolius X X X X Leucopogon dielsianus X X X X X Leucopogon fimbriatus X Leucopogon hamulosus X X Leucopogon obtusatus X X X Leucopogon sp. Forrestania X Leucopogon tamminensis X X X X Lobelia gibbosa X Lobelia heterophylla X Lobelia tenuior X Lomandra collina X Lomandra effusa X X Lomandra mucronata X Lomandra sp. X Loxocarya cinerea X Loxocarya sp. Lycium australe Lyginia barbata X X Leucopogon sp. Levenhookia dubia X X Lepidosperma rigidulum Leucopogon oxycedrus WL X Lepidosperma pubisquameum Lepidosperma resinosum TH X Lepidosperma leptostachyum Lepidosperma longitudinale SAL X X X X X X X X 390 Species GR Lysinema ciliatum HEA MAL X X SAL X Maireana brevifolia X Maireana carnosa X Maireana erioclada X Maireana oppositifolia X X Malleostemon tuberculatus X Melaleuca acuminata X X X Melaleuca adnata X X X Melaleuca atroviridis X Melaleuca brevifolia X Melaleuca calyptroides X Melaleuca cardiophylla X X X X X Melaleuca carrii X Melaleuca condylosa Melaleuca cordata X X X X X X X X Melaleuca cuticularis X Melaleuca densa X Melaleuca depauperata X Melaleuca eleuterostachya X X X Melaleuca elliptica X X Melaleuca fulgens X X X X X X Melaleuca grieveana X Melaleuca halmaturorum X Melaleuca hamata X X X Melaleuca hamulosa X X Melaleuca haplantha X X Melaleuca lanceolata X X X X X Melaleuca lateriflora X X Melaleuca lateritia X X Melaleuca laxiflora X X Melaleuca lecanantha X Melaleuca leptospermoides X X X X Melaleuca macronychia WL X Maireana amoena Maireana triptera TH X X X X X X X X X Melaleuca marginata X Melaleuca nesophila X X Melaleuca oldfieldii X Melaleuca pauperiflora X X X Melaleuca pentagona X X X Melaleuca platycalyx X X X X X X X Melaleuca pungens Melaleuca radula X Melaleuca scabra X X X X X X X 391 Species GR HEA MAL Melaleuca scalena SAL TH X Melaleuca seriata X Melaleuca spathulata X X X X Melaleuca spicigera X X X X Melaleuca subtrigona X X X X Melaleuca thyoides Melaleuca uncinata X X X X Melaleuca urceolaris X Melaleuca villosisepala X Melaleuca viminea X X X X X X Microcorys ericifolia X Microcorys exserta X Microcorys sp. stellate X Microcorys subcanescens X X Microcybe ambigua X Microcybe multiflora X X X Micromyrtus obovata X X X X X Microtis eremaea X Microtis media X Millotia myosotidifolia X Millotia tenuifolia X X X X X Micromyrtus racemosa X X X X Mirbelia dilatata Mirbelia floribunda X X Mirbelia ramulosa X X Mirbelia spinosa X X Mirbelia trichocalyx X X X X X X Monotaxis grandiflora X X Muehlenbeckia adpressa X Neurachne alopecuroidea X Nicotiana cavicola X Olax benthamiana X X X X X Olearia axillaris X Olearia homolepis X Olearia incondita X Olearia muelleri X Olearia muricata Olearia revoluta X X Mesomelaena stygia Mirbelia spp. X X X Mesomelaena preissii Micromyrtus imbricata X X Melaleuca undulata Menkea spp. WL X X X X X X X 392 Species GR Olearia sp. Eremicola HEA MAL X Opercularia vaginata X Ophioglossum lusitanicum X X Orianthera flaviflora X Orianthera nuda X X Orianthera tortuosa X X Oxalis perennans X X Patersonia juncea X X X Pauridia occidentalis Persoonia coriacea X X X Persoonia hakeiformis X X X X X Persoonia inconspicua Persoonia quinquenervis WL X Opercularia hispidula Pauridia glabella TH X Olearia sp. Kennedy Range Parietaria debilis SAL X X X X X Persoonia saundersiana X X Persoonia striata X Persoonia teretifolia X Persoonia trinervis X X Petrophile circinata X X Petrophile conifera X Petrophile divaricata X X X X X X X X X X X X X Petrophile squamata X X X X Petrophile striata X Phebalium filifolium X X X X Phebalium lepidotum X X X Phebalium megaphyllum X X X X X X Petrophile ericifolia X Petrophile scabriuscula X Petrophile seminuda X Phebalium tuberculosum X Pheladenia deformis X Philotheca gardneri X Philotheca tomentella Phyllangium paradoxum X Phyllanthus calycinus X Phylloglossum drummondii X Physopsis lachnostachya X X X X X X X Pimelea angustifolia X Pimelea argentea X Pimelea graniticola X X X X 393 Species GR Pimelea imbricata HEA MAL X X Pimelea suaveolens X Pimelea sulphurea X Pimelea sylvestris X TH X X X X X Plantago debilis X Platysace commutata X X Platysace deflexa X Platysace effusa X X X X X X X X Platysace juncea X Platysace maxwellii X Platysace trachymenioides X Podolepis capillaris X X X X Podolepis lessonii X X X X Podolepis tepperi X X X Podotheca angustifolia X Podotheca gnaphalioides X Poranthera microphylla X Prasophyllum gracile X Prasophyllum triangulare X Psammomoya choretroides X X X X Pterostylis barbata X Pterostylis pyramidalis X Pterostylis recurva X X Pterostylis sanguinea X X X X X X Pterostylis scabra X Pterostylis sp. X Ptilotus fasciculatus X X Ptilotus gaudichaudii X X Ptilotus holosericeus X Ptilotus manglesii X Ptilotus spathulatus X X X X X X X X X Puccinellia stricta X Pultenaea sp. X Pultenaea verruculosa Regelia inops Rhagodia drummondii X X Pterochaeta paniculata Quinetia urvillei X X Pleurosorus rutifolius Ptilotus polystachyus X X Pityrodia lepidota Platysace compressa WL X X Pittosporum angustifolium Pittosporum phillyreoides SAL X X X X X X X 394 Species GR HEA Rhagodia preissii Rhagodia spp. X Rhodanthe citrina X Rhodanthe laevis Rhodanthe manglesii MAL SAL X X X X X X X X X X X Rhodanthe pygmaea Ricinocarpos glaucus TH X X Rinzia carnosa X X Rinzia crassifolia X X Rinzia polystemonea X Roycea pycnophylloides X Roycea spinescens X Ruppia maritima X Rytidosperma caespitosum X Rytidosperma setaceum X Rytidosperma sp. Goomalling X Santalum acuminatum X Santalum murrayanum Santalum spicatum X X X X X X X Sarcocornia globosa X Scaevola restiacea X X X X Schoenus calcatus X Schoenus globifer X Schoenus hexandrus X X X Schoenus sp. X X X X X Scholtzia sp. X Sclerolaena costata X Sclerolaena diacantha X Sebaea ovata X Senecio glossanthus X Senecio sp. X Senna artemisioides X Senna cardiosperma X Senna charlesiana X Siloxerus multiflorus X Spartochloa scirpoidea X Spergularia marina Spiculaea ciliata X X Schoenus armeria Scholtzia parviflora X X X Schoenus nanus X X Sarcocornia blackiana Scaevola spinescens WL X X X X X X X X X 395 Species GR HEA MAL Spyridium subochreatum X Stachystemon polyandrus X Stackhousia huegelii X X Stackhousia monogyna X X Stackhousia pubescens X Stirlingia simplex X Stylidium breviscapum X Stylidium bulbiferum X Stylidium calcaratum X X Stylidium dichotomum X Stylidium dielsianum X X X X X X X Stylidium involucratum X Stylidium leptophyllum X Stylidium neglectum X X Stylidium nungarinense X Stylidium petiolare X Stylidium piliferum X X X Stylidium repens X X X Stylidium sacculatum X X X X X Stypandra glauca X Stypandra jamesii X X Styphelia tenuiflora X Synaphea interioris X Synaphea petiolaris X X X X X Tecticornia doliiformis X Tecticornia halocnemoides X Tecticornia indica X Tecticornia lepidosperma X Tecticornia leptoclada X Tecticornia loriae X Tecticornia lylei X Tecticornia moniliformis X Tecticornia pergranulata X Tecticornia pterygosperma X Tecticornia undulata X Templetonia sulcata X X Stylobasium australe Templetonia aculeata X X Stylidium yilgarnense Synaphea spinulosa WL X Stenanthemum stipulosum Stylidium squamellosum TH X Stackhousia scoparia Stylidium ecorne SAL X X X X 396 Species Tetrapora preissiana GR HEA MAL X X X Tetrapora tenuiramea SAL X Thelymitra antennifera X Thelymitra flexuosa X TH X Thelymitra latiloba Thelymitra macrophylla X X Thelymitra sargentii X X Thelymitra spiralis X Thomasia spp. X Threlkeldia diffusa Thryptomene australis X X X Thryptomene cuspidata X Thryptomene kochii X Thryptomene racemulosa X X X X X Thysanotus manglesianus X X X X Thysanotus sparteus X X X Thysanotus thyrsoideus X Trachymene cyanopetala X Trachymene ornata X X X Trachymene spp. X X Tricoryne tenella X X Tricostularia compressa X X Triglochin calcitrapa X Triglochin centrocarpa X Triglochin sp. X X Triodia scariosa X Trioidia rigidissima X Tripterococcus brunonis X Trymalium daphnifolium X Urodon dasyphyllus X X X X Verticordia auriculata X Verticordia brownii X Verticordia chrysantha X X Verticordia chrysanthella X X Verticordia densiflora X Verticordia drummondii X Verticordia endlicheriana X Verticordia eriocephala X X Velleia cycnopotamica Verticordia acerosa X X Trachymene sp. Utricularia tenella X X Thryptomene urceolaris Thysanotus patersonii WL X X X X X X X X X X X X X X 397 Species GR Verticordia gracilis HEA MAL X X Verticordia habrantha TH WL X X Verticordia inclusa X Verticordia insignis X Verticordia integra X Verticordia mitodes X Verticordia multiflora SAL X X X X Verticordia pennigera X Verticordia picta X X Verticordia plumosa X X Verticordia roei X X X X Verticordia serrata X X X X Verticordia tumida X X X Vittadinia australasica X Vittadinia cuneata X X X X X X Wahlenbergia preissii X X X X Waitzia acuminata X X X X X X X X X X X Westringia cephalantha Westringia rigida Westringia sp. X X Wilsonia humilis X Wurmbea graniticola X Wurmbea sinora X Xanthorrhoea nana X X X X 398 APPENDIX B DISTRIBUTION OF PLANT FOODS IN THE STUDY AREA The identification of food plants and the seasonal availability of their exploited parts have been based largely on databases and publications, including Atkins and Ruan (2010-2016), Bell and Williams (1997), Bindon (1996), Bird (1985), Bird and Beeck (1988), Brand (2005), Brown et al. (2008), Cribb and Cribb (1975), Daw et al. (1997), Dortch (2004), Florabank (n.d.), Hopper and Lambers (2014), Low (1991), Meagher (1974), Native Grass Resources Group (n.d.), Nussinovitch (2010), Paczkowska and Chapman (2000), Pate and Dixon (1982), Sweedman and Merritt (2006), Webb (2000), Western Australian Herbarium (1998-) and Wildlife Society of Western Australia (2013); only those species present in the study area (Appendix A) are included below. Where species- or genera-specific information was unavailable, periods of availability were determined using the assumptions and generalisations outlined in Chapter 5.3.3. Species are listed alphabetically (within food categories for Appendix B.2–B.7), with taxonomy and naming conventions as per Florabase (Western Australian Herbarium (1998-). Any non-current names encountered in the references above have been updated; taxonomic changes made after 1 April 2017 will not be reflected. Where an individual plant species has been identified as a food source, it has a separate entry. Where a genus was identified, all members have been incorporated within a single entry; the figure in parentheses indicates the number of species present in the study area, excluding those that have a separate entry. Genus-level identifications were treated as per Appendix A and contributed to the totals for the study area (Appendix B.1) only where they were the sole representation of a particular genus, or the named species did not reflect the extent of diversity. Again, spp. entries were assumed to reflect two distinct species, and totals were calculated as such for each Landscape Division (Appendix B.2–B.7). A cross within a month column denotes that the particular food item was available during that month; blank cells indicate that the item was unavailable at that time. For genus entries, the period of availability has been defined as that which applies to the majority of species – in most cases there was minimal variation between members of the same genus. The only exception is Banksia spp. (n=9), as each species flowered 399 at many different times throughout the year, and no dominant period could be identified either within or between species; these flowers have been marked as available yearround. Blue filled cells represent periods of high water-content in storage organs. Plant food categories have been abbreviated as below: o FL = Flowers o FR = Fruit o GM = Gum o L/M = Lerp/manna o L/R/S = Leaves/roots/stems o SE = Seeds o SO = Storage organs 400 B.1 MONTHLY AVAILABILITY OF ALL PLANT FOODS Oct Nov X SE X X X X X Acacia beauverdiana SE X X X X X Acacia microbotrya GM X X X X X X X X X X X X X X X X X SE Acacia saligna Acacia spp. (66) X GM X X X SE X X X GM X X X SE X X Acrotriche sp. FR Allocasuarina spp. (9) SE X X Amyema spp. (2) FR X X Aristida sp. SE Astroloma serratifolium FR X X Astroloma spp. (2) FR X X Atriplex spp. (5) L/R/S X Atriplex vesicaria L/R/S X Aug X Jul X Jun GM May Acacia acuminata Apr Part eaten (category) Mar Species Sep Spring Feb Winter Jan Autumn Dec Summer X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X SE Baeckea spp. (3) FL Banksia sphaerocarpa FL Banksia spp. (9) FL X Beaufortia spp. (5) FL X Billardiera lehmanniana FR X Billardiera variifolia FR Brachyloma spp. (2) FR X Caesia micrantha SO X X X X Caladenia spp. (15) SO X X X X Calandrinia polyandra L/R/S X X X X X X X X X X X X SO X X X X X X X X X X X X FL X X X X X L/R/S X X X X X X X X X X X X SE X X X X X SO X X X X X Callistemon sp. FL X X X X Calothamnus spp. (4) FL X X X X X Calytrix spp. (8) FL X X X X Carpobrotus spp. (2) FR X X X L/R/S X X X X X X X Calandrinia