<<

DECLARATIONS

The work presented in this thesis is, to the best of my knowledge and belief, original except as acknowledged in the text. The material has not been previously submitted, either in whole or in part, for a degree at this or any other university.

Jacqueline Matthews

31 May 2013

Brisbane, QLD

I certify that I have read the final draft of this thesis and it is ready for submission in accordance with the thesis requirements as set out in the School of Social Science policy documents.

Ass/Prof. Christopher Clarkson

31 May 2013

Brisbane, QLD

Cover image: Nawarla Gabarnmang, north facing side looking East from above escarpment (photograph Jacqueline Matthews).

I

TABLE OF CONTENTS

List of Figures V List of Tables VII Acknowledgements VIII Abstract IX Chapter 1: Introduction 1 Research Question and Aims 1 Context and Approach 2 Scope 4 Significance of Study 4 Thesis Organisation 5 Chapter 2: Literature Review 6 Introduction 6 Stone Artefacts in Australian 6 Traditions and Sequence Models 7 Questioning Typology 8 Recent Approaches and Gaps in Current Knowledge 8 Gaps and Opportunities 10 Archaeology in Jawoyn Country 11 Nawarla Gabarnmang 11 Expectations for 13 Summary 16 Chapter 3: Methods 17 Introduction 17 Analysis 17 Replicative Experiments 18 Procedures for Analysis 20 The Nawarla Gabarnmang Assemblage 22 Testing Technological Expectations 24 Statistical Analyses 27

II

Summary 27 Chapter 4: Results 28 Introduction 28 Occupational Intensity 28 In Situ Flaking Intensity 29 Raw Material Richness 30 Reduction Intensity 34 Core Preparation and Flake Removal Strategy 37 Objective Piece and Reduction Stage 38 Summary of Results 42 Summary 44 Chapter 5: Discussion 45 Introduction 45 Testing Predictions at Nawarla Gabarnmang 45 45,000–30,000 cal. BP: Initial Occupation pre-LGM 45 (Periods 4 and 5) 21,500–15,200 cal. BP: Last Glacial Maximum (Period 3) 47 14,800–8500 cal. BP: Terminal Pleistocene/Early 49 (Period 2) 2500–250 cal. BP: Late Holocene (Period 1) 51 Summary 53 Readdressing the Research Problem 55 Research Approaches 55 Simple vs Complex Technology 56 and Organic Technology 56 Addressing the Research Problem 57 Conclusion 59 Future Research 60 References 61 Appendix 1: Description of Debitage Attributes 71 Appendix 2: Description of MANA Attributes 73

III

Appendix 3: Square A Artefact Counts 75 Appendix 4: Further Results 78 Appendix 5: Description of MANA Groupings 80 Appendix 6: Analytical Nodules per Spit 85 Appendix 7: Artefact Illustrations 87

IV

LIST OF FIGURES

Figure 1 Location of Jawoyn Lands and Nawarla Gabarnmang in the 2 Northern Territory (Gunn et al. 2012:55). Figure 2 Location of the different types of sites in the Nawarla Gabarnmang 12 site complex, the star marks the locations of the main shelter where excavations took place (Gunn et al. 2012:56). Figure 3 Flow chart demonstrating how MANA divides materials using a 21 simplified example of quartzite from the Nawarla Gabarnmang assemblage; lines that do not extend to a pattern category are unpatterned only. Expanded description of all attributes can be found in Appendix 2. Figure 4 Volumetrically adjusted artefact discard rate for Square A. Refer to 28 Appendix 3 for raw data. Figure 5 Breakdown of the analysed Square A assemblage, comparing the 29 frequency of cores, debitage, retouched artefacts and flake pieces (see Appendix 3 for raw data). Figure 6 Ratio of complete/proximal to broken flakes. 30 Figure 7 Ratio of identifiable (i.e. complete, proximal, and broken flakes) to 30 flake pieces. Figure 8Percentage of each main raw material type across all spits. All 31 complete/proximal flakes, broken flakes, retouched artefacts, and cores are included here. Figure 9 Comparison of local to exotic raw material. 32 Figure 10 Raw material richness based on MANA. 32 Figure 11 Ratio of exotic flakes relative to the number of exotic nodule 33 groups present in each period. Figures 12 and 13 Index of invasiveness results for all retouched flakes (left) 35 and all points (right) in Square A. Figure 14 Box plot showing dorsal scar numbers. 36 Figure 15 Ratio of multiple to single dorsal scar directions. 36 Figure 16 Percentage of different platform preparation types. ‘Other’ refers 37 to platforms that were facetted, ground or a featured a combination of