spp. (3) X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X 401 Species Part eaten (category) Mar Apr May Jun Jul Aug Sep Oct Nov Spring Feb Winter Jan Autumn Dec Summer Carpobrotus virescens FR X X X L/R/S X X X X X X X X X X X X Cassytha glabella FR X X X X X X X X X X X X Cassytha racemosa FR X X X X X X X X X X X X Cassytha spp. (4) FR X X X X X X Chamelaucium spp. (4) FL X X X X X X X X X X X X Cyanicula spp. (3) SO X X X X Cyathostemon sp. FL Darwinia spp. (2) FL X Dianella revoluta FR X X X X L/R/S X X X X X X X X X X X X Dichopogon fimbriatus SO X X X X X X X X X X X X Disphyma crassifolium L/R/S X X X X X X X X X X X X Diuris spp. (3) SO X X X X Dodonaea viscosa FR Drosera spp. (9) SO X X X X Einadia nutans L/R/S X X X X Elythranthera sp. SO X X X X Enchylaena tomentosa FR X L/R/S X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X Enekbatus sp. FL Enneapogon sp. SE Eragrostis dielsii SE Eremaea spp. (2) FL X X X X Eriachne sp. SE X X X X Ericomyrtus sp. FL X X X Erodium cygnorum L/R/S X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X SE Eucalyptus loxophleba FL X X X L/R/S X X X Eucalyptus occidentalis GM X X X Eucalyptus oleosa L/M X X X Eucalyptus spp. (45) FL X X X L/M X X X X X X Eucalyptus wandoo FL Exocarpos aphyllus FR X X Exocarpos sparteus FR X X Frankenia spp. (5) FL X X X X X X X X X X X X X X X X L/R/S X X X X X X X X X X X X Gahnia spp. (3) L/R/S X X X X X X X X X X X X Grevillea eriostachya FL X X X X X X X X X X X X 402 GM X X X SE X X X X X X X X X Hypocalymma spp. (2) FL Kennedia prostrata FL Kunzea spp. (4) FL Lepidium sp. L/R/S X X SE X Leptospermum spp. (6) FL X Leucopogon spp. (11) FR X Lomandra spp. (3) FL Lycium australe X X X X X X X X L/R/S X X FR X X Jul Aug Sep Oct Nov X X X X X X X L/R/S X X X Melaleuca spp. (47) FL X X X Micromyrtus spp. (3) FL Microtis spp. (2) SO X X X X Oxalis sp. L/R/S X X X X Persoonia spp. (8) FR X X X Pittosporum phillyreoides GM X X X Platysace deflexa SO X X X Platysace effusa SO X X Platysace maxwellii SO X Platysace spp. (4) SO Prasophyllum spp. (2) X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X Spring Jun FL Winter May X Apr FL Hakea spp. (30) Autumn Mar Grevillea spp. (24) Feb Part eaten (category) Jan Species Dec Summer X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X SO X X X X X Pterostylis recurva SO X X X X X Pterostylis spp. (4) SO X X X X X Regelia sp. FL X Rinzia spp. (3) FL Santalum acuminatum FR Santalum murrayanum FR L/R/S X X X X X X X X X X X X X X X X X X X X X X X X X X Santalum spicatum FR Scaevola spinescens FR Scholtzia sp. FL Spiculaea sp. SO Styphelia tenuiflora FR Tecticornia indica L/R/S X X X X Thelymitra spp. (6) SO X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X 403 Aug Sep Oct Nov X Jul X Jun X X Spring May L/R/S X Winter Apr FR Autumn Mar Threlkeldia diffusa Feb Part eaten (category) Jan Species Dec Summer X X X X X X X X X Thryptomene spp. (5) FL Thysanotus manglesianus L/R/S X X X X X X X X X X X X SO X X X X X X X X X X X X X X X X X X Thysanotus patersonii FL L/R/S X X X X X X X X X X X X X SO X X X X Thysanotus sp. SO X X X X X X X X X X X X Thysanotus thyrsoideus SO X X X X X X X X X X X X Triglochin spp. (2) SO X X X X X X X X X X X X Verticordia spp. (22) FL X X X X Wurmbea spp. (2) SO X Xanthorrhoea sp. FL X X X X X X X X 404 B.2 MONTHLY AVAILABILITY OF PLANT FOODS IN GRANITE Banksia sphaerocarpa FL Banksia sp. FL X Calandrinia spp. (3) FL X Calothamnus sp. FL X Calytrix spp. (2) FL Cyathostemon sp. X X X X X X X X X X X X X X FL X X X Ericomyrtus sp. FL X X X Eucalyptus loxophleba FL X X X X X X Eucalyptus spp. (3) FL X X X Grevillea spp. (5) FL X Hakea spp. (9) FL Hypocalymma sp. FL Kennedia prostrata FL Kunzea spp. (2) FL Leptospermum spp. (2) FL Lomandra spp. (2) FL Melaleuca spp. (10) FL Micromyrtus sp. FL Scholtzia sp. FL Thryptomene sp. FL Thysanotus patersonii FL Verticordia spp. (9) FL Xanthorrhoea sp. FL Astroloma serratifolium FR X X X X X X X Astroloma sp. FR X X X X X X Billardiera variifolia FR X X X Carpobrotus virescens FR X X X Cassytha glabella FR X X X X X X Dianella revoluta FR X X X X Dodonaea viscosa FR Enchylaena tomentosa FR X Exocarpos aphyllus FR X X Exocarpos sparteus FR X X Leucopogon spp. (3) FR X X Lycium australe FR X X Persoonia spp. (2) FR Santalum acuminatum FR Santalum spicatum FR Scaevola spinescens FR X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X Jul X Jun X Apr X Mar X Feb X Jan Nov X Oct FL Spring Sep Baeckea spp. (2) Winter X X Aug Part eaten (category) Autumn May Species Dec Summer X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X 405 Species Part eaten (category) Mar Apr May Jun Jul Aug Sep Oct Nov Spring Feb Winter Jan Autumn Dec Summer Acacia acuminata GM X X X Acacia microbotrya GM X X X Acacia saligna GM X X X Acacia spp. (14) GM X X X Hakea spp. (9) GM X X X Pittosporum phillyreoides GM X X X Atriplex sp. L/R/S X X X X X X X X X X X X Calandrinia spp. (3) L/R/S X X X X X X X X X X X X Carpobrotus virescens L/R/S X X X X X X X X X X X X Dianella revoluta L/R/S X X X X X X X X X X X X Enchylaena tomentosa L/R/S X X X X X X X X X X X X Erodium cygnorum L/R/S X X X X X X X X X X X X Eucalyptus loxophleba L/R/S X X X X X X X X X X X X Gahnia spp. (2) L/R/S X X X X X X X X X X X X Lomandra spp. (2) L/R/S X X X X X X X X X X X X Lycium australe L/R/S X X X X X X X X X X X X Thysanotus patersonii L/R/S X X X X X X X X X Eucalyptus spp. (3) L/M X X Acacia acuminata SE X X X X X Acacia microbotrya SE X X X Acacia saligna SE X X Acacia spp. (14) SE X X Allocasuarina spp. (4) SE X X Calandrinia spp. (3) SE X X Eriachne sp. SE X Erodium cygnorum SE Hakea spp. (9) SE X X X X X X X Caesia micrantha SO X X X X X X Caladenia spp. (11) SO X X X X Calandrinia spp. (3) SO X X X X X X Cyanicula sp. SO X X X X Dichopogon fimbriatus SO X X X X Diuris spp. (3) SO X X X X X Drosera spp. (7) SO X X X X X Elythranthera sp. SO X X X X X Microtis spp. (2) SO X X X X X Platysace maxwellii SO X X X X X X X X X X X X Platysace spp. (2) SO X X X X X X X X X X X X Prasophyllum spp. (2) SO X X X X X Pterostylis recurva SO X X X X X Pterostylis spp. (3) SO X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X 406 X Thelymitra spp. (4) SO X X X X X Thysanotus patersonii SO X X X X X Triglochin spp. (2) SO X X X X Wurmbea spp. (2) SO X X X X X X X X X X X Nov X Oct X Sep X Aug X Jul SO Jun Spiculaea sp. May Part eaten (category) Apr Species Mar Spring Feb Winter Jan Autumn Dec Summer X X 407 B.3 MONTHLY AVAILABILITY OF PLANT FOODS IN HEATH X Callistemon sp. FL X Calothamnus spp. (4) FL X Calytrix spp. (6) FL Chamelaucium spp. (4) FL X Eremaea spp. (2) FL X Ericomyrtus sp. FL Eucalyptus loxophleba FL X X X Eucalyptus spp. (15) FL X X X Grevillea eriostachya FL X X X Grevillea spp. (14) FL X Hakea spp. (23) Nov FL Oct Beaufortia spp. (5) Sep X Spring Aug FL Jul Banksia spp. (9) Winter Jun FL May Banksia sphaerocarpa Apr Baeckea spp. (3) Part eaten (category) FL Species Autumn Mar Feb Jan Dec Summer X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X FL X X X Kunzea spp. (3) FL X X X X Leptospermum spp. (5) FL X X X X Lomandra spp. (2) FL X X X Melaleuca spp. (26) FL X X X Micromyrtus spp. (3) FL X X X Regelia sp. FL X X Rinzia sp. FL X X Scholtzia sp. FL X X X Thryptomene spp. (4) FL X X X X X Thysanotus patersonii FL X X X X X Verticordia spp. (20) FL X X X Xanthorrhoea sp. FL X X X Acrotriche sp. FR X X X X Astroloma serratifolium FR X X X X X X X X X X X X Astroloma sp. FR X X X X X X X X X X X X Brachyloma sp. FR X X X X X X X Cassytha glabella FR X X X X X X X X X X X X Cassytha spp. (4) FR X X X X X X Dianella revoluta FR X X X X Exocarpos aphyllus FR X X X Exocarpos sparteus FR X X Leucopogon spp. (10) FR X X Persoonia spp. (7) FR X X X X X X X X X X X X X X X X X X X X X X X X 408 GM X X X Acacia microbotrya GM X X X Acacia saligna GM X X X Acacia spp. (26) GM X X X Hakea spp. (23) GM X X X Dianella revoluta L/R/S X X X X X Eucalyptus loxophleba L/R/S X X X X Gahnia spp. (2) L/R/S X X X Lomandra spp. (2) L/R/S X X Santalum murrayanum L/R/S X X Thysanotus patersonii L/R/S Eucalyptus spp. (15) L/M X X Acacia acuminata SE X Acacia beauverdiana SE X Acacia microbotrya SE Acacia saligna SE X X Acacia spp. (26) SE X X Allocasuarina spp. (8) SE X X Aristida sp. SE Hakea spp. (23) SE X X Caladenia spp. (2) SO X Drosera spp. (2) SO Platysace deflexa X Nov Acacia acuminata X Oct FR X Sep Styphelia tenuiflora Spring Aug FR Jul Santalum murrayanum Winter Jun Santalum acuminatum Part eaten (category) FR Species May Apr Autumn Mar Feb Jan Dec Summer X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X SO X X X X X X X X X X X X Platysace effusa SO X X X X X X X X X X X X Platysace maxwellii SO X X X X X X X X X X X X Platysace sp. SO X X X X X X X X X X X X Pterostylis recurva SO X X X X X Pterostylis sp. SO X X X X X Thelymitra sp. SO X X X X X Thysanotus patersonii SO X X X X X Thysanotus sp. SO X X X X X X X X X X X X Thysanotus thyrsoideus SO X X X X X X X X X X X X 409 B.4 MONTHLY AVAILABILITY OF PLANT FOODS IN MALLEE Sep Oct Nov Spring Aug Jun Winter May Apr Autumn Mar Feb Jan Dec Summer X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X Part eaten (category) FL Banksia sphaerocarpa FL Banksia spp. (6) FL X Beaufortia spp. (5) FL X Calandrinia sp. FL X Callistemon sp. FL X Calothamnus spp. (3) FL X Calytrix spp. (4) FL Chamelaucium spp. (2) FL Cyathostemon sp. FL Eremaea sp. FL Ericomyrtus sp. FL Eucalyptus loxophleba FL X X X Eucalyptus spp. (31) FL X X X Grevillea eriostachya FL X X X Grevillea spp. (13) FL X Hakea spp. (20) FL Hypocalymma sp. FL Leptospermum spp. (4) FL Lomandra spp. (2) FL Melaleuca spp. (33) FL Micromyrtus spp. (2) FL Regelia sp. FL Rinzia spp. (2) FL Thryptomene spp. (2) FL Thysanotus patersonii FL Verticordia spp. (12) FL Xanthorrhoea sp. FL X Acrotriche sp. FR Astroloma serratifolium FR X X X X X X X Astroloma spp. (2) FR X X X X X X Billardiera variifolia FR X X Cassytha racemosa FR X X X X Cassytha sp. FR X X X X Dianella revoluta FR X X X X Dodonaea viscosa FR Exocarpos aphyllus FR X X Exocarpos sparteus FR X X Leucopogon spp. (8) FR X X Persoonia spp. (4) FR Jul Species Baeckea spp. (2) X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X 410 Sep Oct Nov Spring Aug Jun Winter May Apr Autumn Mar Feb Jan Dec Summer X X X Part eaten (category) FR Santalum murrayanum FR Styphelia tenuiflora FR Acacia acuminata GM X X X Acacia microbotrya GM X X X Acacia spp. (38) GM X X X Hakea spp. (20) GM X X X Atriplex sp. L/R/S X X X X X Calandrinia sp. L/R/S X X X X Dianella revoluta L/R/S X X X Eucalyptus loxophleba L/R/S X X Gahnia spp. (2) L/R/S X X Lomandra spp. (2) L/R/S X Santalum murrayanum L/R/S X Thysanotus patersonii L/R/S Eucalyptus oleosa L/M X X X Eucalyptus spp. (31) L/M X X X Acacia acuminata SE X X Acacia beauverdiana SE X X Acacia microbotrya SE Acacia spp. (38) SE X X Allocasuarina spp. (6) SE X X Calandrinia sp. SE X X Hakea spp. (20) SE X X X X X X X X Calandrinia sp. SO X X X X X X X X Drosera spp. (2) SO X X X X Platysace effusa SO X X X X X X X X X X X X Platysace maxwellii SO X X X X X X X X X X X X Platysace sp. SO X X X X X X X X X X X X Pterostylis sp. SO X X X X X Thysanotus patersonii SO X X X X X Thysanotus sp. SO X X X X Jul Species Santalum acuminatum X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X 411 B.5 MONTHLY AVAILABILITY OF PLANT FOODS IN SALINE X X Eucalyptus spp. (17) FL X X X Frankenia spp. (5) FL X X Grevillea spp. (3) FL X Hakea spp. (2) FL Leptospermum sp. FL Lomandra sp. FL Melaleuca spp. (18) FL Rinzia sp. FL Verticordia sp. FL X Amyema sp. FR X X Billardiera lehmanniana FR X X X Carpobrotus spp. (2) FR X X X Dianella revoluta FR X X X Dodonaea viscosa FR Enchylaena tomentosa FR X Exocarpos aphyllus FR X X Leucopogon sp. FR X X Lycium australe FR X X Santalum acuminatum FR X Santalum spicatum FR Scaevola spinescens FR Threlkeldia diffusa FR Acacia spp. (6) GM Hakea spp. (2) GM Atriplex spp. (5) Nov X Oct X FL Sep FL Eucalyptus loxophleba Aug Darwinia sp. Spring X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X Jul FL Jun Calytrix sp. Winter May X Apr FL Autumn Mar Calandrinia sp. Feb Part eaten (category) FL Jan Species Baeckea sp. Dec Summer X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X L/R/S X X Atriplex vesicaria L/R/S X Calandrinia polyandra L/R/S Calandrinia sp. X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X L/R/S X X X X X X X X X X X X Carpobrotus spp. (2) L/R/S X X X X X X X X X X X X Dianella revoluta L/R/S X X X X X X X X X X X X Disphyma crassifolium L/R/S X X X X X X X X X X X X Einadia nutans L/R/S X X X X X X X X X X X X Enchylaena tomentosa L/R/S X X X X X X X X X X X X Erodium cygnorum L/R/S X X X X X X X X X X X X 412 Species Eucalyptus loxophleba Part eaten (category) L/R/S Mar Apr May Jun Jul Aug Sep Oct Nov Spring Feb Winter Jan Autumn Dec Summer X X X X X X X X X X X X Frankenia spp. (5) L/R/S X X X X X X X X X X X X Gahnia spp. (2) L/R/S X X X X X X X X X X X X Lepidium sp. L/R/S X X X X X X X X X X X X Lomandra sp. L/R/S X X X X X X X X X X X X Lycium australe L/R/S X X X X X X X X X X X X Tecticornia indica L/R/S X X X X X X X X X X X X Threlkeldia diffusa Thysanotus manglesianus Eucalyptus spp. (17) L/R/S X X X X X X X X X X X X L/R/S X X X X X X X X X X X X L/M X X X Acacia spp. (6) SE X X X X X X Allocasuarina spp. (2) SE X X X X X Atriplex vesicaria SE Calandrinia sp. SE X X X X X Eragrostis dielsii SE X X X Erodium cygnorum SE X X X X Hakea spp. (2) SE X X X X X Lepidium sp. SE X X X X X Caladenia spp. (3) SO X X X X Calandrinia polyandra SO X X X X X X X X X X X X Calandrinia sp. SO X X X X X X X X X X X X Drosera spp. (2) SO X X X X Platysace effusa SO X X X X X X X X X X X X Platysace maxwellii Thysanotus manglesianus Triglochin sp. SO X X X X X X X X X X X X SO X X X X X X X X X X X X SO X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X 413 B.6 MONTHLY AVAILABILITY OF PLANT FOODS IN THICKET Sep Oct Nov Spring X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X Part eaten (category) FL Banksia sphaerocarpa FL Banksia spp. (8) FL X Beaufortia spp. (5) FL X Calothamnus spp. (2) FL X Calytrix spp. (4) FL Chamelaucium spp. (3) FL X Darwinia sp. FL X Enekbatus sp. FL Eremaea sp. FL Ericomyrtus sp. FL Eucalyptus loxophleba FL X X X Eucalyptus spp. (15) FL X X X Grevillea eriostachya FL X X X Grevillea spp. (12) FL X Hakea spp. (13) FL Leptospermum spp. (2) FL Lomandra sp. FL Melaleuca spp. (20) FL Micromyrtus spp. (3) FL Scholtzia sp. FL Thryptomene spp. (3) FL Thysanotus patersonii FL Verticordia spp. (11) FL X Astroloma serratifolium FR X Brachyloma spp. (2) FR X Cassytha glabella FR X X X X X X Cassytha spp. (2) FR X X X X X X Dianella revoluta FR X X X X Exocarpos aphyllus FR X X X Leucopogon spp. (4) FR X X Persoonia spp. (6) FR Santalum acuminatum FR X Acacia acuminata GM X X X Acacia spp. (16) GM X X X Hakea spp. (13) GM X X X Atriplex sp. L/R/S X X X X X X X X Dianella revoluta L/R/S X X X X X X X X Eucalyptus loxophleba L/R/S X X X X X X X X Jul Species Baeckea spp. (2) Aug Jun Winter May Apr Autumn Mar Feb Jan Dec Summer X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X 414 Species Gahnia sp. Part eaten (category) L/R/S Mar Apr May Jun Jul Aug Sep Oct Nov Spring Feb Winter Jan Autumn Dec Summer X X X X X X X X X X X X Lomandra sp. L/R/S X X X X X X X X X X X X Oxalis sp. L/R/S X X X X X X X X X X X X Thysanotus patersonii L/R/S X X X X X X X X X Eucalyptus oleosa L/M X X X Eucalyptus spp. (15) L/M X X X Acacia acuminata SE X X X X X Acacia beauverdiana SE X X X X X Acacia spp. (16) SE X X X X X X Allocasuarina spp. (3) SE X X X X X X X X X X X X Hakea spp. (13) SE X X X X X X X X X X X X Drosera spp. (2) SO X X X X Platysace effusa SO X X X X X X X X X X X X Platysace maxwellii SO X X X X X X X X X X X X Pterostylis sp. SO X X X X X Thysanotus patersonii SO X X X X X Thysanotus sp. SO X X X X X X X X X X X X X 415 B.7 MONTHLY AVAILABILITY OF PLANT FOOD IN WOODLAND Sep Oct Nov Spring Aug Winter May Autumn X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X FL X Calandrinia sp. FL X Calothamnus sp. FL X Calytrix spp. (2) FL Cyathostemon sp. FL Darwinia sp. FL X Eremaea sp. FL X Eucalyptus loxophleba FL X X X Eucalyptus spp. (25) FL X X X Eucalyptus wandoo FL X X Grevillea spp. (8) FL Hakea spp. (9) FL Kunzea sp. FL Leptospermum spp. (3) FL Lomandra sp. FL Melaleuca spp. (18) FL X Regelia sp. FL X Rinzia sp. FL X X Thryptomene sp. FL X X Thysanotus patersonii FL X X Verticordia spp. (2) FL Acrotriche sp. FR Amyema sp. FR X X Astroloma serratifolium FR X X X X Astroloma spp. (2) FR X X X Brachyloma sp. FR X Cassytha racemosa FR X X Cassytha sp. FR X Dianella revoluta FR X Dodonaea viscosa FR Enchylaena tomentosa FR X Exocarpos aphyllus FR X X Exocarpos sparteus FR X X Leucopogon spp. (2) FR X X Persoonia spp. (3) FR Santalum acuminatum FR Santalum spicatum FR Jul Beaufortia spp. (2) Jun X Apr FL Mar Banksia spp. (3) Feb Part eaten (category) FL Jan Species Baeckea sp. Dec Summer X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X 416 Jun Jul Aug Sep Oct Nov X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X L/R/S X X X X X X X X X X X X Enchylaena tomentosa L/R/S X X X X X X X X X X X X Erodium cygnorum L/R/S X X X X X X X X X X X X Eucalyptus loxophleba L/R/S X X X X X X X X X X X X Gahnia sp. L/R/S X X X X X X X X X X X X Lomandra sp. L/R/S X X X X X X X X X X X X Thysanotus patersonii L/R/S X X X X X X X X X Eucalyptus spp. (25) L/M X X Acacia acuminata SE X X X X X Acacia microbotrya SE X X X Acacia saligna SE X X X X X Acacia spp. (24) SE X X X X X X Allocasuarina spp. (4) SE X X X X X X Calandrinia sp. SE X X X X X Enneapogon sp. SE Erodium cygnorum SE X X X X Hakea spp. (9) SE X X X X X Caladenia spp. (2) SO X X X X Calandrinia sp. SO X X X X Cyanicula spp. (2) SO X X X X X Diuris sp. SO X X X X X Drosera spp. (2) SO X X X X X Platysace maxwellii SO X X X X Pterostylis spp. (2) SO X X X X X Thelymitra spp. (2) SO X X X X X Thysanotus patersonii SO X X X X Triglochin sp. SO X X X X Part eaten (category) FR Jan X Species Scaevola spinescens Dec May Spring Apr Winter Mar Autumn Feb Summer X X Acacia acuminata GM X X X Acacia microbotrya GM X X X Acacia saligna GM X X X Acacia spp. (24) GM X X X Hakea spp. (9) GM X X X Eucalyptus occidentalis Pittosporum phillyreoides Atriplex spp. (3) GM X X X GM X X X L/R/S X X Calandrinia sp. L/R/S X Dianella revoluta L/R/S Einadia nutans X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X 417 APPENDIX C DISTRIBUTION OF ANIMAL FOODS IN THE STUDY AREA The table below indicates fauna present in the study area – all of which are considered food sources (see Chapter 5.3.2). The list was compiled based on surveys conducted in the study area (Chapman and Dell 1977; Dell 1977; Kitchener and Chapman 1977; McKenzie et al. 1973; Muir 1978, 1979), with habitat and behavioural information from other databases and publications (Atlas of Living Australia, n.d.; Cogger 2014; Cronin 2008, 2014; Nevill 2014; Simpson and Day 2004; Storr 1991; Western Australian Museum n.d.). For ease of reference, species are listed alphabetically within animal type, with taxonomy and naming conventions as per the Australian Faunal Directory (Australian Biological Resources Study 2009); taxonomic changes made after August 2018 will not be reflected. The first series of columns indicates the month/s during which a particular species was present in the study area – a cross indicates that the animal was present in the study area during the majority of the specific month. Yellow filled cells indicate the months during which a particular species or bird or reptile may be laying eggs; grey filled cells indicate the period during which female amphibians may contain eggs. Green filled cells indicate when certain macropods are in peak condition. The second series of columns (to the right of the thick black line) refers to the presence of each species within the six Landscape Divisions. A cross indicates that the species had been observed in a particular area, while an asterisk indicates that it likely inhabited this area, based on generalised habitat information sourced from the references listed above. A blank cell indicates that the animal was not found in this particular area. The seasonal availability of animals and associated products will be identical for each Landscape Division in which the animal can be found. Landscape Divisions have been abbreviated, as below: o GR = Granite o SAL = Saline o HEA = Heath o TH = Thicket o MAL = Mallee o WL = Woodland 418 Jul Aug Sep Oct Nov GR X X X X X X X X X X X X Heleioporus albopunctatus Amphibian X X X X X X X X X X X X X Myobatrachus gouldii Amphibian X X X X X X X X X X X X Pseudophryne guentheri Amphibian X X X X X X X X X X X X X Pseudophryne occidentalis Amphibian X X X X X X X X X X X X X Acanthagenys rufogularis Bird X X X X X X X X X X X X * * * X * * Acanthiza apicalis Bird X X X X X X X X X X X X X X X X X X Acanthiza chrysorrhoa Bird X X X X X X X X X X X X X * X X Acanthiza pusilla Bird X X X X X X X X X X X X * * * Acanthiza uropygialis Bird X X X X X X X X X X X X X X X Aegotheles cristatus Bird X X X X X X X X X X X X Anas gracilis Bird X X X X X X Anas superciliosa Bird X Anthochaera carunculata Bird Anthrochaera chrystoptera Bird X X X Anthus novaeseelandiae Bird X X Aquila audax Bird X X Ardea pacifica Bird Artamus cyanopterus Bird X X Artamus personatus Barnardius zonarius semitorquatus Cacomantis flabelliformis Bird X X Bird X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X * WL Jun X TH May SAL Animal type Amphibian Apr MAL Species Crinia pseudinsignifera Mar HEA Spring Feb Winter Jan Autumn Dec Summer * X * * * * X X X X * * X * * * X * X * * X * X X X * * X X X X X X * * Bird X X X X X X * Cacomantis pallidus Bird X X X X X X * Calamanthus fuliginosus Bird X X X X X X X X X X X X Charadrius ruficapillus Bird X X X X X X X X X X X X X * * * * * * * X * * * X 419 SAL TH WL * * * * X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X X Bird X X X X X X X X X X X X Cygnus atratus Bird X X X X X X X X X X X X Dicaeum hirundinaceum Bird X X X X X X X X X X X X Dromaius novaehollandiae Bird X X X X X X X X X X X X Drymodes brunneopygia Bird X X X X X X X X X X X X Egretta novaehollandiae Bird X X X X X X Eopsaltria australis Bird X X X X X X X X X X X X * * X Epthianura albifrons Bird X X X X X X X X X X X X * X X Eurostopodus argus Bird X X X X X X * X Falco berigora Bird X X X X X X X X X X X X Falco cenchroides Bird X X X X X X X X X X X X * * * X * * Gavicalis virescens Bird X X X X X X X X X X X X * X X X * X Gerygone fusca Bird X X X X X X X X * * Gliciphila melanops Bird X X X X X X X X X Grallina cyanoleuca Bird X X X X X X X X X Himantopus himantopus Bird X X X X X Cinclosoma castanotum Bird X X X X X X Climacteris rufa Colluricincla harmonica rufiventris Coracina novaehollandiae Bird X X X X X Bird X X X X Corvus coronoides Bird X X Cracticus nigrogularis Bird X Cracticus tibicen dorsalis Bird Cracticus torquatus Bird X X X X Apr X Mar Bird Feb Cincloramphus cruralis Jan Animal type Bird Jun X Species Chrysococcyx basalis Dec Nov X Oct * Sep * Aug * Jul MAL Spring HEA Winter GR Autumn May Summer * * X X * X X * * * * X X X * X X * X X X X * X X X X X * X X * * * * X X * * * * * X X * * X X X X X X * X X X * * * * X 420 X X X X X X X X X X Hylacola cauta Bird X X X X X X X X X X X X Lalage suerii Bird X X X X X X X Leipoa ocellata Bird X X X X X X X X X X X X Lichenostomus cratitius Bird X X X X X X X X X X X X Lichmera indistincta Bird X X X X X X X Lophoictinia isura Bird X X X X X Malurus lamberti Bird X X X X X X X X X X X X * * * * Malurus pulcherrimus Bird X X X X X X X X X X X X * X X X * Manorina flavigula Bird X X X X X X X X X X X X * * * * * Melithreptus brevirostris Bird X X X X X X X X X X X X * * X * X Merops ornatus Bird X X X X X Microeca fascinans Bird X X X X X X X X X X X X * Myiagra inquieta Bird X X X X X X X X X X X X * * Neophema elegans Bird X X X X X X X X X X X X * * X X Nesoptilotis leucotis Bird X X X X X X X X X X X X * * X Ninox novaeseelandiae Bird X X X X X X X X X X X X * Nymphicus hollandicus Bird X X X X X X X X X X X X * Oreoica gutturalis Bird X X X X X X X X X X X X * X X Pachycephala pectoralis Bird X X X X X X X X X X X X * * X Pachycephala rufiventris Bird X X X X X X X X X X X X * * Pardalotus striatus Bird X X X X X X X X X X X X Parvipsitta porphyrocephala Bird X X X X X X X X X X X X * Petrochelidon nigricans Bird X X X X X X X * X Petroica goodenovii Bird X X X X X X X * X X X X X X * X * X * * X X X X WL X TH Nov SAL Oct MAL Sep X GR Aug X Jul * Jun X May * Animal type Bird Apr * Species Hirundo neoxena Mar HEA Spring Feb Winter Jan Autumn Dec Summer * * * * * * * * X X X * * * * * * X * * * X X X X X X * * X * X X * X X * * * 421 Species Phaps chalcoptera Animal type Bird Mar Apr May Jun Jul Aug Sep Oct Nov GR HEA MAL SAL TH WL Spring Feb Winter Jan Autumn Dec Summer X X X X X X X X X X X X * * * X X * Phylidonyris niger Bird X X X X X X X X X X X X * * * Platycercus icterotis Bird X X X X X X X X X X X X X X X Platycercus varius Bird X X X X X X X X X X X X * * * X * * Platycercus zonarius Bird X X X X X X X X X X X X * X X X X X Podargus strigoides Bird X X X X X X X X X X X X * Polytelis anthopeplus Bird X X X X X X X X X X X X * * * X * * Pomatostomus superciliosus Bird X X X X X X X X X X X X * * * X X * Ptilotula ornata Bird X X X X X X X X X X X X * * * * * Pyrrholaemus brunneus Bird X X X X X X X X X X X X X * X Rhipidura leucophrys Bird X X X X X X X X X X X X X Sericornis frontalis Bird X X X X X X X X X X X X * Smicrornis brevirostris Bird X X X X X X X X X X X X * Stipiturus malachurus Bird X X X X X X X X X X X X Strepera versicolor Bird X X X X X X X X X X X X Tadorna tadornoides Bird X X X X Turnix varius