V techniques, these are presented as one category as they represented such a small proportion of the assemblage. Figure 17 Percentages of different initiation types. 38 Figure 18 Percentage of different platform types; ‘other’ refers to focalised 39 and mixed platform types. Figure 19 Ratio of flakes without bulbs to those with bulbs. 40 Figure 20 Percentage of different bulb types. 40 Figure 21 Mean platform size. 41 Figure 22 Mean flake mass. 41 Figure 23 Mean relative thickness. Relative thickness is calculated as 42 (Length+Width)/Thickness.

VI

LIST OF TABLES

Table 1 Percentage of bending to hertizan initiation types across the four 19 replicated and compared to the archaeological assemblage (note that multiple experiments are pooled here). Statistical tests confirm the validity of these results and show that only the unifacial point assemblage was not significantly different from the Geleji assemblage (unifacial assemblage 1 p > 0.993 and unifacial assemblage 2 p > 0.419). Table 2 Attributes recorded in analysis, see Appendix 1 for expanded 20 description of each individual attribute. Table 3 Division of Square A into analytical periods, based on dates and 23 stratigraphic units, from David et al. (2011). Table 4 Technological tests used to ascribe meaning to the data collected in 25 analysis. A description of each attribute is presented in Appendix 1. Table 5 Counts and percentages of the four main artefact types for the entire 29 Square A assemblage (see Appendix 3). Table 6 Number of analytical nodules for each raw material in the Square A 31 assemblage and indication of which periods they occur in. Table 7 flakes from Square A; note that the flake from Spit 30 was 33 previously described by Geneste et al. 2010. Table 8 Frequency of different retouched artefact types across the different 34 occupation periods and their percentage of the total retouched assemblage. Table 9 The extent of core reduction for each time period containing cores, 35 including the mean mass (g), number of scars >15mm, number of core rotations, frequency of cores and the raw material types used. Table 10 Summary of key findings, interpretations and inferred provisioning 54 strategies for Nawarla Gabarnmang.

VII

ACKNOWLEDGEMENTS

I would like to thank my supervisor, Chris Clarkson, for guiding me through this project and for introducing me to the world of Australian as a first year student. This project is a culmination of many years of teaching and mentoring (not to mention thousands of artefacts) for which I am extremely grateful.

Sincere thanks to Dr Bruno David, the rest of the Connecting Country researchers and the Jawoyn people for allowing me to work on the Nawarla Gabarnmang assemblages and to be a small part of this expansive collaborative project. Thanks especially to the 2012 field crew who I excavated at Gabarnmang with and to Jawoyn Elder and site owner Margaret Katherine for introducing me to this site.

I am particularly grateful to Lynley Wallis for taking the time and care to read over drafts of this thesis, and who has provided me with so much guidance and advice as I find my place in archaeology.

Thanks also to my fabulous friends, little brother, and extended families for their kindness and support, and for doing their utmost to keep me sane for the past year and a half while I worked on this thesis. And finally, heartfelt thanks to my wonderful parents, Colleen and Gerard, for their continued support in all things, for taking a genuine interest in my project, and for not judging when things inevitably became stressful towards the end.

VIII

ABSTRACT

This thesis examines the of Nawarla Gabarnmang, a rockshelter in

Jawoyn Country, Arnhem Land, with evidence for occupation spanning ~45,000 years ago to the present. The colonisation of Sahul was an important event in the story of anatomically and behaviourally modern . Despite the demonstrated behavioural and symbolic complexity of the earliest occupants, traditional approaches to Australian lithic technology have emphasised the static and simplistic nature of Pleistocene technology, contrasting it with an inventive Holocene period. This project provides new evidence on this issue by examining the lithic assemblage from Nawarla Gabarnmang to investigate technological complexity and change in Australian prehistory.