Bird X X X X X X X X X X X X Vanellus tricolor Bird X X X X X X X X X X X X Zosterops lateralis Bird X X X X X X X X X X X X Antechinomys laniger spenceri Mammal X X X X X X X X X X X Cercartetus concinnus Mammal X X X X X X X X X X Chalinolobus gouldii Mammal X X X X X X X X X X Dasyurus geoffroii Mammal X X X X X X X X X Isoodon obesulus Mammal X X X X X X X X Macropus fuliginosus Mammal X X X X X X X X * X * * X X * X X X X X X X X X * X X X * X X X * * * * X * * * * * * X X * * X X * X X X X * X X X X X * X X X X * * * X X X X X X X X * * X X 422 Jul Aug Sep Oct Nov X X X X X X X X X X X Notamacropus irma Mammal X X X X X X X X X X X X X X X X X Notomys mitchellii Mammal X X X X X X X X X X X X X * X X X Nyctophilus geoffroyi Mammal X X X X X X X X X X X X X Osphranter robustus Mammal X X X X X X X X X X X X * Phascogale calura Mammal X X X X X X X X X X X X X Pseudomys occidentalis Mammal X X X X X X X X X X X X X Sminthopsis crassicaudata Mammal X X X X X X X X X X X X Sminthopsis granulipes Mammal X X X X X X X X X X X X * X Sminthopsis murina Mammal X X X X X X X X X X X X X * Tachyglossus aculeatus Mammal X X X X X X X X X X X X * * Tarsipes rostratus Mammal X X X X X X X X X X X X Trichosurus vulpecula Mammal X X X X X X X X X X X X X Vespadelus pumilus Mammal X X X X X X X X X X X X X Anilios australis Reptile X X X X X X X X X X X X * Crenadactylus ocellatus Cryptoblepharus plagiocephalus Ctenophorus cristatus Reptile X X X X X X X X X X X X * Reptile X X X X X X X X X X X X X Reptile X X X X X X X X X X X X * Ctenophorus maculatus Reptile X X X X X X X X X X X X Ctenophorus ornatus Reptile X X X X X X X X X X X X X Ctenophorus salinarum Reptile X X X X X X X X X X X X X Ctenotus impar Reptile X X X X X X X X X X X Ctenotus pantherinus Reptile X X X X X X X X X X Ctenotus schomburgkii Reptile X X X X X X X X X Delma australis Reptile X X X X X X X X X WL Jun X TH May SAL Apr MAL Animal type Mammal GR Species Myrmecobius fasciatus Mar HEA Spring Feb Winter Jan Autumn Dec Summer * * * * X X X X X X X * * * X * X X * * * * X * X * X * X X X * X * X X * X * X X X * X X X X * * * X * * 423 May Jun Jul Aug Sep Oct Nov GR HEA X X X X X X X X X X X * * Demansia sp. Reptile X X X X X X X X X X X X Egernia kingii Reptile X X X X X X X X X X X X * Gehyra variegata Reptile X X X X X X X X X X X X X Hesperoedura reticulata Reptile X X X X X X X X X X X X Lerista distinguenda Reptile X X X X X X X X X X X X * * Lialis burtonis Reptile X X X X X X X X X X X X * * * Liopholis multiscutata Reptile X X X X X X X X X X X X * X X Lucasium maini Reptile X X X X X X X X X X X X * X Morethia obscura Reptile X X X X X X X X X X X X X * * Parasuta gouldii Reptile X X X X X X X X X X X X * * * * * Pogona minor Reptile X X X X X X X X X X X X X * X X Strophurus spinigerus Reptile X X X X X X X X X X X X * X * Tiliqua occipitalis Reptile X X X X X X X X X X X X * * X Tiliqua rugosa Reptile X X X X X X X X X X X X * * X Underwoodisaurus milii Reptile X X X X X X X X X X X X * * * Varanus gouldii Reptile X X X X X X X X X X X X * * * WL Apr X TH Animal type Reptile SAL Species Delma fraseri Mar MAL Spring Feb Winter Jan Autumn Dec Summer * X X X * * * * * * * * X X * X * X * * X 424 APPENDIX D MEASURING ARTEFACT ATTRIBUTES Artefactual material was analysed by XU within each test-pit. Relevant attributes were recorded in a database of Microsoft Excel spreadsheets. Due to the issues associated with often poor-quality quartz, some attributes could not be observed on particular artefacts. Selected attributes are discussed in more detail below, where earlier methodological information (see Chapter 9.3.3) was insufficient. D.1 METRIC ATTRIBUTES Artefacts were weighed on digital scales that measured weight in 0.05 g increments. All artefacts were weighed individually except for flaked fragments and flakes measuring less than 10 mm in percussion/maximum length – these were weighed as a group containing other artefacts of the same raw material and type, from the same pit and XU. Length, width and thickness measurements were taken using digital calipers that measured in 0.01 mm increments. Maximum dimensions were recorded for flaked fragments and cores as the former cannot be oriented, and there is no orientation system that applies to all varieties of core (Holdaway and Stern 2004:186). Oriented dimensions were recorded for flakes and retouched flakes where the flaking axis could be identified (i.e. platform to termination). Where this was not possible (often due to breakage or extensive retouch) maximum dimensions were recorded, but in a slightly different manner than for cores and flaked fragments due to the presence of distinct dorsal and ventral surfaces. These dimensions were measured as below: Maximum dimensions: o Maximum length: a line drawn between the two points separated by the greatest distance on the same surface (flakes, retouched flakes) or any surface (cores, flaked fragments); o Maximum width: the longest line that can be drawn at 90 degrees to maximum length; 425 o Maximum thickness: the longest line that can be drawn at 90 degrees to maximum width (cores, flaked fragments) or the maximum distance between the dorsal and ventral surfaces (flakes, retouched flakes). Oriented dimensions: o Percussion length: a line measured from the platform (bisecting bulb of percussion/impact point, where present) to termination, following the flaking axis (i.e. the direction that the fracture has progressed from impact until exiting the core); o Width at midpoint: measured at the midpoint of percussion length, at right angles; o Thickness at midpoint: distance between the dorsal and ventral surfaces, measured at the point where percussion length intersects with width at midpoint. D.2 RETOUCH As discussed earlier (see Chapter 9.3.2), retouch was identified where additional flake removals occurred after the ventral flake surface was created – negative flake scars should therefore occur on the ventral surface itself or be initiated from this newly created surface (Hiscock 2007). In the latter case, at least some complete scars should remain (i.e. the latest in the sequence, if scar superimposition is evident); micro-scars resulting from edge damage are not considered retouch. If only partial negative scars are preserved on flake margins, these instead represent flake removals that occurred when the artefact was still attached to the parent core – scars that are subsequently split when the new flake is created – not retouch (Hiscock 2007; Holdaway and Stern 2004:161). While it should therefore be relatively easy to separate retouch from edge damage and previous flake removals, in practice, poor-quality raw material hampered some efforts; where retouch was ambiguous, pieces were classified as flakes. Two types of retouch were encountered, defined below: o Backing: steep uni- or bidirectional retouch that unites the dorsal and ventral surfaces by creating a blunt margin located opposite the working edge – retouch can be created via the bipolar technique, so crushing may be evident (Hiscock 1993; Holdaway and Stern 2004:159). 426 o Unifacial retouch: retouch occurring on only one surface of the flake, either dorsal or ventral, modifying the inferred working edge (Holdaway and Stern 2004:158–159). D.3 BREAKAGE The presence/absence of breakages were noted for flakes and retouched flakes. The type of break was quantified, where possible, but in many cases poor quality raw material prevented this. Breakage types and preserved portions were classified following Hiscock (2002b), as illustrated in Figure D.1. Figure D.1 Breakage types and the resultant flake fragments produced (Hiscock 2002b:253). 427 D.4 FLAKE SCAR ORIENTATION Negative flake scar orientation was recorded following Holdaway and Stern (2004:146–147), via the use of a quadrat system – this method could only be applied to artefacts that could be oriented along the flaking axis, where flake scar directionality was not obscured by poor-quality raw material. An 'x' is superimposed over the artefact, with the central point positioned in the midpoint of the dorsal surface, and the end of the upper arms joining each edge of the platform; where the platform was missing, the piece was oriented following the flaking axis and the arms positioned at the 'corners' of the flake, i.e. where the left and right lateral margins intersected with the break. Quadrats are numbered in a clockwise direction, beginning with the quadrat that contained/would have contained the platform (Figure D.2). Figure D.2 Method of recording flake scar orientation on flakes and retouched flakes (Holdaway and Stern 2004:147). Each quadrat was evaluated to determine whether flakes had been initiated therein, based on negative bulbs of percussion, ripple mark orientation, and/or general scar topography. It was important to specify the precise quadrats containing initiations – rather than simply counting how many quadrats contain them – as the number of core rotations represented will depend on whether the platform quadrat (Q1) preserved evidence of prior flake removals. For example, a flake exhibiting flake removals from just one quadrat may indicate no core rotation (when Q1 contains the initiations, thus 428 previous flakes were detached from the same platform as the current flake) or a single rotation (Q2–Q4, when the quadrat exhibiting initiations does not contain the platform). Therefore, the presence/absence of flake initiations was recorded for quadrats one to four on each suitable flake. D.5 BIPOLAR REDUCTION Bipolar flaking involves three separate components: a core, a hammerstone and an anvil, but the arrangement of these elements can vary. The bipolar technique is often separated into axial and non-axial reduction, depending on the spatial relationship between the impact point (where the hammerstone strikes the core) and the area where the core contacts the anvil. In axial reduction, the blow is directed towards the anvil, compressing the core between the hammerstone and anvil, and detaching a flake (Diez-Martin et al. 2011; Hiscock 2015). In non-axial or oblique reduction, core placement and/or direction of blow means that force exits the core without reaching the anvil – the only thing that differentiates this technique from freehand knapping is the use of an anvil to stabilise the core. These major methodological differences affect the form of flaked products, and indeed the lack of compressive forces in non-axial techniques makes those products difficult to differentiate from those created via freehand flaking (Diez-Martin et al. 2011; Hiscock 2015). Therefore, only axial bipolar is considered here; Hiscock (2015) considers this a 'strong' definition for bipolar flaking. There is no single feature that signifies the use of the axial bipolar technique. Instead, raw material type and quality dictate bipolar and freehand signatures, so diagnostic features must be appropriate to a particular raw material type (De La Pena 2015; Jeske and Lurie 1993, Pargeter and De La Pena 2017). For quartz, the features that most accurately separate the products of bipolar from freehand knapping are crushing and termination type (Figure D.3): bipolar terminations are generally axial or splintered and flakes often preserve platform segments at the distal end, along with extensive crushing at the platform and distal margin; freehand flakes and cores rarely exhibit crushing, instead preserving feather, step, hinge or overshot (aka plunging) terminations (Diez-Martin et al. 2011; Pargeter and De La Pena 2017). The bipolar technique was identified by the presence of one or more of the diagnostic features noted above. Those lacking any such features (i.e. exhibiting the freehand counterpart) 429 were considered to be the product of freehand reduction. As with other attributes, the presence/absence of bipolar reduction could not always be observed on poor-quality raw material. A B Figure D.3 Flake terminations. A: side views of different terminations showing the path the fracture takes through the core (Cotterell and Kamminga 1987:684). B: dorsal surface of a flake with a splintered termination – distal end is at the bottom of the photograph (Pargeter and De La Pena 2017:10). 430 APPENDIX E ANDERSON ROCKS STONE ARTEFACT ASSEMBLAGE This appendix contains the raw artefact data for the Anderson Rocks stone artefact assemblages. Weight (grams) and dimensions (millimetres) have been rounded to two decimal places. Blank cells denote unusable or missing data. Unusable data was that deemed not representative of a particular attribute, i.e. mm/scar for a flake that could not be oriented. Where data were absent, this indicated that a particular attribute could not be confidently measured on a certain specimen, often due to poor-quality raw material. Artefact types have been ordered by pit/scatter (beginning with surface material), XU depth and raw material. See Tables 9.1–9.2 and Appendix D for details on attributes recorded for each artefact type. Abbreviations are as follows: o Bck. = backed o Ret. = retouch type o BIF = Banded Iron Formation o RM = raw material o Bk = breakage o Sc. = scatter o Chal. = chalcedony o Sc1 = Scatter 1 o Comb wt. = combined weight o Sc2 = Scatter 2 o Dims = dimensions o Sc3 = Scatter 3 o ED = edge damage o Sil. = silcrete o HQ = high-quality raw material o T = thickness o L = length o Trans. = transitional o Lng. = longitudinal o Trv. = transverse o Max. = maximum dimensions o Uni = unifacial o Mrg. = marginal o %Unmod = proportion of o #NFS = no. of negative flake scars o Ort. = oriented dimensions o W = width o OSM = other surface material o Wt = weight o #Plat = no. of platforms o XU = excavation unit o Qtz = quartz unmodified surface 431 Pit/Sc. XU RM HQ Wt (g) L (mm) W (mm) T (mm) #NFS g/scar #Plat %Unmod Bipolar E.1 CORES Sc1 - Qtz N 4.9 23.62 16.39 15.13 5 0.98 2 < 25 Y Sc1 - Qtz N 5.3 24.43 20.82 15.02 8 0.66 4 < 25 Y - trans Sc2 - Qtz N 2.65 20 15.22 8.99 7 0.38 3 < 25 N Sc3 OSM - Qtz Qtz N N 5.3 46.65 25.59 41.9 21.53 35.04 10.51 32.39 6 9 0.88 5.18 4 4 < 25 < 25 N Y - trans AR2-3 6 Qtz N 6.2 25.05 19.72 10.86 4 1.55 2 AR2-3 8 Qtz Y 2.45 22.67 12.97 7.47 4 0.61 4 < 25 Y AR2-3 AR2-3 9 13 Qtz Qtz Y N 1.85 3.4 15.7 22.62 13.06 14.94 9.57 10.54 8 8 0.23 0.43 2 3 < 25 < 25 Y AR2-3 15 Qtz Y 1.15 14.56 13.2 6.7 4 0.29 3 25-50 Y AR2-3 16 Qtz N 1.45 15.69 11.08 9.56 9 0.16 4 < 25 Y AR2-3 16 Qtz N 1.6 17.9 14.11 6.55 3 0.53 3 25-50 Y AR2-3 17 Qtz N 5.5 22.95 21.65 12.91 5 1.10 3 AR2-3 17 Qtz N 10.65 31.53 25.74 14.21 5 2.13 3 25-50 N AR2-3 18 Qtz N 5.3 22.93 22.55 12.51 4 1.33 2 25-50 Y AR2-3 AR2-3 18 19 Qtz Qtz N Y 6.25 2.25 26.93 22.68 19.69 17.85 10.35 7.64 7 4 0.89 0.56 4 4 < 25 < 25 Y Y AR2-3 20 Qtz N 3.35 23.55 18.77 8.69 8 0.42 4 < 25 Y AR2-3 21 Qtz Y 6.15 30.33 17.28 10.78 7 0.88 4 < 25 N 432 E.2 SMALL FLAKED FRAGMENTS XU RM No. Comb. Wt (g) No. HQ Sc1 - Qtz 1 0.5 0 Sc3 - Qtz 2 0.35 2 AR1 1 Qtz 3 0.25 0 AR1 2 Qtz 3 0.15 0 AR1 3 Qtz 3 0.2 0 AR1 4 Qtz 4 0.45 3 AR1 5 Qtz 7 0.7 0 AR1 6 Qtz 3 0.35 0 AR1 7 Qtz 7 0.3 0 AR1 8 BIF 1 0.05 1 AR1 8 Qtz 8 0.45 0 AR1 AR1 9 10 Qtz Qtz 8 2 0.8 0.15 0 0 AR1 11 Qtz 4 0.7 0 AR1 12 Qtz 1 0.05 0 AR1 AR2-3 13 1 Qtz Qtz 1 1 0.2 < 0.05 0 1 AR2-3 2 Qtz 1 0.1 1 AR2-3 3 Qtz 1 0.05 0 AR2-3 AR2-3 5 6 Qtz Qtz 1 7 < 0.05 1.45 1 2 AR2-3 7 Qtz 1 0.15 0 AR2-3 9 Qtz 1 0.1 0 AR2-3 AR2-3 10 11 Qtz Qtz 3 1 0.1 0.05 3 0 AR2-3 12 Qtz 2 0.3 0 AR2-3 13 Qtz 3 0.35 2 AR2-3 AR2-3 15 16 Qtz Qtz 5 3 0.3 0.15 0 1 AR2-3 17 Qtz 2 0.05 2 AR2-3 18 Qtz 25 2 9 AR2-3 AR2-3 19 20 Qtz Qtz 5 2 0.3 0.2 1 0 AR2-3 21 Qtz 3 0.2 3 AR2-3 22 Qtz 2 0.35 0 AR2-3 23 Qtz 3 0.2 3 AR2-3 25 Qtz 1 0.25 0 AR2-3 26 Qtz 1 < 0.05 0 AR2-3 28 Qtz 1 < 0.05 0 AR2-3 AR2-3 29 31 Qtz Qtz 1 1 < 0.05 0.05 1 1 Pit/Sc. 433 Pit/Sc. XU RM HQ Wt (g) L (mm) W (mm) T (mm) #NFS g/scar #Plat E.3 LARGE FLAKED FRAGMENTS Sc1 - Qtz N 4.6 21.56 17.05 11.04 2 2.30 2 Sc2 Sc2 - Qtz Qtz N N 0.7 1.75 12.9 20.57 9.5 12.99 6.2 6.8 1 2 0.70 0.88 1 1 Sc2 - Qtz N 2.35 17.44 12.7 9.3 3 0.78 3 Sc2 - Qtz N 2.55 22.71 12.91 8.48 2 1.28 1 Sc3 Sc3 - Qtz Qtz Y Y 0.35 0.35 11.85 11.26 7.34 6.93 4.54 5.04 2 3 0.18 0.12 2 2 OSM - Qtz N 6.25 23.57 22.05 11.64 7 0.89 2 OSM - Qtz N 6.75 30.68 20.27 10.56 AR1 AR1 3 8 Qtz Qtz N N 10.45 0.4 37.51 11.62 22.96 7.16 10.81 4.49 1 10.45 1 AR1 9 Qtz N 0.15 10.16 4.72 3.77 AR1 9 Qtz N 0.2 10.19 5.18 4.16 2 0.10 2 AR1 9 Qtz N 0.45 11.8 10.86 3.73 AR1 9 Qtz N 2.15 23.02 11.22 7.43 3 0.72 2 AR1 10 Qtz N 0.4 11.15 7.18 5.16 3 0.13 1 AR1 13 Qtz N 0.4 12.58 6.07 4.32 2 0.20 1 AR1 13 Qtz N 0.6 11.46 5.7 5.19 AR2-3 2 Qtz N 0.2 12.71 4.41 2.85 3 0.07 1 AR2-3 5 Qtz N 0.4 11.99 6.45 4.8 3 0.13 2 AR2-3 6 Qtz N 6.3 27.66 20.23 12.1 4 1.58 2 AR2-3 AR2-3 7 11 Qtz Qtz Y N 0.5 0.25 14.09 12.56 6.09 5.54 4.95 4.81 1 2 0.50 0.13 1 2 AR2-3 14 Qtz N 0.65 14.7 8.53 5.53 4 0.16 2 AR2-3 14 Qtz N 1.1 15.03 12.54 6.5 2 0.55 2 AR2-3 AR2-3 17 18 Qtz Qtz Y N 0.2 0.2 15.01 11.11 4.43 5.46 3.52 3.77 3 1 0.07 0.20 2 1 AR2-3 18 Qtz N 0.35 13.01 7.79 4.83 3 0.12 2 AR2-3 18 Qtz N 0.85 13.44 10.15 6.8 4 0.21 3 AR2-3 AR2-3 19 19 Qtz Qtz N N 0.15 0.2 10.38 10.3 6.95 4.65 2.97 3.61 2 3 0.08 0.07 2 2 AR2-3 19 Qtz N 0.6 12.14 10.16 4.97 2 0.30 2 AR2-3 19 Qtz N 0.65 11.84 9.68 5.18 3 0.22 2 AR2-3 AR2-3 19 19 Qtz Qtz N N 3.15 5.15 19.56 26.73 14.15 18.44 9.98 10.33 3 2 1.05 2.58 2 2 AR2-3 20 Qtz N 2.25 22.9 15.56 5.79 4 0.56 3 AR2-3 23 Qtz N 0.25 12.53 6.3 3.83 2 0.13 2 AR2-3 23 Qtz N 0.5 10.01 6.63 8.24 2 0.25 2 AR2-3 25 Qtz N 2.15 27.43 11.92 7.17 434 E.4 SMALL FLAKES XU - RM Qtz No. 4 Comb. Wt (g) 0.55 No. HQ 1 No. broken 0 AR1 2 Qtz 1 < 0.05 0 0 AR1 3 Qtz 4 0.15 0 3 AR1 AR1 4 5 Qtz Qtz 1 2 0.05 0.35 1 1 1 1 AR1 6 Sil. 1 < 0.05 0 0 AR1 8 Qtz 2 0.2 1 1 AR1 AR1 9 9 Qtz Sil. 4 1 0.45 0.05 1 0 1 0 AR1 10 Qtz 1 < 0.05 1 1 AR1 11 Qtz 1 0.05 1 1 AR1 AR2-3 12 2 Qtz Qtz 3 2 0.1 0.2 3 2 1 1 AR2-3 3 BIF 1 < 0.05 1 0 AR2-3 3 Qtz 6 0.35 2 1 AR2-3 AR2-3 4 5 Qtz Qtz 3 9 0.05 0.2 3 6 3 5 AR2-3 6 BIF 1 0.1 1 0 AR2-3 6 Qtz 7 0.5 0 1 AR2-3 AR2-3 7 8 Qtz Qtz 10 8 0.5 0.35 1 4 5 6 AR2-3 9 Qtz 10 0.55 10 8 AR2-3 10 Qtz 8 0.35 8 7 AR2-3 AR2-3 11 12 Qtz Qtz 7 14 0.15 0.7 7 7 6 9 AR2-3 13 Chal. 2 0.05 2 0 AR2-3 13 Qtz 10 0.75 8 8 AR2-3 AR2-3 14 15 Qtz BIF 20 2 1.8 < 0.05 10 2 11 0 AR2-3 15 Qtz 30 1.5 18 14 AR2-3 16 Qtz 17 0.85 9 5 AR2-3 AR2-3 17 17 BIF Qtz 1 31 < 0.05 1.35 1 25 0 12 AR2-3 18 BIF 1 0.35 1 0 AR2-3 18 Qtz 71 4.7 55 41 AR2-3 AR2-3 19 20 Qtz Chal. 36 1 2.55 0.1 33 1 25 1 AR2-3 20 Qtz 29 2 19 18 AR2-3 21 Qtz 26 1.8 11 19 AR2-3 22 Qtz 21 0.95 17 12 AR2-3 22 Sil. 1 < 0.05 0 0 Pit/Sc. Sc3 435 Pit/Sc. AR2-3 XU 23 RM BIF No. 1 Comb. Wt (g) < 0.05 No. HQ 1 No. broken 0 AR2-3 23 Qtz 10 1 0 3 AR2-3 24 Qtz 1 < 0.05 0 0 AR2-3 AR2-3 25 26 Qtz BIF 5 1 0.7 < 0.05 0 1 2 0 AR2-3 26 Qtz 1 < 0.05 0 0 AR2-3 27 Qtz 2 0.75 2 0 AR2-3 AR2-3 28 28 BIF Qtz 1 1 0.05 0.15 1 0 0 0 AR2-3 29 Qtz 1 < 0.05 1 1 AR2-3 30 Qtz 3 0.1 3 2 AR2-3 AR2-3 31 32 Qtz Qtz 1 2 < 0.05 0.1 1 2 0 2 436 HQ Wt (g) Dims. L (mm) W (mm) T (mm) Bk #NFS mm/scar #Plat %Unmod Bipolar ED - Qtz. N 0.4 Ort. 13.02 5.96 3.52 N 3 4.34 1 < 25 N N Sc1 Sc1 - Qtz. Qtz. N N 0.55 1.5 Ort. Ort. 10.42 10.15 8.36 13.4 4.83 6.36 N N 3 4 3.47 2.54 2 2 < 25 < 25 Y Y N N Sc2 - Qtz. N 0.8 Ort. 14.54 12.66 3.55 N 2 7.27 N N Sc2 - Qtz. N 1.25 Ort. 14.83 11.35 6.2 N 1 14.83 1 Y N Sc2 Sc2 - Qtz. Qtz. N N 1.65 2.3 Ort. Ort. 15.11 30.3 11.56 11.78 5.47 5.14 N 4 2 3.78 15.15 2 3 < 25 Y Y N N Sc3 - Chal. Y 0.25 Ort. 10.17 10.48 1.83 Y 1 10.17 < 25 N N Sc3 - Qtz. Y 0.15 Max. 12.38 6.06 3.98 Y 4 < 25 N Sc3 Sc3 - Qtz. Qtz. Y Y 0.15 0.15 Max. Ort. 12.14 10.02 7.97 7.86 1.58 1.42 Y N 1 1 10.02 1 < 25 < 25 N N Y Sc3 - Qtz. N 0.25 Ort. 11.9 10.97 2.21 N 4 2.98 3 < 25 N Y Sc3 - Qtz. Y 0.25 Ort. 12.38 7.7 2.25 N 3 4.13 2 < 25 Y Y Sc3 Sc3 - Qtz. Qtz. N Y 0.25 0.35 Ort. Ort. 12.73 13.44 5.6 5.98 2.91 3.34 N N 2 3 6.37 4.48 1 3 < 25 < 25 N N N N Sc3 - Qtz. Y 0.4 Ort. 11.92 10.66 2.48 N 4 2.98 2 < 25 N N Sc3 - Qtz. N 0.5 Ort. 13.39 8.59 2.54 N 2 6.70 1 < 25 Y N Sc3 Sc3 - Qtz. Qtz. N Y 0.5 0.65 Max. Ort. 19.53 10.82 12.2 12.21 1.99 4.13 Y Y 1 3 3.61 2 < 25 < 25 N N N Sc3 - Qtz. Y 1.1 Ort. 22.36 8.71 4.34 N 3 7.45 3 < 25 Y N Sc3 - Qtz. N 2.4 Ort. 20.02 18.08 5.35 N 4 5.01 2 < 25 Y N OSM AR1 1 Qtz. Qtz. N N 0.4 0.25 Ort. Ort. 10.54 10.87 11.64 11.3 2.32 2.33 N N 5 3 2.11 3.62 2 2 < 25 < 25 N N N Bk type RM Sc1 Pit/Sc. XU E.5 LARGE FLAKES Trv. Trv. 437 HQ Wt (g) Dims. L (mm) W (mm) T (mm) Bk #NFS mm/scar #Plat %Unmod Bipolar ED Y 0.55 Ort. 12.8 7.68 3.89 N 2 6.40 2 < 25 Y N AR1 4 Qtz. N 0.5 Ort. 11.98 7.72 3.88 N 2 5.99 2 < 25 N N AR1 AR1 4 4 Qtz. Sil. N N 4.7 0.2 Ort. Ort. 26.57 16 18.6 8.98 6.36 1.09 N 1 2 26.57 8.00 1 1 < 25 < 25 Y N N N AR1 8 Qtz. Y 0.3 Ort. 12.44 8.04 1.88 N 2 6.22 1 < 25 N Y AR1 9 Qtz. N 0.15 Ort. 9.86 6.2 2.21 Y 3 3.29 2 < 25 N N AR1 AR2-3 12 3 Qtz. Qtz. Y N 0.2 1.55 Max. Ort. 11.43 20.11 6.83 15.9 2.85 4.31 Y N AR2-3 4 Qtz. N 0.1 Ort. 10.87 3.57 1.37 Y AR2-3 4 Qtz. N 0.2 Ort. 13.26 4.8 2.33 N AR2-3 4 Qtz. N 0.85 Ort. 13.44 12.74 6.45 AR2-3 5 BIF Y 0.15 Max. 10.62 5.7 2.36 Y 2 AR2-3 6 BIF Y 0.15 Ort. 11.71 6.32 1.82 N 2 5.86 AR2-3 6 Qtz. Y 0.05 Ort. 10.23 2.95 1.15 1 AR2-3 AR2-3 6 6 Qtz. Qtz. N Y 0.3 0.3 Ort. Ort. 17.22 13.05 7.8 4.01 2.39 5.95 Y AR2-3 6 Qtz. N 0.35 Ort. 13.81 8.29 2.94 N AR2-3 6 Qtz. N 0.4 Ort. 11.85 7.72 4.39 Y AR2-3 AR2-3 6 6 Qtz. Qtz. N N 0.7 0.9 Ort. Ort. 12.89 11.03 8.56 11.87 5.04 5.16 AR2-3 7 Qtz. N 0.15 Max. 10.62 7.87 2.82 1 AR2-3 7 Qtz. N 0.35 Ort. 13.34 5.33 3.12 3 4.45 1 < 25 AR2-3 AR2-3 7 7 Qtz. Qtz. N N 1.3 1.55 Ort. Ort. 20.69 17.02 10.72 10.27 5.67 5.41 N N 3 1 6.90 17.02 2 1 < 25 Y Y N N AR2-3 7 Qtz. N 1.75 Ort. 18.04 11.65 7.58 N 6 3.01 4 < 25 Y N N Bk type RM Qtz. XU 2 Pit/Sc. AR1 Trv. 3 6.70 2 < 25 Y N Y Y N 2 6.63 3 < 25 N N 2 6.72 2 < 25 Lng. Trv. Lng. N < 25 N N 1 < 25 N N 10.23 2 < 25 N N 2 3 8.61 4.35 1 1 < 25 < 25 N Y N 3 4.60 2 N N 2 5.93 2 < 25 N N 3 7 4.30 1.58 2 3 < 25 < 25 Y N N N Y N 438 T (mm) 5.88 AR2-3 11 Chal. Y 0.15 Ort. 11.36 5.84 0.85 Y AR2-3 AR2-3 11 12 Qtz. Qtz. N Y 0.15 0.15 Max. Max. 11.07 11 5.22 9 2.62 1.42 Y Y AR2-3 12 Qtz. N 0.15 Ort. 11.91 4.46 1.83 AR2-3 12 Qtz. N 0.6 Ort. 13.74 8.4 3.57 AR2-3 AR2-3 12 12 Qtz. Qtz. N N 1.05 2.5 Ort. Ort. 14.71 19.76 11.75 15.88 4.47 6.68 AR2-3 13 Qtz. Y 0.15 Max. 10.83 6.92 2.41 Y AR2-3 13 Qtz. N 0.6 Ort. 19.73 6.71 5.31 N AR2-3 13 Qtz. N 1.5 Ort. 16.16 16.02 4.76 AR2-3 13 Qtz. N 1.65 Ort. 19.55 12.72 5.53 AR2-3 13 Qtz. N 1.9 Ort. 12.03 12.37 AR2-3 14 BIF Y 0.45 Ort. 15 AR2-3 AR2-3 14 14 Qtz. Qtz. Y N 0.05 0.15 Ort. Max. AR2-3 14 Qtz. N 0.3 AR2-3 14 Qtz. N AR2-3 AR2-3 14 15 Qtz. Qtz. AR2-3 15 AR2-3 Lng. ED W (mm) 11.02 Bipolar L (mm) 13.02 %Unmod Dims. Ort. #Plat Wt (g) 0.75 mm/scar HQ N #NFS RM Qtz. Bk type XU 8 Bk Pit/Sc. AR2-3 4 3.26 3 < 25 N N 1 11.36 2 < 25 N N < 25 < 25 N N N 1 1 1 11.91 N 5 2.75 2 < 25 Y N N N 4 4 3.68 4.94 3 4 < 25 25-50 Y Y N Y Trv. 2 N 2 < 25 N N 2 9.87 2 < 25 Y N 2 8.08 3 N 4 4.89 2 6.63 N 5 2.41 2 12.58 2.92 N 3 5.00 1 11.64 10.9 2.59 9.74 1.76 2 Y Y Trv. 4 2 2.91 Ort. 10.39 7.2 2.73 Y Trv. 2 0.8 Ort. 15.54 8.49 4.78 N N Y 3.35 0.15 Ort. Ort. 18.92 13.24 16.12 6.62 7.83 1.09 N Y Qtz. N 0.2 Ort. 15.3 7.41 2.52 Y 15 Qtz. N 0.25 Ort. 10.06 6.46 2.06 Y AR2-3 AR2-3 15 16 Qtz. BIF N Y 1.1 0.25 Ort. Ort. 14.01 10.01 12 14.74 5.62 1.44 N AR2-3 16 Qtz. Y 0.1 Ort. 10.47 4.93 1.68 Y < 25 N N Y N < 25 N N 2 < 25 < 25 N N N N 5.20 2 < 25 N N 4 3.89 3 < 25 Y N Trv. 4 4 4.73 3.31 3 3 < 25 < 25 Y N N Y Mrg. 4 3.83 2 < 25 N N Trv. 3 3.35 3 < 25 N N 3 3 4.67 3.34 2 2 < 25 < 25 N N N N 2 5.24 1 < 25 N 439 Wt (g) Dims. L (mm) W (mm) T (mm) Bk %Unmod Bipolar ED 0.15 Max. 12.73 4.68 2.17 Y 1 < 25 N N AR2-3 16 Qtz. Y 0.15 Max. 12.62 4.26 2.09 Y 2 < 25 N N AR2-3 AR2-3 16 16 Qtz. Qtz. Y N 0.25 0.9 Ort. Ort. 13.09 16.72 5.71 16.09 2.06 2.33 < 25 N Y N N AR2-3 17 Qtz. N 0.1 Max. 10.9 7.42 2.54 Y 1 < 25 N N AR2-3 17 Qtz. Y 0.25 Max. 11.38 9.2 3.2 Y 3 < 25 N N AR2-3 AR2-3 17 17 Qtz. Qtz. N N 0.3 0.35 Ort. Ort. 13.1 11.23 6.07 6.26 2.57 3.11 Y < 25 < 25 N N N AR2-3 17 Qtz. N 1.45 Ort. 31.19 8.48 5.18 AR2-3 17 Sil. N 0.3 Ort. 12.64 6.7 2.1 AR2-3 18 BIF Y 0.2 Ort. 10.79 10.32 AR2-3 18 Qtz. N 0.1 Max. 13.11 AR2-3 18 Qtz. N 0.15 Ort. AR2-3 18 Qtz. N 0.15 AR2-3 AR2-3 18 18 Qtz. Qtz. N N AR2-3 18 Qtz. AR2-3 18 AR2-3 AR2-3 Trv. Trv. 2 1 6.55 16.72 #Plat HQ Y mm/scar RM Qtz. #NFS XU 16 Bk type Pit/Sc. AR2-3 2 2 3 4 4.37 2.81 2 3 N 1 31.19 2 Y N N 2 6.32 1 < 25 N N 1.47 N 1 10.79 1 < 25 N N 4.9 2.04 Y 3 < 25 N N 10.4 6.69 2.18 N 2 5.20 2 < 25 N N Ort. 10.32 7.24 1.57 N 3 3.44 3 < 25 N N 0.2 0.25 Ort. Max. 13.22 13.06 5.49 6.7 2.23 2.4 Y 3 2 4.41 2 < 25 N N N Y 0.25 Ort. 11.08 7.66 2.3 N 4 2.77 3 < 25 N N Qtz. N 0.25 Ort. 14.18 5.22 2.69 N 0 1 > 75 N N 18 18 Qtz. Qtz. N N 0.3 0.4 Ort. Ort. 13.79 12.46 5.97 10.38 2.68 2.46 N N 3 1 4.60 12.46 2 1 < 25 25-50 Y Y N N AR2-3 18 Qtz. N 0.5 Ort. 14.8 7.7 4.65 N 2 7.40 2 < 25 N N AR2-3 18 Qtz. Y 0.55 Max. 15.24 12.23 3.57 Y 2 AR2-3 AR2-3 18 18 Qtz. Qtz. N Y 0.65 0.65 Ort. Ort. 11.5 16.24 10.73 12.11 3.56 3.05 Y 3 2 3.83 8.12 AR2-3 18 Qtz. N 0.7 Ort. 17.99 6.49 5.11 N 3 6.00 Trv. Trv. < 25 Y 2 3 < 25 N N N 2 < 25 Y N 440 mm/scar #Plat %Unmod 4.44 3.92 N N 2 5 7.11 4.96 1 2 < 25 < 25 N Y N N 10.7 3.87 N 3 4.68 2 < 25 N N 10.27 4.91 1.87 Y 3 < 25 N Y Ort. Ort. 10.1 10.6 5.34 10.73 0.85 2.8 Y N Lng. 1 2 10.10 5.30 2 2 < 25 < 25 N N N Y 0.4 Ort. 20.18 6.56 3.06 Y Trv. 4 5.05 2 < 25 N Y 0.55 Ort. 10.14 10.04 2.94 Y Trv. 3 3.38 2 < 25 N N 0.7 Ort. 15.9 10.23 2.95 Y Trv. 2 7.95 1 < 25 N 1.2 Ort. 21.86 14.17 3.1 N 3 7.29 2 < 25 Y N Qtz. N 0.1 Ort. 12 5.35 1.25 N 2 6.00 1 < 25 N Y 20 Qtz. N 0.15 Max. 11.81 4.91 3.06 Y 3 AR2-3 AR2-3 20 20 Qtz. Qtz. N Y 0.15 0.15 Ort. Ort. 14.31 10.24 4.32 3.99 1.3 2.04 N Y AR2-3 20 Qtz. N 0.15 Max. 11.64 8.08 1.48 Y 1 AR2-3 20 Qtz. N 0.2 Ort. 11.61 7.01 1.93 N 0 AR2-3 AR2-3 20 20 Qtz. Qtz. Y N 0.2 0.2 Ort. Ort. 13.78 13.02 6.23 6.18 1.73 1.91 N Y 2 3 AR2-3 20 