This project utilised debitage attribute analysis coupled with the theoretical framework of technological organisation to explain the nature of variation in the lithic assemblage.

Initial results indicate that the sequence from Nawarla Gabarnmang demonstrates a flexible and adaptive approach to technology throughout time with changing technological strategies emerging in response to variation in environmental contexts.

Overall this thesis demonstrates that Pleistocene technology in Australia was a part of a flexible land-use strategy that solved the problem of making a living in a landscape of changing resource structure. The findings of this thesis emphasise the value of continuing to test previous assumptions about the nature of Pleistocene Aboriginal societies and has provided new evidence that further challenges the notion of a simple two-stage sequence and a unilinear trajectory towards complexity in Australian lithic technology.

IX

CHAPTER 1: INTRODUCTION

This research project examines the lithic technology of Nawarla Gabarnmang, a rockshelter on the southwest Arnhem Land Plateau, in the Northern Territory of Australia. Previous research in this region, and across Australia more broadly, has tended to adhere to a metanarrative of progressive change through time (Przywolnik 2005). This tendency has emphasised the ‘static’ and ‘simplistic’ nature of Pleistocene technology and contrasted it against more dynamic and complex mid- to late Holocene technologies (and, by inference, society) (e.g. Bowler et al. 1970; Mulvaney 1975; Schrire 1982). However, more recent research has suggested that factors such as taphonomy and sample size (Langley et al. 2011), the assumptions of previous researchers (Hiscock 2007; Holdaway 1995), and a focus on antiquity rather than behavioural variation (Frankel 1993; Ulm 2013), could all be influencing interpretations of the nature of human behaviour during the Pleistocene in Australia. Hiscock (2012) has argued that the continued interpretation of Pleistocene lithic technology as being ‘simple’ (e.g. Moore 2013; O'Connell and Allen 2012) is out-dated and at odds with mounting evidence for early complexity. Indeed, recent evidence of early ground-edge axe technology from this site supports an argument for complexity in Pleistocene lithic technology (Geneste et al. 2010, 2012).

Nawarla Gabarnmang is dated to ~45,000 BP (David et al. 2011) and contains evidence for human occupation throughout key periods of change. This site’s age, location in the North of Australia (giving it a prime position to answer questions relating to initial colonisation), and large lithic assemblage, which has been an issue for previous researchers investigating sites of this age (Law et al. 2010; Slack 2007), make it positioned to address questions of technological change over the full sweep of human occupation on this continent.

Research Question and Aims

This research provides a new perspective on Australian lithic technology by examining the lithic assemblage from Nawarla Gabarnmang to determine whether or not it parallels trends identified at other sites across Australia. It allows me to examine the question; how might have changing environmental contexts influenced the technological strategies people used over time at Nawarla Gabarnmang?

1 The Nawarla Gabarnmang lithic assemblage was analysed with several research aims in mind:

 To identify its technological nature and assemblage composition;  To use debitage attribute analysis to identify possible change through time;  To compare all occupation periods from the site to establish whether the directionality of change emphasised by previous research in Australia, is apparent or not; and,  To explain any apparent technological change with reference to hunter-gatherer theory concerned with technological organisation and provisioning strategies.

Context and Approach

Nawarla Gabarnmang is a large open-ended rockshelter located in the traditional lands of the Jawoyn people (Figure 1). This site has been excavated since 2010, and has been demonstrated to be among the few known archaeological sites in Australia dated to or beyond 45,000 years BP (O’Connell and Allen 2012) and further has an archaeological sequence that continues through the Last Glacial Maximum, the Holocene and into the contact period (David et al. 2011). Nawarla Gabarnmang is the first site to be excavated as part of the current archaeological research in Jawoyn lands and already this site has produced significant findings the most striking being the oldest, until recently, known evidence of ground edge axe technology (Geneste et al 2010, 2012).