Qtz. N 0.25 Ort. 13.13 7.8 2.47 N AR2-3 20 Qtz. Y 0.35 Ort. 12.83 7.14 3.28 AR2-3 AR2-3 20 20 Qtz. Qtz. N Y 0.4 0.45 Ort. Ort. 10.75 12.12 10.42 12.76 AR2-3 20 Qtz. N 0.7 Ort. 12.63 10.01 Qtz. N 0.75 Ort. 16.17 11.58 4.31 Y AR2-3 18 Qtz. N 1.05 Ort. 15.26 9.52 4.6 AR2-3 AR2-3 18 18 Qtz. Qtz. N Y 1.15 1.55 Ort. Ort. 14.22 24.82 10.14 12.59 AR2-3 19 Chal. Y 1.2 Ort. 14.04 AR2-3 19 Qtz. Y 0.05 Max. AR2-3 AR2-3 19 19 Qtz. Qtz. Y N 0.1 0.3 AR2-3 19 Qtz. N AR2-3 19 Qtz. Y AR2-3 19 Qtz. Y AR2-3 19 Qtz. AR2-3 20 AR2-3 N N N < 25 < 25 N N N N 25-50 N N 1 > 75 N N 6.89 4.34 1 2 < 25 < 25 N N N N 1 13.13 1 25-50 N Y N 4 3.21 2 < 25 N N 2.8 2.87 N Y 3 4 3.58 3.03 3 2 25-50 < 25 Y N N N 4.23 N 2 6.32 3 < 25 Y N Trv. Trv. Trv. 2 3 ED #NFS N 18 Bipolar Bk type Y Bk < 25 T (mm) 2 W (mm) 3.05 L (mm) 5 Dims. N Wt (g) < 25 HQ 2 RM 5.39 XU 3 Pit/Sc. Mrg. AR2-3 7.16 3.41 1 2 441 XU RM HQ Wt (g) Dims. L (mm) W (mm) T (mm) Bk #NFS mm/scar #Plat %Unmod Bipolar ED 20 Qtz. N 1.8 Ort. 17.05 13.3 5.45 N 2 8.53 3 < 25 N N AR2-3 20 Qtz. N 1.9 Ort. 20.6 12.13 5.41 N 2 10.30 2 25-50 Y AR2-3 AR2-3 20 20 Qtz. Qtz. N N 2.85 9.45 Ort. Ort. 28.47 29.97 12.47 24.31 6.67 8.75 N N 3 6 9.49 5.00 3 4 25-50 < 25 N Y N N AR2-3 21 Qtz. Y 0.05 Ort. 11.37 5.97 1.61 Y 2 5.69 < 25 N N AR2-3 21 Qtz. N 0.15 Ort. 14.19 6.71 1.88 N 2 7.10 2 25-50 N N AR2-3 AR2-3 21 21 Qtz. Qtz. Y N 0.2 0.25 Ort. Ort. 10.88 10.18 7.52 8.88 2.53 1.59 Y 3 2 3.63 5.09 1 3 < 25 < 25 Y N N N AR2-3 21 Qtz. N 0.3 Ort. 12.41 5.34 3.82 N 4 3.10 2 < 25 Y N AR2-3 21 Qtz. N 0.5 Ort. 11.16 12.33 2.63 N 2 5.58 1 < 25 N N AR2-3 21 Qtz. N 1.4 Ort. 14.07 14.83 4.22 N 3 4.69 2 < 25 Y N AR2-3 21 Qtz. N 2.35 Ort. 18.5 17.82 4.47 N 4 4.63 2 < 25 Y N AR2-3 23 Qtz. N 0.5 Ort. 11.33 7.81 4.28 N 1 11.33 2 < 25 Y Y AR2-3 23 Qtz. N 0.75 Ort. 19.14 9.36 3.78 N 2 9.57 1 Y Y AR2-3 AR2-3 23 24 Qtz. Qtz. N N 8.8 0.25 Ort. Ort. 30.26 12.63 20.97 8.45 12.1 2.12 N N 2 3 15.13 4.21 2 > 50-75 < 25 N N N N AR2-3 24 Qtz. N 1.65 Ort. 17.2 18.79 4.31 N 3 5.73 2 < 25 N N AR2-3 24 Qtz. N 1.8 Ort. 20.52 11.85 4.34 N 3 6.84 2 < 25 Y N AR2-3 AR2-3 25 30 Qtz. Qtz. N Y 0.65 0.2 Ort. Ort. 18.08 10.12 8.18 10.69 4.4 1.49 N Y 2 2 9.04 5.06 1 1 < 25 Y N N N AR2-3 30 Qtz. N 0.25 Max. 18.52 8.03 1.73 Y 1 N N AR2-3 30 Qtz. N 1.05 Ort. 19.37 11.63 3.37 N 2 9.69 2 25-50 Y N AR2-3 AR2-3 31 32 Qtz. Qtz. Y Y 0.1 0.15 Ort. Ort. 10.76 13.2 4.87 4.01 1.34 1.73 N N 3 3 3.59 4.40 1 2 < 25 < 25 N Y N N Bk type Pit/Sc. AR2-3 Trv. Trv. Trv. 442 Wt (g) Dims. L (mm) W (mm) T (mm) Ret. Bk Qtz. Y 0.2 Ort. 8.73 7.43 1.98 Bck. N N Sc1 Sc3 - Qtz. Qtz. N Y 0.85 0.15 Ort. Ort. 14.48 10.09 10.95 5.15 3.9 2.15 Bck. Bck. N N N Sc3 - Qtz. Y 0.5 Ort. 11.39 10.46 3.28 Bck. N N OSM - Qtz. N 1.35 Ort. 21.87 12.45 3.66 Bck. N N AR1 AR1 3 4 Qtz. Qtz. N N 0.85 0.65 Ort. Ort. 12.91 16.55 9.91 9.3 5.15 3.36 Bck. Bck. Y N Trv. N Y AR1 5 Qtz. Y 0.1 Ort. 7.5 4.77 1.37 Bck. Y Lng., Trv. N AR1 8 Qtz. N 0.05 Ort. 7.84 4.47 1.68 Bck. N N AR1 AR1 8 10 Qtz. Qtz. N Y 0.15 < 0.05 Ort. Ort. 8.91 6.54 6.99 4.99 1.81 0.83 Bck. Bck. N Y Lng., Trv. N N AR1 11 Qtz. Y 0.15 Ort. 7.22 4.73 2.74 Bck. Y Trv. N AR2-3 10 Qtz. Y 0.15 Ort. 9.92 7.68 1.77 Bck. N N AR2-3 AR2-3 11 13 Qtz. Qtz. Y Y 0.3 0.2 Ort. Ort. 17.38 14.27 5.14 4.6 2.44 2.18 Bck. Bck. N N N N AR2-3 14 Qtz. N 1.85 Ort. 15.62 12.63 4.81 Uni. N N AR2-3 15 Qtz. Y 0.15 Ort. 12.49 5.71 2.11 Uni. Y AR2-3 AR2-3 16 16 Qtz. Qtz. N Y 0.3 0.45 Ort. Ort. 13.14 14.72 8.25 7.21 2.72 3.38 Bck. Bck. N N Y N AR2-3 16 Qtz. N 0.65 Max. 15.78 9.6 4.62 Uni. AR2-3 17 Chal. Y 0.5 Ort. 17.81 8.62 2.86 Bck. N N AR2-3 AR2-3 17 18 Qtz. Chal. Y Y 0.3 0.2 Ort. Ort. 9.28 14.43 6.88 5.09 2.36 1.7 Bck. Bck. N Y N N Trv. Trv. ED HQ - Bk type RM Sc1 Pit/Sc. XU E.6 RETOUCHED FLAKES N 443 Pit/Sc. XU RM HQ Wt (g) Dims. L (mm) W (mm) T (mm) Ret. Bk Bk type ED AR2-3 18 Qtz. Y 0.25 Ort. 11.68 7.44 1.51 Bck. Y Mrg. N AR2-3 18 Qtz. Y 0.3 Ort. 8.22 7.68 3.96 Bck. Y Trv. N AR2-3 AR2-3 18 18 Qtz. Qtz. N N 0.35 0.4 Ort. Ort. 14.49 17.53 6.18 7.9 2.56 2.35 Bck. Bck. Y Y Trv. Trv. N N AR2-3 18 Qtz. Y 0.4 Ort. 12.95 7.22 3.99 Bck. Y Trv. N AR2-3 18 Qtz. Y 0.5 Ort. 15.7 10.04 2.48 Bck. Y Trv. Y AR2-3 AR2-3 18 18 Qtz. Qtz. Y N 0.5 0.9 Ort. Ort. 11.02 18.8 10.42 9 3.74 3.23 Bck. Uni. N Y Trv. N N AR2-3 18 Qtz. N 2.7 Ort. 18.95 15.43 5.34 Bck. N AR2-3 19 Chal. Y < 0.05 Ort. 4.71 4.89 0.89 Bck. Y Trv. N AR2-3 19 Qtz. N 0.1 Ort. 6.11 5.57 2.33 Bck. Y Trv. N AR2-3 20 Qtz. N 0.2 Ort. 12.58 6.25 1.41 Bck. N N AR2-3 20 Qtz. Y 0.9 Ort. 20.37 9.61 3.47 Bck. N N AR2-3 20 Qtz. N 1.8 Ort. 22.67 14.5 4.41 Bck. N N AR2-3 AR2-3 21 21 Qtz. Sil. Y N 0.55 4.9 Ort. Ort. 12.1 43.05 13.29 16.93 2.39 6.52 Uni. Uni. N N N AR2-3 23 Qtz. N 0.35 Ort. 9.61 7.96 2.63 Bck. N N N 444 APPENDIX F GIBB ROCK STONE ARTEFACT ASSEMBLAGE This appendix contains the raw artefact data for the Gibb Rock stone artefact assemblages. Weight (grams) and dimensions (millimetres) have been rounded to two decimal places. Blank cells denote unusable or missing data. Unusable data was that deemed not representative of a particular attribute, i.e. mm/scar for a flake that could not be oriented. Where data were absent, this indicated that a particular attribute could not be confidently measured on a certain specimen, often due to poor-quality raw material. Artefact types have been ordered by pit/scatter (beginning with surface material), XU depth and raw material. See Tables 9.1–9.2 and Appendix D for details on attributes recorded for each artefact type. Abbreviations are as follows: o Bk = breakage o Sc. = scatter o Comb wt. = combined weight o Sc1 = Scatter 1 o Dims = dimensions o Sc2 = Scatter 2 o ED = edge damage o Sil. = silcrete o HQ = high-quality raw material o T = thickness o L = length o Trans. = transitional o Max. = maximum dimensions o %Unmod = proportion of o #NFS = no. of negative flake scars o Ort. = oriented dimensions o W = width o #Plat = no. of platforms o Wt = weight o Qtz = quartz o XU = excavation unit o RM = raw material unmodified surface 445 Pit/Sc. XU RM HQ Wt (g) L (mm) W (mm) T (mm) #NFS g/scar #Plat %Unmod Bipolar F.1 CORES Sc1 Sc1 - Qtz Qtz N N 62.1 143 63.83 86.29 36.02 39.44 19.61 38.88 4 2 15.53 71.50 2 2 > 50-75 > 75 N Sc2 - Qtz N 4.9 20.17 17.74 10.58 9 0.54 4 < 25 Y - trans Sc2 - Qtz N 11.35 28.69 23.37 18.14 4 2.84 Sc2 Sc2 - Qtz Sil. N N 11.6 232 28.7 69.59 23.15 54.69 16.19 59 11 1 1.05 232.00 4 1 < 25 > 75 N N F.2 SMALL FLAKED FRAGMENTS XU RM No. Comb. Wt (g) No. HQ Sc1 GR1 1 Qtz Qtz 1 4 0.2 0.35 0 0 GR1 2 Qtz 3 0.15 0 GR1 3 Qtz 4 0.4 0 GR1 GR1 4 5 Qtz Qtz 1 1 0.1 0.25 0 0 Pit/Sc. 446 RM HQ Wt (g) L (mm) W (mm) T (mm) #NFS g/scar #Plat Sc1 - Qtz N 0.3 12.58 5.72 3.78 3 0.10 2 Sc1 - Qtz N 0.4 10.18 8.09 5.92 1 0.40 1 Sc1 Sc1 - Qtz Qtz N N 0.5 0.55 14.13 14.56 5.84 9.53 5.42 4.59 2 0.28 2 Sc1 - Qtz N 0.7 13.11 9.71 6.8 2 0.35 2 Sc1 - Qtz N 0.7 13.31 9.04 7.24 Sc1 Sc1 - Qtz Qtz N N 0.7 1 15.8 16.81 7.94 11.85 5.44 5.49 1 2 0.70 0.50 1 1 Sc1 - Qtz N 1.25 15.59 10.96 7.15 2 0.63 2 Sc1 - Qtz N 1.3 22.83 8.85 5.27 Sc1 - Qtz N 1.35 20.27 7.8 6.38 Sc1 - Qtz N 1.45 16.48 13.06 5.59 1 1.45 1 Sc1 - Qtz N 1.5 18.16 12.42 8.19 Sc1 - Qtz N 1.7 18.28 9.66 7.05 1 1.70 1 Sc1 Sc1 - Qtz Qtz N N 2.85 2.9 18.23 22.43 11.55 17.32 8.55 10.04 2 1 1.43 2.90 1 1 Sc1 - Qtz N 4.45 25.43 13.1 9.3 2 2.23 2 Sc1 - Qtz N 10.9 39.28 17.31 14.21 1 10.90 1 Sc1 Sc1 - Qtz Qtz N N 15.2 15.65 37.94 39.77 29.41 22.65 13.09 17.04 7 2.24 3 Sc1 - Qtz N 45.75 45.71 34.32 24.13 Pit/Sc. XU F.3 LARGE FLAKED FRAGMENTS F.4 SMALL FLAKES XU RM No. Comb. Wt (g) No. HQ Sc1 - Qtz 3 1.55 0 Sc2 GR1 1 Qtz Qtz 1 8 0.25 0.6 0 0 0 GR1 2 Qtz 1 < 0.05 0 1 GR1 3 Qtz 2 0.15 0 GR1 4 Qtz 1 0.05 0 Pit/Sc. No. broken 0 447 #Plat %Unmod Bipolar ED #NFS - Qtz N 0.2 Ort. 12.43 6.46 1.96 N 2 Sc1 Sc1 - Qtz Qtz N N 0.2 0.35 Ort. Ort. 12.24 13.34 8.84 5.72 2.52 2.98 Y N 1 2 12.24 6.67 Sc1 - Qtz N 0.45 Ort. 10.04 7 4.23 N 0 Sc1 - Qtz N 0.45 Max. 10.87 9.37 4.87 1 Sc1 Sc1 - Qtz Qtz N N 0.45 0.45 Ort. Ort. 10.95 12.13 14.08 5.89 2.68 3.16 N 2 2 5.48 6.07 3 1 Sc1 - Qtz N 0.5 Ort. 14.82 7.68 3.55 N 1 14.82 1 Sc1 - Qtz N 0.5 Ort. 13.08 9 3.1 N 2 6.54 2 < 25 Sc1 Sc1 - Qtz Qtz N N 0.55 0.55 Ort. Ort. 11.91 13 8.19 8.97 4.12 4.53 N 1 11.91 2 > 50-75 Sc1 - Qtz N 0.55 Ort. 12.57 8.2 4.16 N 2 6.29 1 < 25 N N Sc1 - Qtz N 1 Ort. 12.19 10.57 4.21 N 3 4.06 1 < 25 Y N Sc1 Sc1 - Qtz Qtz N N 1.1 1.15 Ort. Ort. 18.17 16.99 10.7 9.64 3.08 3.99 N N 3 3 6.06 5.66 2 2 < 25 < 25 Y N N Sc1 - Qtz N 1.4 Ort. 16.49 13.62 3.79 N 1 16.49 1 25-50 N N Sc1 - Qtz N 1.65 Ort. 18.23 10.72 6.96 N 3 6.08 2 < 25 Sc1 Sc1 - Qtz Qtz N N 1.7 1.7 Ort. Ort. 23.28 17.24 10.26 9.78 4.77 6.43 N N 3 3 7.76 5.75 2 1 < 25 < 25 Y Y N N Sc1 - Qtz N 1.8 Ort. 13.63 18.73 4.58 N 3 4.54 3 < 25 Y N Sc1 - Qtz N 2.55 Ort. 16.22 16.13 7.1 N 5 3.24 3 < 25 Y N Sc1 Sc1 - Qtz Qtz N N 2.6 3 Ort. Ort. 20.91 18.68 16.3 19.57 5.57 5.3 N N 4 2 5.23 9.34 2 1 < 25 < 25 N Y N N Bk type Sc1 Pit/Sc. Bk N > 75 T (mm) Y 1 W (mm) 1 < 25 < 25 L (mm) N Dims. Y Wt (g) < 25 HQ 2 RM 6.22 XU mm/scar F.5 LARGE FLAKES N N < 25 < 25 N Y N N N N Y N N N N 448 RM HQ Wt (g) Dims. L (mm) W (mm) T (mm) Bk #NFS mm/scar #Plat %Unmod Bipolar ED Qtz N 3.85 Ort. 25.77 17.32 7.42 N 2 12.89 1 25-50 Y N Sc1 Sc1 - Qtz Qtz N N 3.9 4.35 Ort. Max. 22.84 30.37 18.7 22.44 6.77 6.66 N N 1 1 22.84 1 25-50 > 50-75 Y N N Sc2 - Qtz N 0.95 Ort. 11.16 13 5.06 N 3 3.72 2 < 25 N N Sc2 - Qtz N 3.6 Ort. 27.08 17.93 5.44 N 0 1 > 75 Y N GR1 GR1 1 1 Qtz Qtz N N 0.2 0.35 Ort. Ort. 10.48 11.78 7.63 6.82 2 4.64 N N 2 3 N N N GR1 2 Qtz N 0.35 Ort. 11.33 9.62 2.33 N Bk type XU - Pit/Sc. Sc1 5.24 3.93 2 2 < 25 N 449 APPENDIX G MULKA’S CAVE STONE ARTEFACT ASSEMBLAGE This appendix contains the raw artefact data for the Mulka’s Cave stone artefact assemblages. Weight (grams) and dimensions (millimetres) have been rounded to two decimal places. Blank cells denote unusable or missing data. Unusable data was that deemed not representative of a particular attribute, i.e. mm/scar for a flake that could not be oriented. Where data were absent, this indicated that a particular attribute could not be confidently measured on a certain specimen, often due to poor-quality raw material. Artefact types have been ordered by site area, pit/scatter (beginning with surface material), XU depth and raw material. See Tables 9.1–9.2 and Appendix D for details on attributes recorded for each artefact type. Abbreviations are as follows: o Bck. = backed o Ort. = oriented dimensions o BIF = Banded Iron Formation o Oth. = other surface artefacts o Bk = breakage o #Plat = no. of platforms o CAS = Camping Area Scatter o Qtz = quartz o Chal. = chalcedony o Ret. = retouch type o Cl. = cluster o RM = raw material o Comb wt. = combined weight o Sc. = scatter o CS = Column Sample o Sil. = silcrete o Dims = dimensions o T = thickness o ED = edge damage o Trans. = transitional o EP = Entrance Platforms o Trv. = transverse o HQ = high-quality raw material o Uni = unifacial o HS = Humps Scatter o %Unmod = proportion of o L = length o Max. = maximum dimensions o W = width o Mrg. = marginal o Wt = weight o #NFS = no. of negative flake scars o XU = excavation unit unmodified surface 450 Pit/Sc. XU/Cl. RM HQ Wt (g) L (mm) W (mm) T (mm) #NFS g/scar #Plat %Unmod Bipolar G.1 CORES EP CS MC1 HS HS 8 4 - Qtz Qtz Qtz Qtz Qtz N Y Y N Y 1.75 3.25 0.65 2.2 2.55 16.28 24.11 13.29 18.72 19.89 11.89 13.12 11.65 16.37 17.39 10.51 8.49 4.15 10.39 9.33 4 6 3 6 6 0.44 0.54 0.22 0.37 0.43 3 4 1 3 2 < 25 < 25 25-50 < 25 < 25 Y Y - trans Y Y Y - trans HS CAS B Qtz Qtz Y Y 3.05 2.35 18.36 16.58 14.14 16.54 10.16 11.57 7 10 0.44 0.24 3 4 < 25 < 25 Y - trans Y CAS CAS CAS CAS CAS CAS CAS CAS B B B B D D F Oth. Qtz Qtz Sil. Sil. Qtz Qtz Qtz Qtz N N N N N Y N N 3.05 6.05 14.15 60.2 0.85 2.65 16.55 14.3 19.09 21.96 35.43 59.53 10.93 17.69 33.78 31.02 18.47 19.77 30.18 45.51 11.53 14.76 21.92 27.55 10.28 16.37 15.59 27.63 8.58 10.33 16.65 19.17 11 9 10 4 6 10 3 11 0.28 0.67 1.42 15.05 0.14 0.27 5.52 1.30 4 3 4 1 2 6 1 5 < 25 < 25 < 25 > 50-75 < 25 < 25 > 75 < 25 Y - trans N N N Y Y Y N CA1 CA1 CA1 CA1 CA1 CA2 3 3 4 4 4 4 Qtz Qtz Qtz Qtz Qtz Qtz N N N N N N 2.5 16.15 0.95 4.3 82.15 23.15 18.23 43.29 14.81 25.38 58.5 34.74 16.28 21.41 9.37 15.44 52.4 27.93 8.27 19.77 9.14 13.86 40.23 20.68 6 12 4 5 9 16 0.42 1.35 0.24 0.86 9.13 1.45 2 5 4 3 4 6 < 25 < 25 < 25 25-50 25-50 < 25 Y N Y Y - trans N N 451 G.2 SMALL FLAKED FRAGMENTS XU/Cl. 5 9 RM Qtz Qtz No. 1 2 Comb. Wt (g) < 0.05 0.15 No. HQ 1 2 11 - Qtz Qtz 2 1 0.15 0.1 2 1 TH1 TH1 TH1 TH1 TH1 TH1 CAS CAS CA1 CA1 CA1 CA1 CA1 CA1 CA1 1 2 3 4 5 6 C Oth. 1 2 3 5 6 7 8 Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz 8 15 10 13 6 3 2 1 4 5 10 7 8 11 4 0.6 1.25 1.25 1.4 0.4 0.3 0.2 0.35 0.45 0.85 1.35 0.7 0.95 1.5 0.4 4 7 0 0 0 0 2 0 0 1 3 0 0 3 2 CA1 CA2 CA2 CA2 CA2 CA2 CA2 9 1 2 3 4 5 6 Qtz Qtz Qtz Qtz Qtz Qtz Qtz 1 5 6 7 13 12 3 0.2 0.6 0.6 0.9 2.35 2 0.45 0 0 0 0 3 2 2 Pit/Sc. CS CS CS HS 452 XU/Cl. RM HQ Wt (g) L (mm) W (mm) T (mm) #NFS g/scar EP EP CS CS CS CS CS MC1 5 5 6 7 9 3 Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz N N N Y N N Y N 0.9 4.25 0.4 1.8 0.25 1.25 0.1 2.25 15.27 25.32 11.81 16.63 11.9 17.73 10.38 18.56 9.31 18.14 6.93 15.79 5.05 9.82 5.73 13.89 7.43 10.56 5.08 7.88 4.22 6.01 2.75 7.67 1 1 3 6 2 1 2 3 0.90 4.25 0.13 0.30 0.13 1.25 0.05 0.75 1 1 2 3 MC1 HS 4 - Qtz Qtz N Y 8 2 32.7 21.26 20.11 12.93 14 7.16 1 7 8.00 0.29 1 3 HS TH1 TH1 TH1 TH1 TH1 TH1 TH1 TH1 TH1 TH1 TH1 TH1 TH1 TH1 TH1 TH1 TH1 TH1 TH1 CAS CAS CAS CAS CAS CAS CAS CAS CAS CAS CAS 1 1 1 2 2 2 3 3 3 3 4 4 4 4 4 4 5 5 5 A A A B B B C C C D D Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz N N N N N N N N N N N N N N N N N N N N N N N Y N N N N N N N 2.15 0.25 0.45 0.55 0.25 0.3 0.5 0.3 0.6 0.65 0.7 0.15 0.25 0.3 0.3 0.5 0.55 0.15 0.2 0.3 0.6 0.9 1.7 1.55 4.25 6.6 0.8 2.65 3.95 1.55 7.05 19.23 10.76 13.03 15.63 11.18 10.5 15.71 11.65 13.6 12.33 14.55 12.62 12.7 12.02 11.14 13.49 12.98 11.39 10.7 8.78 13.91 16.8 19.45 22.58 24.58 29.19 17.37 21.39 24.77 17.32 27.57 17.37 5.47 9.42 7.26 5.86 6.31 6.24 6.07 9.05 8.25 9.19 4.45 5.04 6.54 6.63 6.88 6.83 5.05 6.89 6.36 6.67 8.95 10.71 10.49 18.87 20.57 7.19 14.83 18.74 16.68 19.93 6.58 4.01 3.68 5.39 4.2 3.5 4.18 3.93 6.49 6.94 4.59 3.06 4.6 4.16 3.49 3.67 5.01 2.86 3.24 3.67 6.2 5.9 9.84 7.85 12.01 13.19 6.41 9.64 11.3 6.62 13.54 5 2 2 3 1 2 1 1 5 1 2 1 2 2 1 1 0.43 0.13 0.23 0.18 0.25 0.15 0.50 0.30 0.12 0.65 0.35 0.15 0.13 0.15 0.30 0.50 2 2 1 2 1 2 1 1 3 1 2 1 2 2 1 1 1 1 1 3 3 2 4 2 9 3 2 7 4 5 0.15 0.20 0.30 0.20 0.30 0.85 0.39 2.13 0.73 0.27 1.33 0.56 0.39 1.41 1 1 1 1 2 1 2 1 4 1 2 2 2 2 #Plat Pit/Sc. G.3 LARGE FLAKED FRAGMENTS 1 1 3 453 RM HQ Wt (g) L (mm) W (mm) T (mm) #NFS g/scar Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz N N Y N N N N N 4.75 6.75 0.15 0.4 0.6 0.65 0.8 0.85 22.04 27.2 10.99 10.26 12.08 12.12 16.71 14.84 18.02 26.45 5.92 9.34 6.98 11.72 7.71 9.28 11.81 11.26 2.71 4.96 5 6.21 6.29 5.56 6 1 2 2 2 3 4 4 0.79 6.75 0.08 0.20 0.30 0.22 0.20 0.21 3 1 2 1 CAS CAS Oth. Oth. Qtz Qtz N Y 1.55 2 17.77 20.34 9.67 12.53 9.31 7.95 3 3 0.52 0.67 1 CAS CAS CAS CAS Oth. Oth. Oth. Oth. Qtz Qtz Qtz Qtz N N N N 2.15 2.45 2.85 3.15 23.03 16.36 20.25 22.62 12.29 15.02 12.37 16.61 6.47 11.9 11.34 9.94 3 3 4 3 0.72 0.82 0.71 1.05 1 2 2 1 CAS CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 Oth. 2 2 2 4 4 4 5 5 5 5 5 5 Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz N N N N N N N N N N N N N 4.05 0.8 1.1 1.6 0.1 0.15 1.75 0.15 0.3 0.65 1.15 1.6 2 32.5 14.4 21.81 18.48 10.15 11.6 21.28 11.83 10.84 18.21 17.97 17.65 20.01 13.66 7.31 7.2 13.03 5.48 6.26 8.71 3.77 6.21 7.68 14.56 12.82 14.61 11.02 7.23 7.07 7.94 2.32 5.31 7.48 3.24 4.16 4.77 4.89 8.67 9.27 1 2 4.05 0.40 1 2 3 0.53 2 3 1 0.05 1.75 3 1 2 0.15 2 1 3 1 1.15 0.53 2.00 1 2 1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA2 CA2 CA2 6 7 7 7 7 7 7 7 7 8 8 8 1 1 2 Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz N N N N N N N N N Y N N Y Y Y 0.9 0.3 0.45 0.45 0.55 0.55 0.6 0.85 0.9 0.25 0.35 1.65 0.25 0.5 0.2 11.39 13.69 10.69 10.81 12.02 12.63 13.93 12.72 16.99 10.31 11.42 15.29 12.91 12.69 13.39 10.94 6.35 7.37 7.83 6.32 6.25 7.11 10.02 6.77 5.27 5.42 11.09 4.01 7.78 4.36 7.87 3.67 5.97 5.3 6.08 6.07 6.16 5.81 6.3 3.97 4.55 8.89 3.41 5.03 4.19 1 1 2 1 4 0.90 0.30 0.23 0.45 0.14 1 1 3 2 0.20 0.43 5 0.05 3 2 1 1 0.55 0.13 0.50 0.20 #Plat XU/Cl. E F Oth. Oth. Oth. Oth. Oth. Oth. Pit/Sc. CAS CAS CAS CAS CAS CAS CAS CAS 2 1 3 2 1 1 454 Pit/Sc. XU/Cl. RM HQ Wt (g) L (mm) W (mm) T (mm) #NFS g/scar #Plat CA2 CA2 CA2 CA2 CA2 CA2 CA2 CA2 2 3 4 4 4 5 5 5 Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz N N N N Y N N N 1.1 0.15 0.2 0.35 1.5 0.4 0.5 0.6 18.24 11.14 10.34 10.66 20.62 12.58 11.65 11.45 7.94 5.16 5.73 7.04 10.1 5.47 11.17 9.32 7.66 3.06 3.64 5.55 8.65 4.26 4.67 5.42 3 1 2 3 3 1 2 3 0.37 0.15 0.10 0.12 0.50 0.40 0.25 0.20 2 1 2 3 3 1 2 2 CA2 CA2 6 6 Qtz Qtz N N 0.45 0.55 10.62 13.42 6.55 6.24 5.1 5.61 3 2 0.15 0.28 CA2 CA2 CA2 6 6 6 Qtz Qtz Qtz N N N 0.9 2 2.5 13.62 15.43 19.65 9.17 11.61 14.29 6.26 11.07 8.1 4 2 3 0.23 1.00 0.83 1 2 455 G.4 SMALL FLAKES XU/CL. RM No. Comb. Wt (g) No. HQ No. broken CS CS CS CS CS CS CS CS CS 2 4 5 6 7 8 9 10 11 Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz 2 3 2 6 6 4 5 7 1 0.05 0.15 0.15 0.45 0.5 0.3 0.5 0.3 0.3 1 0 0 2 6 4 5 7 1 0 0 0 0 4 3 3 1 1 MC1 MC1 5 6 Qtz Qtz 1 1 0.4 0.35 0 0 1 1 HS TH1 TH1 TH1 TH1 TH1 TH1 CAS CAS CAS CAS CAS CAS CA1 CA1 1 2 3 4 5 6 A B C D F Oth. 1 2 Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Chal. 4 14 14 6 9 6 3 4 3 2 2 1 4 3 1 1.65 1.2 2.25 0.45 1.2 0.55 0.2 1.2 1.1 0.3 0.5 0.05 0.85 0.3 0.1 3 0 0 4 3 0 1 1 3 2 1 1 1 0 1 0 CA1 CA1 2 2 Qtz Sil. 15 3 1.3 0.1 8 0 3 1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA2 CA2 3 3 3 4 5 5 5 6 6 7 8 9 1 1 BIF Qtz Sil. Qtz BIF Qtz Sil. Chal. Qtz Qtz Qtz Qtz BIF Qtz 1 31 1 29 2 24 3 1 13 11 3 5 1 6 0.1 4.25 0.15 2.55 0.15 2.5 0.25 0.1 1.1 1.9 0.35 0.45 0.1 0.95 1 16 0 20 2 5 0 1 13 7 2 0 1 4 1 6 0 6 0 3 2 0 5 1 0 0 0 1 Pit/Sc. 3 0 2 0 0 1 0 1 456 Pit/Sc. CA2 CA2 CA2 CA2 CA2 CA2 CA2 XU/CL. 2 2 3 3 4 5 6 RM BIF Qtz Qtz Sil. Qtz Qtz Qtz No. 1 17 28 1 25 11 1 Comb. Wt (g) 0.05 1.05 3.45 0.15 4.2 1.45 0.05 No. HQ 1 14 18 0 5 3 1 No. broken 1 6 9 1 3 2 0 457 5.88 5.18 7 9.84 8.59 5.32 1.29 0.93 2.19 3.8 4.5 3.53 3.81 N N N N N N N N CS CS CS CS CS CS CS MC1 MC1 MC1 MC1 9 9 9 9 9 9 9 3 3 4 4 Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz N N Y Y N Y N Y N Y N 0.2 0.2 0.25 0.25 0.3 0.3 1 0.65 1.45 0.75 0.75 Ort. Ort. Ort. Max. Ort. Ort. Ort. Ort. Ort. Ort. Ort. 12.8 10.45 10.28 12.08 10.01 12.71 14.77 14.65 15.37 16.04 15.33 4.71 5.26 9.28 7.61 6.68 7.72 9.63 10.13 16.9 10.19 8.01 3.3 2.73 1.89 3.81 3.64 2.04 6.19 3.39 3.02 3.24 3.78 N N Y N N Y N N N N N N 3 3 4 5 2 4 2 3 2 4 1 2 2 5.72 5.31 5.03 3.15 7.86 3.80 5.06 3.93 3 2 2 2 2 3 N N N Y Y N N N N 3 < 25 < 25 < 25 < 25 < 25 < 25 < 25 < 25 < 25 < 25 4.65 6.03 1 < 25 25-50 N N 2 2 2 1 0 2 4 3 6.40 5.23 5.14 1 2 1 N 6.36 3.69 4.88 1 1 2 3 < 25 25-50 < 25 < 25 > 75 < 25 < 25 < 25 3 2 5.35 7.67 1 3 < 25 < 25 2 N N N N N N N ED 12.21 14.4 15.41 12.82 14.41 7.35 6.67 4.06 7.31 9.23 11.94 5.9 9.54 Bipolar T (mm) 17.16 15.94 20.11 15.77 15.71 15.21 10.12 11.78 12.81 18.6 18.38 11.2 12.05 %Unmod W (mm) Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Max. Ort. Max. Max. Ort. #Plat L (mm) 1.5 1.55 2.45 2.6 2.65 0.6 0.1 0.05 0.2 0.85 0.95 0.1 0.5 mm/scar Dims. N N N N N Y N Y N N N N Y #NFS Wt (g) Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Bk type HQ 5 6 7 7 7 7 8 8 Bk RM EP EP EP EP EP CS CS CS CS CS CS CS CS Pit/Sc. XU/Cl. G.5 LARGE FLAKES Y Y N N N N N N N N N N N N N N N N Y N Y N N N 458 W (mm) T (mm) Bk 21.66 12.01 12.74 16.91 6.76 15.22 9.49 9.5 4.88 3.94 4.58 5.69 3.28 5.65 2.06 3.3 N N N N Y N Y Y HS HS - Qtz Qtz Y Y 0.85 1.25 Ort. Ort. 17.04 15.65 8.66 11.85 2.98 4.91 N Y HS HS HS HS HS TH1 TH1 TH1 TH1 TH1 TH1 TH1 1 1 1 1 2 2 2 Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Y Y N N N Y N N N N N N 1.3 1.45 2 2.9 2.9 0.2 0.55 0.6 3.7 0.15 0.45 2.05 Max. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. 20.56 14.59 28.5 22.47 16.33 11.04 10.29 10.19 22.95 10 13.58 23.53 12.35 10.05 11.88 15.47 11.24 6.44 13.4 12.26 15.03 5.18 5.6 15.5 4.74 5.71 5.32 5.8 8.98 2.13 4.24 4.22 10.84 2 4.06 5.11 N N N N Y Y N N N N N TH1 TH1 TH1 TH1 3 3 4 5 Qtz Qtz Qtz Qtz N N N N 0.15 0.25 1.1 0.05 Ort. Ort. Ort. Ort. 10.81 11.99 19.88 10.11 6.43 6.22 12.22 3.62 2.07 2.84 4.46 1.13 27.30 5.13 9.55 19.19 2 2 2 1 Trv. Trv. 1 3 2 1 2 4 4 2 4.83 3.57 6.49 Trv. 2 2 Trv. Trv. N Y Trv. Y Trv. ED L (mm) 27.3 15.39 19.1 19.19 11.54 19.33 14.28 12.97 Bipolar Dims. Ort. Ort. Ort. Ort. Max. Ort. Ort. Ort. %Unmod Wt (g) 4 0.85 1.4 2.45 0.25 2.5 0.4 0.5 #Plat HQ N N N N N N Y Y mm/scar RM Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz #NFS XU/Cl. 4 5 5 5 6 6 - Bk type Pit/Sc. MC1 MC1 MC1 MC1 MC1 MC1 HS HS 2 2 2 < 25 < 25 < 25 < 25 < 25 < 25 < 25 N N N N N N N N N N 8.52 7.83 3 2 < 25 25-50 N N N N 5 2 4 5 5 3 3 2 4 1 2 4 7.30 7.13 4.49 3.27 3.68 3.43 5.10 5.74 10.00 6.79 5.88 2 3 2 2 1 3 < 25 < 25 < 25 < 25 < 25 < 25 < 25 25-50 < 25 Y N Y Y N N Y 3 < 25 N 1 1 3 2 10.81 11.99 6.63 5.06 2 1 2 2 25-50 25-50 < 25 < 25 N N N N 2 1 Y Y Y N N N N N N N N N N N N N N N 459 #NFS mm/scar #Plat %Unmod Bipolar ED N N N N 5.3 9.2 5.75 2.52 2.53 2.83 4.5 2.8 3.33 3.01 4.29 3.73 N N Y N N N N N Y N N 7 4 2 3 3 2 3 4 3 4 4 2.00 6.88 6.68 3.94 5.06 6.66 5.51 3.35 4.96 3.69 4.64 3 3 1 3 2 1 2 2 2 3 2 < 25 < 25 < 25 < 25 < 25 < 25 < 25 < 25 < 25 < 25 < 25 Y N N N N N N N N N N N N N N N N N Y Y Y N N 4.06 3.63 3.38 4.06 Y N N N 3 2 7 4 5.64 5.24 1.68 4.38 1 1 4 2 < 25 < 25 < 25 < 25 N N N N N N Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz N N N N N N N N 0.2 0.25 0.3 0.3 0.4 0.4 0.6 0.7 Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. 11.21 10.67 11.36 11.69 14.38 11.67 12.98 13.16 6.77 8.38 7.9 6.07 8.12 6.92 10.78 12.89 2.58 1.38 2.28 3.55 2.01 3.27 3.72 3.57 CAS CAS A A Qtz Qtz N N 0.85 1.15 Ort. Ort. 20.47 11.26 9.25 12.6 4.71 5.02 CAS CAS CAS CAS CAS CAS CAS CAS CAS CAS CAS CAS A A A B B B B B B B B B Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz N Y N Y N N N N Y N N Y 1.5 1.85 2.2 0.35 0.5 0.5 0.6 0.6 0.6 0.65 0.7 0.75 Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. 16.35 13.99 27.53 13.36 11.82 15.17 13.31 16.54 13.4 14.87 14.76 18.56 13.85 9.91 20.11 8.13 11.85 10.72 7.24 12.77 11.11 11.84 12.86 7.97 CAS CAS CAS CAS B B B B Qtz Qtz Qtz Qtz N Y Y N 0.8 0.8 0.85 0.9 Ort. Ort. Ort. Ort. 16.91 10.47 11.75 17.51 8.54 11.96 16.05 10.8 Bk type < 25 < 25 5 5 A A A A A A Bk 2 2 T (mm) 6.82 5.63 W (mm) 3 2 L (mm) N N Dims. N Y Y N Y N N N Wt (g) N N N N N N N N HQ < 25 < 25 < 25 < 25 < 25 < 25 < 25 < 25 RM 2 2 2 2 3 2 1 2 XU/Cl. 3.74 3.56 5.68 2.92 3.60 3.89 4.33 6.58 Pit/Sc. 3 3 2 4 4 3 3 2 TH1 TH1 CAS CAS CAS CAS CAS CAS Y N Trv. Y N N Trv. Trv. Trv. Trv. 460 T (mm) 5.15 3.31 5.16 5.18 7.11 6.01 6.08 6.01 CAS CAS B B Qtz Qtz N N 3 3.55 Ort. Ort. 21.44 17.35 15.87 18.13 CAS CAS CAS CAS CAS CAS CAS CAS CAS CAS CAS CAS B B B C C C C C C C C C Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz N N N N N Y N Y N N N Y 4.25 4.3 5.15 0.2 0.4 0.65 1.2 1.2 2.3 2.6 3.3 4.15 Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Max. Ort. 20.24 22.08 20.04 10.22 10.14 12.26 14.72 16.66 17.43 25.43 22.91 23.32 CAS CAS CAS CAS C C C C Qtz Qtz Sil. Sil. N N N N 4.25 4.9 0.4 10.75 Max. Ort. Ort. Ort. 24.09 26.74 11.47 35.6 ED W (mm) 15.76 9.62 15.21 10.6 12.38 15.18 15.56 17.83 Bipolar L (mm) 16.21 19.33 10.74 18.27 15.86 11.91 18.69 19.4 %Unmod Dims. Max. Ort. Ort. Ort. Ort. Ort. Ort. Ort. #Plat Wt (g) 1 1.05 1.2 1.4 1.55 1.8 2.3 2.35 2 2 3 3 4 1 4 < 25 < 25 < 25 < 25 < 25 < 25 < 25 < 25 N N N N N N N N Y N N N N N N N 5.36 2.48 2 3 < 25 < 25 N N Y Y 5 3 5 2 4 4 4.05 7.36 4.01 5.11 2.54 3.07 3 2 2 1 2 2 < 25 < 25 < 25 < 25 < 25 < 25 Y N 6 5 2 2.78 3.49 12.72 3 2 1 < 25 < 25 < 25 N Y N N 2 11.66 2 25-50 N N N N Y N N N N N Y Y N N N Y N 3 7 1 5 3 1 2 < 25 < 25 < 25 < 25 N 3.82 11.47 7.12 mm/scar HQ N Y N Y N N N N #NFS RM Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Bk type XU/Cl. B B B B B B B B Bk Pit/Sc. CAS CAS CAS CAS CAS CAS CAS CAS N N N N N N N 4 3 3 7 5 4 3 5 6.44 3.58 2.61 3.17 2.98 6.23 3.88 8.87 7.51 N N 4 7 16.35 17.41 19.33 6.92 6.85 11.29 13.44 12.32 6.81 15.63 13.82 18.08 8.43 8.2 7.91 1.79 3.46 4.64 4.54 4.44 13.32 3.82 7.97 7.37 N N N N N N N N Y 19.9 10.49 13.27 18.74 8.45 13.53 3.15 10.81 Trv. Mrg. N N N Y N N N N 461 Dims. L (mm) W (mm) T (mm) Bk #NFS mm/scar 0.2 0.45 0.55 0.95 1.7 2.2 2.35 2.7 Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. 10.41 14.53 12.67 13.45 14.15 16.49 22.48 18.31 7.03 10.61 13.39 12.79 11.75 10.53 12.01 18.37 2.38 2.23 1.92 3.51 7.1 7.92 3.6 6.43 N N N N N N N N 2 2 2 2 5.21 7.27 6.34 6.73 1 1 1 < 25 < 25 < 25 < 25 5 3 3 3.30 7.49 6.10 2 2 2 CAS CAS D D Qtz Qtz N N 2.75 4.5 Ort. Ort. 20.08 33.57 15.54 12.36 7.7 10.12 N N 5 3 4.02 11.19 4 2 CAS CAS CAS CAS CAS CAS CAS CAS CAS CAS CAS CAS D D E E E E F F F Oth. Oth. Oth. Qtz Sil. Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz N N N Y N N N N N N Y N 6.1 5.65 1.55 1.8 3.95 6.25 0.75 0.75 1.85 0.2 0.2 0.25 Max. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Max. Ort. Ort. 34.15 24.84 20.66 25.5 13.89 21.84 14.08 17.34 19.11 10.09 10.92 13.4 16.91 29.3 13.53 8.25 20.11 25.66 14.16 14.45 13.58 6.13 6.71 7.31 10.21 6.46 4.78 6.33 9.13 8.52 3.56 2.12 6.78 2.94 2.26 2.43 N N N N N N N N N Y Y N 4 4 3 4 6 9 2 6.21 6.89 6.38 2.32 2.43 7.04 2 2 2 4 4 1 6.37 2 5.46 6.70 CAS CAS CAS CAS Oth. Oth. Oth. Oth. Qtz Qtz Qtz Qtz Y N N N 0.3 0.3 0.3 0.35 Max. Max. Ort. Ort. 11.46 10.89 13.15 10.91 6.1 8.52 6.5 11.27 4.27 3.54 2.54 2.55 Y N N Trv. 3 3 2 2 6 2 3 3 4.38 3.64 ED Wt (g) N N N N N N N Y Bipolar HQ %Unmod RM Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz #Plat XU/Cl. D D D D D D D D Bk type Pit/Sc. CAS CAS CAS CAS CAS CAS CAS CAS < 25 < 25 < 25 N N N N N N Y Y N Y Y N Y N N N < 25 N N N < 25 < 25 < 25 25-50 < 25 < 25 < 25 N 1 1 25-50 < 25 < 25 < 25 N N Y N N N N N N N N Y N N N N N N N N Y 2 2 < 25 < 25 < 25 < 25 N N N N N N N 462 T (mm) Bk 3.49 3.64 3.39 3.27 3.06 3.49 3.77 3.27 N Y Y N N N N N CAS CAS Oth. Oth. Qtz Qtz N N 0.6 0.6 Max. Ort. 16.63 12.67 7.69 8.55 4.82 5.63 CAS CAS CAS CAS CAS CAS CAS CAS CAS CAS CAS CAS Oth. Oth. Oth. Oth. Oth. Oth. Oth. Oth. Oth. Oth. Oth. Oth. Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz N N N N N Y N N N N N N 0.6 0.65 0.7 0.75 0.8 0.9 0.9 0.9 0.9 0.95 0.95 0.95 Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. 15.8 11.39 12.63 11.85 24.16 13.58 15.26 11.36 11.68 17.06 16.45 15.1 7.73 8.93 14.45 12.35 15.51 12.17 14.34 12.2 10.67 10.26 10.9 11.37 CAS CAS CAS CAS Oth. Oth. Oth. Oth. Qtz Qtz Qtz Qtz N N N N 1 1.1 1.1 1.1 Ort. Ort. Ort. Ort. 12.99 13.11 21.02 16.24 12.31 9.35 13.12 13.7 Trv. 3 1 4 2 2 3 6 5 4.06 2 3.38 5.84 8.00 5.23 2.40 2.79 2 2 2 2 4 3 < 25 < 25 < 25 > 50-75 < 25 < 25 < 25 < 25 N N ED W (mm) 6.83 8.1 7.06 8.27 8.95 8.04 6.98 8.67 Bipolar L (mm) 12.19 14.32 13.51 11.68 16 15.7 14.41 13.93 %Unmod Dims. Ort. Max. Ort. Ort. Ort. Ort. Ort. Ort. #Plat Wt (g) 0.35 0.4 0.45 0.5 0.5 0.5 0.55 0.55 mm/scar HQ Y Y N N N Y Y Y #NFS RM Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Bk type XU/Cl. Oth. Oth. Oth. Oth. Oth. Oth. Oth. Oth. Pit/Sc. CAS CAS CAS CAS CAS CAS CAS CAS N N N N N N N N Y N N N N N N N N 1 3 4.22 1 > 50-75 < 25 4.02 4.58 4.24 4.45 3.24 3.1 3.14 5.15 6.04 3.86 3.86 5.41 N N N N N N N N N Y N N 2 3 4 3 3 4 3 2 3 3 3 3 7.90 3.80 3.16 3.95 8.05 3.40 5.09 5.68 3.89 5.69 5.48 5.03 1 2 3 2 2 2 4 1 2 2 3 2 < 25 < 25 < 25 < 25 < 25 < 25 < 25 < 25 < 25 < 25 < 25 < 25 N N N N N N N Y Y N N Y N N Y N 3.7 6.5 3.57 2.03 N N N N 2 3 4 3 6.50 4.37 5.26 5.41 2 1 3 2 < 25 < 25 < 25 < 25 N N N N N N Y N Trv. N N N N N N N 463 Bipolar 1.3 1.4 1.45 1.7 1.75 1.9 2.25 2.4 Ort. Ort. Ort. Ort. Ort. Ort. Ort. Max. 16.4 23.81 19.48 18.24 22.7 16.59 23.42 20.5 13.3 10.83 10.59 16.93 12.82 13.99 16.97 17.01 4.51 5.08 5.98 4.32 6.62 6.32 5.08 7.3 N N N N N N N N 4 3 3 3 3 4 3 4 4.10 7.94 6.49 6.08 7.57 4.15 7.81 2 2 3 2 2 3 2 CAS CAS Oth. Oth. Qtz Qtz Y Y 2.4 2.65 Ort. Ort. 20.04 24.62 12.13 10.45 5.14 5.73 N N 5 3 4.01 8.21 CAS CAS CAS CAS CAS CAS CA1 CA1 CA1 CA1 CA1 CA1 Oth. Oth. Oth. Oth. Oth. Oth. 1 1 2 2 2 2 Qtz Qtz Qtz Sil. Sil. Sil. Qtz Qtz Qtz Qtz Qtz Qtz Y N N N N N Y Y Y N Y Y 3.8 7.05 12.55 0.85 3.25 5.05 0.2 0.4 0.15 0.15 0.25 0.45 Ort. Ort. Ort. Ort. Ort. Ort. Max. Ort. Ort. Max. Max. Max. 22.52 26.2 33.6 14.29 19.77 30.99 10.6 14.69 10.28 10.16 10.73 11.8 25.26 30.6 24.38 14.3 23.85 20.74 7.83 7.79 9.7 7.18 7.91 11.77 5.79 7.55 14.96 3.5 5.39 7.55 2.42 2.69 1.75 2.7 3.88 3.8 N N N N Y N 5 5 3 2 6 5 2 4 CA1 CA1 CA1 CA1 2 2 2 3 Qtz Qtz Qtz BIF Y Y N Y 1.1 1.35 1.9 0.35 Ort. Ort. Ort. Ort. 14.71 15.04 16.79 12.7 13.05 11.99 13.72 6.95 4.27 4.72 6.23 2.57 N N N N N N Y Trv. #NFS Y N N N N Y N N Bk type Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Pit/Sc. Oth. Oth. Oth. Oth. Oth. Oth. Oth. Oth. N 2 3 < 25 < 25 N < 25 < 25 < 25 N Y Y N Y 3 5 3 5.01 3.36 4.23 3 1 ED %Unmod 2 N N N N N N N N Y Y N N #Plat N N N N N N 3.67 < 25 < 25 < 25 < 25 < 25 < 25 < 25 < 25 mm/scar 3 3 2 1 2 4 Bk 4.50 5.24 11.20 7.15 3.30 6.20 T (mm) N N W (mm) N N L (mm) < 25 < 25 Dims. 3 2 Wt (g) N N N N N HQ < 25 < 25 < 25 < 25 25-50 N N N N N N Y N RM N N XU/Cl. < 25 < 25 CAS CAS CAS CAS CAS CAS CAS CAS N N N N 464 5.52 2 N N N N 4 3 3 3 2.70 4.07 4.16 5.02 2 2 2 2 Ort. Ort. Ort. Max. Max. Max. Ort. Ort. 11.19 10.28 10.17 10.44 11.9 10.63 11.52 10.94 4.02 4.6 4.86 4.63 4.56 6.44 4.52 7.26 0.96 2.59 1.61 2.2 2.52 1.36 2.44 2.05 N N Y Y Y Y N N CA1 CA1 3 3 Qtz Qtz N Y 0.3 0.35 Max. Ort. 11.24 14.29 7.74 5.86 4.42 3.79 Y N 3 3 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 3 3 3 3 3 3 3 3 3 3 4 4 Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz N Y Y N Y N Y Y Y N Y Y 0.35 0.5 0.6 0.75 1 1.3 1.35 1.85 2.55 4.05 0.1 0.15 Max. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Max. Ort. 10.93 12.48 12.18 16.99 19.02 11.97 13.68 20.97 16.72 25.79 11.96 11.04 9.84 10.02 9.51 11.52 10.85 13.37 13.86 10.5 15.77 18.01 4.67 6.52 4.09 3.71 3.93 2.71 3.09 5.12 5.07 4.08 6.55 8.41 2.48 2.47 N N N N N N N N N CA1 CA1 CA1 CA1 4 4 4 4 Qtz Qtz Qtz Qtz N N N N 0.2 0.2 0.25 0.3 Ort. Ort. Ort. Ort. 10.78 12.22 12.48 15.05 8.58 7.96 7.39 5.02 2.44 1.85 2.22 3.57 Trv. #NFS 0.05 0.1 0.1 0.1 0.1 0.1 0.15 0.15 Bk type Y N Y N Y Y Y Y 2 3 2 3 4 1 1 2 N N N N N ED 2 Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Bipolar #Plat N 3 3 3 3 3 3 3 3 Pit/Sc. %Unmod mm/scar 3 1 2 2 < 25 < 25 < 25 < 25 < 25 < 25 < 25 < 25 < 25 < 25 Bk 2 3 1 2 T (mm) 6.24 3.05 5.66 6.34 3.99 4.56 6.99 2.39 8.60 W (mm) 5 2 4 3 3 3 3 3 7 3 L (mm) 2 Dims. 4.76 < 25 < 25 Wt (g) 2 2 HQ 11.52 5.47 < 25 < 25 < 25 < 25 < 25 < 25 25-50 < 25 RM 1 2 1 XU/Cl. 5.60 3.43 5.09 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 N N N N N N N N N Y N Y N N N N N N N N < 25 N N N N N N N N N Y N N < 25 < 25 < 25 < 25 N N N N N Y N N N N N N 465 T (mm) Bk 2.26 2.36 3.82 3.8 2.27 3.45 1.77 3.24 Y N N N Y N N N CA1 CA1 4 4 Qtz Qtz Y N 0.75 0.85 Max. Ort. 15.81 16.92 12.15 12.26 3.52 3.22 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 4 4 4 4 4 4 4 4 5 5 5 5 Qtz Qtz Qtz Qtz Qtz Qtz Qtz Sil. BIF Qtz Qtz Qtz N Y N N N N N N Y Y N N 1 1.1 1.15 1.75 2.75 3.8 4.45 0.3 0.45 0.1 0.15 0.35 Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. 11.84 15.12 14.81 18.04 18.28 18.87 18.25 10.61 13.55 10.58 10.39 10.3 10.93 10.79 15.08 18.25 15.98 14.53 19.38 10.8 8.29 5.38 5.38 8.4 CA1 CA1 CA1 CA1 5 5 5 5 Qtz Qtz Qtz Qtz N N N N 0.35 0.4 0.45 0.45 Ort. Ort. Ort. Ort. 11.09 13.57 13.99 15.06 7.09 7.89 9.39 8.95 Mrg. Trv. 2 2 4 3 2 4 1 3 5.24 10.81 2.73 3.78 6.20 2.86 15.62 5.58 2 1 2 2 2 2 1 1 < 25 < 25 < 25 < 25 < 25 < 25 < 25 < 25 N 2 2 8.46 1 5.39 5.71 4.76 5.58 6.5 9.83 9.28 2.55 2.15 1.27 1.75 3.26 N N N N N N N N N Y N N 7 5 3 3 6 4 4 2 2 2 2 1 1.69 3.02 4.94 6.01 3.05 4.72 4.56 5.31 6.78 5.29 5.20 10.30 2 4 3 3 3 2 2 3 3.26 4.06 3.13 2.61 N N N N 5 2 4 2 2.22 6.79 3.50 7.53 3 1 2 2 Trv. 1 1 1 ED W (mm) 8.1 7.04 7.46 8.41 8.04 8.55 9.43 7.89 Bipolar L (mm) 10.47 21.62 10.9 11.35 12.4 11.42 15.62 16.75 %Unmod Dims. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. #Plat Wt (g) 0.3 0.35 0.35 0.35 0.35 0.4 0.4 0.55 mm/scar HQ N N Y N N Y N Y #NFS RM Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Bk type XU/Cl. 4 4 4 4 4 4 4 4 Pit/Sc. CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 N N < 25 < 25 N N N < 25 < 25 < 25 < 25 < 25 < 25 25-50 < 25 < 25 < 25 < 25 Y N N Y Y Y N N N N N N N Y N N N N N N Y N N Y N Y N N N N N < 25 < 25 < 25 < 25 N N N N N N N N N N N 466 N N N Y N N N N N N N N 2 6 2 4 3 4 2 4 3 6 4 5 8.45 3.32 5.08 3.53 3.79 2.97 5.72 3.39 3.95 1.97 3.42 5.46 2 25-50 < 25 2 1 < 25 < 25 1 3 3 2 2 3 < 25 < 25 < 25 < 25 N Y N N N N N N N Y N N N N N Y N N N Y N N N N 5.82 6.29 2.3 2.34 N N 3 4 2 2 4.32 4.65 2 2 < 25 < 25 < 25 7.60 2 Y Y N N N N N N Ort. Ort. Ort. Ort. Ort. Ort. Ort. Max. 11.82 12.4 16.6 19.29 18.9 11.15 10.81 11.02 6.04 10.57 9.71 9.53 16.02 8.86 4.41 6.42 4.06 2.89 3.75 6.12 7.5 1.3 1.35 3.05 N Y N N N N N Y CA1 CA1 6 6 Qtz Qtz N N 0.3 0.35 Ort. Ort. 10.52 11.64 6.49 7.8 2.83 2.65 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 6 6 7 7 7 7 7 7 7 7 7 7 Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz N N N N N N N N N N N N 0.75 6.45 0.2 0.25 0.3 0.45 0.55 0.65 0.75 1.1 1.4 3.4 Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. 16.89 19.94 10.15 14.1 11.36 11.89 11.44 13.55 11.85 11.8 13.69 27.32 7.82 17.12 4.92 6.72 7.05 7.95 8.72 8.45 12.52 9.37 9.15 14.08 CA1 CA1 CA2 CA2 8 8 1 1 Qtz Qtz Qtz Qtz N N N Y 1.3 1.55 0.2 0.35 Ort. Ort. Max. Ort. 12.95 18.6 10.54 15.2 11.37 8.26 7.6 8.48 N Trv. Trv. < 25 ED 4.34 10.05 2.48 3.01 2.65 4.21 5.09 4.72 3.5 5.54 7.25 7 0.5 0.55 0.75 1.4 2.05 0.15 0.1 0.2 Bipolar #Plat N N N N N N N N N N Pit/Sc. %Unmod mm/scar N Qtz Qtz Qtz Qtz Qtz Sil. Qtz Qtz #NFS > 50-75 < 25 5 5 5 5 5 5 6 6 Bk type 1 1 Bk 10.52 3.88 T (mm) 1 3 W (mm) N L (mm) N N N N N N N N Dims. N N N Y N N Y Wt (g) < 25 < 25 < 25 < 25 > 50-75 < 25 < 25 < 25 HQ 1 3 2 3 1 2 2 RM 5.91 4.13 5.53 3.22 9.45 5.58 3.60 XU/Cl. 2 3 3 6 2 2 3 2 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 467 N N Y CA2 CA2 3 3 Qtz Qtz N N 0.85 4 Ort. Ort. 13.32 18.28 10.51 15.02 7.03 9.9 CA2 CA2 CA2 CA2 CA2 CA2 CA2 CA2 CA2 CA2 CA2 CA2 4 4 4 4 4 4 4 4 4 4 4 4 Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Y N Y Y N N Y Y N N Y Y 0.15 0.2 0.2 0.25 0.3 0.3 0.4 0.5 0.6 0.65 4.75 < 0.05 Max. Ort. Max. Max. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. 10.22 11.28 12.36 11.11 12.34 10.61 10.78 11.72 19.11 10.84 24.56 11.09 5.83 6.51 11.85 6.94 6.22 12.5 9.81 9.49 7.83 13.06 22.28 2.29 2.48 2.4 1.67 3.5 2.68 2.42 3.86 3.15 2.9 4.02 7.36 0.82 CA2 CA2 CA2 CA2 5 5 5 5 Qtz Qtz Qtz Qtz N N N N 0.2 0.25 0.3 1.85 Max. Ort. Ort. Ort. 10.99 11.36 13.8 18.57 4.98 7.07 5.62 13.04 3.34 4.14 4.43 5.42 N Y N N 4 2 3 2 4.09 5.48 2 < 25 < 25 < 25 Y N N 2 3 2 6.66 3.82 6.11 1 2 1 25-50 < 25 < 25 N N N N N N 3 6 4.44 3.05 2 3 < 25 < 25 Y Y Y N Y Y 2 2 3 4 1 3 5 3 3 2 4 2 N N 12.34 3.54 2.16 3.91 6.37 5.42 6.14 5.55 < 25 < 25 < 25 < 25 < 25 < 25 < 25 < 25 < 25 < 25 < 25 2 4 4 4 2.84 3.45 4.64 N N N N N N N N N Trv. 5.64 1 3 2 2 2 3 2 3 < 25 < 25 < 25 < 25 ED T (mm) 8.91 0.98 2.1 2.07 2.31 2.67 3.22 2.41 Bipolar W (mm) 25.91 4.21 6.59 6.96 7.47 4.99 6.88 9.81 %Unmod L (mm) 16.34 10.96 11.2 11.93 10.93 13.31 11.46 12.22 #Plat Dims. Ort. Ort. Max. Max. Ort. Ort. Ort. Ort. mm/scar Wt (g) 5.1 < 0.05 0.1 0.15 0.2 0.2 0.25 0.35 #NFS HQ N Y Y Y N Y Y N Bk type RM Sil. BIF Qtz Qtz Qtz Qtz Qtz Qtz Bk XU/Cl. 1 2 3 3 3 3 3 3 Pit/Sc. CA2 CA2 CA2 CA2 CA2 CA2 CA2 CA2 N N N Y Y N N N Y N N N N N N N N N N N N N N Y Y N N N N N N N N Y N Y 468 Dims. L (mm) W (mm) T (mm) Ret. Y N Y N Y Y N N N N N N Y 0.75 0.3 1.2 0.25 1.3 0.95 2.5 0.1 0.1 0.3 0.1 0.2 0.05 Max. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. 16.42 11.64 12.32 12.2 26.37 12.73 18.17 7.28 5.68 11.78 7.37 8.6 6.13 9.24 7.78 12.57 6.78 12.99 12.93 17.22 3.42 4.69 5.88 4.86 6.28 3.48 4.47 3.98 5.89 2.32 2.52 5.78 5.98 1.67 2.7 4.63 2.35 2.81 1.85 Bck. Bck. Bck. Bck. Uni. Uni. Uni. Bck. Bck. Uni. Bck. Bck. Bck. N N N TH1 TH1 TH1 TH1 TH1 TH1 TH1 TH1 TH1 TH1 TH1 3 3 4 4 4 5 5 5 6 6 6 Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Y Y N N N Y N N N N N 0.1 < 0.05 0.1 0.15 0.2 0.05 0.1 < 0.05 0.1 0.1 0.2 Ort. Max. Max. Ort. Ort. Ort. 7.46 5.12 6.62 9.55 6.37 5.51 5.6 3.64 4.83 5.52 10.44 4.16 1.89 1.84 2.87 2.28 3.08 2.2 N Y Y Ort. Ort. Ort. Ort. 4.95 7.24 5.96 8.44 3.99 5.96 4.5 5.45 2.03 1.9 2.5 2.71 Bck. Bck. Bck. Bck. Uni. Bck. Bck. Bck. Bck. Bck. Bck. N N N Y Y Y N Y N Y Y Y Y Y Y ED Wt (g) Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Bk type HQ 9 3 3 4 4 1 1 1 2 2 3 Bk RM CS MC1 MC1 MC1 MC1 HS HS TH1 TH1 TH1 TH1 TH1 TH1 Pit/Sc. XU/Cl. G.6 RETOUCHED FLAKES N N Trv. Trv. Trv. Trv. Trv. Trv. Trv. Trv. Trv. N N N Y N N N N N N N N N N N N N N N N N 469 Wt (g) Dims. L (mm) W (mm) T (mm) Ret. 0.2 0.35 0.6 0.7 1.1 0.5 0.25 53.65 Ort. Ort. Ort. Ort. Ort. Ort. Max. Ort. 8.77 13.36 20.13 17.93 16.04 13.47 12.96 53.7 7.94 9.48 7.56 8.77 12.93 9.99 9.42 43.42 2.94 2.35 3.09 3.29 3.57 2.21 2.74 16.33 Bck. Bck. Bck. Bck. Bck. Bck. Bck. Uni. N N N N N N N N Y CAS CAS Oth. Oth. Qtz Qtz Y Y 0.15 0.45 Ort. Ort. 12.3 12.84 5.31 8.04 1.95 2.61 Bck. Bck. N N N Y CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA1 CA2 CA2 1 1 3 3 4 4 4 4 4 5 1 2 Chal. Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Y Y N Y Y Y N Y N Y Y Y 0.1 0.1 0.6 0.9 0.05 0.1 0.15 0.4 0.5 0.1 0.3 0.1 Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. Ort. 7.95 11.55 17.04 24.95 11 10.03 15.11 14.66 12.47 9.09 10.8 12.82 6.35 4.76 8.91 7.52 3.36 6.91 4.74 6.96 9.83 4.95 10.35 3.09 1.91 2.05 2.93 4.86 1.91 1.14 1.43 2.32 2.94 2.14 2.23 2.66 Uni. Uni. Bck. Bck. Bck. Bck. Bck. Bck. Bck. Bck. Uni. Bck. N N N N Y N N N N Y N N N N N N N Y N N N N N CA2 CA2 CA2 CA2 2 2 2 2 Qtz Qtz Qtz Qtz Y Y N Y 0.15 0.15 0.15 0.2 Ort. Ort. Ort. Ort. 12.83 8.93 10.86 15.61 4.55 5.91 7.15 4.11 3.39 1.83 1.76 1.83 Bck. Bck. Uni. Bck. N N N N N N N N ED HQ Y Y Y Y Y N Y N Bk type RM Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Bk XU/Cl. A A A A B D E E Pit/Sc. CAS CAS CAS CAS CAS CAS CAS CAS N N N Y Y Trv. Trv. 470 Wt (g) Dims. L (mm) W (mm) T (mm) Ret. 0.4 0.5 0.85 0.65 0.95 0.1 0.3 4.1 Ort. Ort. Ort. Ort. Ort. Max. Max. Ort. 15.13 12.36 19.31 14.73 13.02 7.09 12.26 22.34 7.36 10.76 9.39 8.79 12.61 5.72 6.45 16.63 2.48 2.66 4.54 3.6 4.75 2.08 3.91 8.53 Bck. Uni. Bck. Uni. Uni. Bck. Bck. Uni. N CA2 5 Qtz Y 0.05 Ort. 9.75 3.31 1.7 Bck. N Y N N Y Y N Trv. ED HQ Y Y Y N Y Y Y N Bk type RM Qtz Qtz Qtz Qtz Qtz Qtz Qtz Qtz Bk XU/Cl. 2 2 2 3 3 4 4 4 Pit/Sc. CA2 CA2 CA2 CA2 CA2 CA2 CA2 CA2 N N N N N N Y N N 471