Figure 1 Location of Jawoyn Lands and Nawarla Gabarnmang in the Northern Territory (Gunn et al. 2012:55). Due to time restrictions this research focuses on the artefact assemblage from one excavated square (Square A) of the site. The assemblage from Square A is principally

2 composed of debitage (98.95%), that is, any flake that has not been utilized or retouched, as such the methods employed are tailored to this artefact category, even though some retouched artefacts and cores are present.

Debitage is often a useful means of inferring the technologies employed at a site when formal components such as retouched artefacts and cores are rare or even absent from assemblages (Shott 1994). Debitage attribute analysis can provide an understanding of the technological activities that occurred at a site, even if there are only small flakes left behind (Andrefsky 2001). In order to understand technology using this approach the researcher infers a relationship between distinctive flake attributes (e.g. platform type) and the technologies or core forms that most likely generated such attributes. These inferences are based on comparison and analogy with indicative experimental and archaeological assemblages.

Another technique applied in this research is minimal analytical nodule analysis (MANA). This technique extrapolates the minimum number of distinctive raw material types flaked at a site by identifying the range of distinctive stone types on the basis of colour and grain size. MANA provides a way to explore past mobility and land-use by determining the different kinds and amounts of local and exotic stone flaked at the site (Hall 2004; White 2012). This technique was coupled with debitage attribute analysis and other measures of reduction intensity, such as core morphology (rotations, number of removals and metrics) and the index of invasiveness for retouched artefacts (Clarkson 2002). Overall this approach provided a sound means of investigating technological activities carried out at Nawarla Gabarnmang, such as artefact reduction strategies and raw material transport.

The research concentrates on building behavioural interpretations concerning the strategic organisation of technology at the site (Nelson 1991). Technology is understood as a series of strategies and decisions that people made to meet their needs and solve the problems they faced. Specifically, the provisioning model of Kuhn (1995) is employed in this study as a means to understand the relationship between subsistence, mobility and technological strategies (Clarkson 2006), and to investigate how these might have changed in relation to variation in climatic and environmental conditions over time (Reeves et al. in press-a).

3 Scope

Ideally this project should have drawn on multiple lines of archaeological evidence (Stiner and Kuhn 1992) and situated this site within a detailed regional study. Unfortunately this was beyond the scope of this project and the current state of regional archaeological knowledge. Additional research at other sites in Jawoyn Country is required to verify whether the trends identified at Nawarla Gabarnmang are specific to this site or whether they hold true for a larger area. Further, as there is currently no local palaeoenvironmental data, nor any faunal or botanical remains from the site, I relied on general palaeoenvironmental data for northern Australia to make broad inferences about how the availability of subsistence resources may have changed through time. This means that while the analytical results are site specific the interpretations of how these findings relate to technological organisation and provisioning strategies are of a much broader scale.

Significance of Study

A technological study of the lithic assemblage from Nawarla Gabarnmang is important in part because of the rarity of Pleistocene sites in Australia (O'Connell and Allen 2012). However, recent research has highlighted the broader significance of Nawarla Gabarnmang in Australian archaeology. This site has yielded some of the oldest evidence for ground- edge axe technology in Australia and buried fragments of rock-art globally (David et al. 2013b; Geneste et al. 2012). Locally this project will be the first study of technology in Jawoyn lands since the 1950s (cf. McCarthy 1951) as such this research will expand knowledge about the prehistory of human occupation in Jawoyn lands.

The methods employed in this project are noteworthy as this form of analysis is not commonly applied in Australia (cf. Mackay 2005). As such, an exploration of the usefulness of debitage attribute analysis to Australian sites has the potential to expand the range of current approaches to lithic technology in Australia. Coupled with a robust theoretical framework, this project has the potential to provide meaningful explanations of past human behaviour at this site.

The significance of developing a stronger understanding of Pleistocene technology using contemporary analytical methods cannot be understated. Recent research has emphasised the global significance of the colonisation of Sahul (Balme 2013; Davidson 2013; Veth et al. 2011b). However, there has been little recent work on Pleistocene lithic technology and

4 notions of its simplicity and uniformity, whilst considered by some to be out-dated (e.g. Hiscock 2012), are still pervasive (e.g. Moore 2013). Therefore a detailed analysis of the Nawarla Gabarnmang assemblage is an important and timely project given the current research interest in life during Pleistocene Australia (Balme and O'Connor in press; Davidson 2013; Langley et al. 2011; O'Connell and Allen 2012).

Thesis Organisation

Chapter Two conceptualises the thesis, identifying current gaps in our knowledge, and presenting testable predictions. The methods employed to answer the research question are discussed in Chapter Three. Results of the analyses and key trends in the data are presented in Chapter Four. Chapter Five presents an interpretation of the results, answering the research question, and commenting on the project’s contribution to the field and avenues for future research.

5 CHAPTER 2: LITERATURE REVIEW

Introduction

There is ample evidence that the first colonists of Sahul employed complex technology (in the form of watercraft and axes) and settled all of the hospitable and inhospitable lands in only a short period of time, suggesting that dynamic adaptability and complexity were features of early Aboriginal societies (Balme and O'Connor in press; Davidson 2013). Despite this evidence, for decades, researchers asserted that lithic technology in Australia lacked complexity and remained largely unchanging until the late Holocene (Moore 2013; Stern 2009). The question is whether the early static nature of technology is a product of an ongoing research focus on formal retouched types and cores (Holdaway and Douglass 2012), which might be undermining attempts to understand technology as being a strategic response to changing contexts. With its large lithic assemblage, position in northern Australia close to potential entry points into Sahul, and appropriate temporal framework, Nawarla Gabarnmang presents an exciting opportunity to explore this question.

In this chapter I survey approaches to the analysis of lithic technology in Australia. I explore the ways in which lithic technology is connected to broader subsistence and land- use strategies, and consider the contemporary approaches to lithic analysis and interpretation as starting points for my research. I provide an examination of the key literature, which provides the foundation for my original research and highlights gaps where there is opportunity for this project to contribute. Finally, I propose a series of expectations for the technological strategies that might have been employed by Aboriginal people using Nawarla Gabarnmang through time and hypothesize how these might be identified.

Stone Artefacts in Australian Prehistory

One of the main preoccupations in Australian archaeology has been explaining variation in artefact assemblages over time and particularly a focus on the differences between Pleistocene and Holocene assemblages. Stone artefact studies in Australia have recently begun to engage with foraging theory and technological organisation as explanatory frameworks (e.g. Clarkson 2007; Hiscock 1994, 2006; Mackay 2005). This change in explanatory focus represents one of the most important advances in Australian artefact

6 studies since the movement against normative typologies, as it allows researchers to embed technology in broader understandings of subsistence strategies and land-use.

Traditions and Sequence Models The earliest archaeological models of Australian lithic technology were developed by Tindale (1957) and McCarthy (1964) who used a typological and culture history approach. Their use of typology and stratigraphic context to understand change was a substantial improvement to understandings of Aboriginal technology at the time (Holdaway and Stern 2004:287). More pervasive than these typological approaches was Mulvaney (1975) who introduced a three-phase sequence, which was later generally simplified into a two-stage sequence model with an early period of generalised adaptive and a later period of technological innovations. Mulvaney’s approach was also based on formal typology but now with some metrical analysis to define categories, which overall made it a more robust approach to understanding assemblage composition and identifying any changes through time (Holdaway 1995:786).

The two-stage sequence was taken up by most of the archaeological discipline in Australia and refined by Jones and colleagues who named the early (Pleistocene) sequence the ‘core and tradition’ (Bowler et al. 1970:52), and which was characterised by the presence of large core tools and steep edge scrapers. The later sequence, the ‘small-tool tradition’ (Gould 1969:233), was identified based on the appearance of small refined artefacts such as points and backed artefacts in the mid-Holocene (Bowdler and O'Connor 1991:47; Mulvaney and Kamminga 1999).

Key sites for establishing the Pleistocene antiquity of occupation in Australia also provided support for the pan-continental nature of this sequence, including at the Arnhem Land sites, Malangangerr, Nawamoyn and Jimeri II (Schrire 1982). These findings were extended to an earlier time period at the nearby site of Nauwalabila I, where Jones and Johnson (1985) argued that there was clear evidence for the two-stage sequence. However, the Nauwalabila sequence included a typologically unspecific assemblage between the early and late traditions (Jones and Johnson 1985:213), which raised questions as to just how clear the distinction between traditions was at this site. Further supporting evidence was found in Western Australia at Devil’s Lair and Miriwun (Dortch 1977; 1984) and in New South Wales at Lake Mungo (Bowler et al. 1970:52), all of which sustained the claim for a pan-continental, two-stage sequence.

7 Questioning Typology Despite some early dissent (e.g. Kamminga and Allen 1973; Pearce and Barbetti 1981), the two-stage sequence model remained a pervasive notion until the late 1980s and early 1990s. At this time researchers began to question the assumptions and evidence underpinning the model, particularly the use of typology and the assertion that there were unilinear continental traditions.

A critical issue surrounding the use of sequence models is that they reinforce the meta- narrative of unidirectional change in hunter-gatherer societies (Przywolnik 2005:177). Sequence models have tended to emphasize technological stasis and uniformity during the Pleistocene and contrasted a perceived lack of change against the more diverse and apparently richer archaeological record of the Holocene (Hiscock 2008:106). This approach has links with nineteenth century stereotypes of Australian Aborigines as being a ‘primitive’ or ‘static’ culture (David 2002:98) and is at odds with archaeological evidence for the complexity of regional variation spatially and temporally (Ulm 2013:185).

A key to the movement away from progressive sequence models was a rigorous reappraisal of functional artefact typologies (Flenniken and White 1985). The central limitation of typology is that as a normative approach it hides technological variation and infers intentionality to establish artefact categories (Hiscock 2007:198, 219). Nevertheless, other researchers have highlighted the potential for typology to be used as an interpretative tool in Australian lithic studies (Holdaway and Stern 2004:313).

Recent Approaches and Gaps in Current Knowledge The current focus on artefact reduction sequences and materialist technological descriptions to characterise assemblages has been key to advances in Australian lithic studies (Flenniken and White 1985; Hiscock 2007; Hiscock and Clarkson 2000). Further, the incorporation of foraging theory as an explanatory tool in Australian archaeology has introduced ideas about toolkit flexibility, the relationship between tool maintenance and mobility, and forager responses to subsistence risk, which are all critical to the way lithic variability is currently interpreted (Hiscock 2008:108).

Technological Organisation Current approaches to the explanation of lithic technology and its variation have drawn on foraging theory and the theoretical framework of technological organisation (Clarkson 2006). These approaches focus on the economics of technology, viewing it as a series of

8 strategies and decisions that people made in order to survive and solve the problems they faced, such as ensuring access to resources when needed and being able to easily maintain or replace tools.

Technological organisation was defined by Nelson (1991:57) as ‘the selection and integration of strategies for making, using, transporting and discarding tools and the materials needed for their manufacture and maintenance.’ Andrefsky (2008:4) framed it somewhat differently as a way to understand how technology is embedded in the daily lives of hunter-gatherers. Both concepts are employed in this research. As a framework, technological organisation has become very popular as it allows the investigation of what technological strategies might have been used, why strategies may have changed through time, and goes some way to explaining the human behaviour and decision making behind them.

An influential set of concepts is provided by Kuhn’s (1995) technological provisioning model. Kuhn focused on the strategies mobile hunter-gatherers use to solve the problem of not having the tools required when needed, and was concerned with factors such as planning, maintenance, and transport. Two strategies for provisioning were introduced by Kuhn (1995:22). The first was provisioning of individuals, where individuals have tools available at all times by carrying a mobile, multipurpose toolkit that allows them to anticipate most needs. The second strategy is a provisioning of places, where tools or material to make them are placed in specific locations where there is anticipated need e.g. where resources are located, favoured camping places etc. These strategies provide a way to understand the relationship between subsistence, mobility and technological strategies (Clarkson 2007:25). Although they seemingly create a binary opposition, these strategies and an evaluation of their relative emphasis over time can be a useful way to explore how technology was organised in the past (Hiscock 2006:84) and, of relevance to this project, how strategies may relate to changing environmental contexts.

Central to Kuhn’s model is an understanding of mobility and the role it can play in structuring hunter-gatherer technology. One of the main ways mobility is identified in lithic technology is through the transport of lithic raw material, understanding the dynamics of stone movement provides a way to understand how wide-ranging people’s mobility might have been (Andrefsky 1994; Kelly 1992; Smith 2011). Further, as material (in raw form or as tools and cores) may need to be reduced before transport (Beck et al.

9 2002), or be used and reduced along the way (Cole 2009:132; Surovell 2009:189), the extent of artefact reduction can inform on the frequency and intensity of mobility. In terms of expectations for technology, increasing mobility should often result in decreased tool and core size (Pecora 2001; Shott 1986:20).

In Australian studies, mobility is commonly linked to concepts of subsistence risk and procurement cost (Hiscock 1994; 2009). Strong evidence for substantial environmental transformations over the course of human occupation in Australia provide the foundations for such models (Hiscock and Wallis 2005). Increased economic risk resulting from the aridity and unpredictability caused by El Nino Southern Oscillation (ENSO) conditions during the late Holocene is often linked to the emergence of new technologies and changing mobility as a risk-minimising strategy (Veth 2005; Veth et al. 2011a). In terms of generating testable expectations for technology, Veth (2005:103–109) argued that the intensity of stone reduction, diversity of artefact assemblages and the varying dominance of local and exotic raw materials are important for understanding potential risk- minimisation and past mobility through lithic technology (see also Clarkson and Wallis 2003; Hiscock 1994).

Application In northern Australia, a key regional study conducted by Clarkson (2007) at several sites across Wardaman Country (Northern Territory) demonstrates the contemporary practice of lithic analysis and interpretation in Australia. This study sought to provide an accurate description and explanation of technological change in this region, through the application of detailed reduction sequence models and foraging theory to demonstrate and explain artefact variation. In contrast to pan-continental models, Clarkson (2007:159) found that there were gradual and continuous technological changes throughout the span of human occupation in this region. The application of reduction sequence models and utilisation of foraging theory and technological organisation in Clarkson’s study is representative of the current approach to lithic analysis and explanation in Australia.

Gaps and Opportunities Recent research has uncovered more sites with long temporal sequences. Unfortunately the lithic assemblages from these are yet to be presented in any detail so the behaviour and activities of people at these sites is largely unknown (e.g. Law et al. 2010; Veth et al. 2009). GRE8, a ~40,000 year old site in the Riversleigh region of northwest Queensland is

10 an exception. It is fully analysed, but unfortunately an extremely small sample size in the oldest layers constrains its ability to meaningfully contribute to our understandings of Pleistocene technology and how it relates to later periods (Slack 2007:227).

It is apparent that there are some key gaps that still need to be addressed, such as to accurately describe the earliest period of occupation, to focus on behaviour not antiquity (David 2002:150; Frankel 1995), and to look beyond continental narratives to appreciate the complexity of regional variation (Hiscock 2008:128; Przywolnik 2005; Ulm 2013). The application of contemporary methods and the use of technological organisation as an interpretative framework at Nawarla Gabarnmang will provide the opportunity to address these problems and generate testable predictions for technological organisation strategies.

Archaeology in Jawoyn Country

The traditional lands of the Jawoyn people cover a ~50,000 km2 area of the southwest Arnhem Land Plateau, encompassing Nitmiluk National Park, the southern part of Kakadu National Park, and the Katherine region. Owing to the inaccessibility of much of Jawoyn Country, little archaeological work had been conducted here prior to the current project. Two other sites had previously been excavated in this region: Tangtangjal (Macintosh 1951) and Sleisbeck Cave (Mountford 1958). While the large Tangtangjal lithic assemblage was analysed in detail by McCarthy (1951), without any dates from the site it is difficult to compare it with the Nawarla Gabarnmang sequence. As only initial observations about the assemblage and excavation of Sleisbeck Cave have been published (e.g. Mulvaney and Kamminga 1999:239); it is not a reliable source of information about technology or behaviour in Jawoyn lands.

Nawarla Gabarnmang As shown in Figure 2, Nawarla Gabarnmang is part of a larger site complex comprising 12 smaller shelters, three areas of grinding patches, four standing stones and a ritual Dreaming stone (Gunn et al. 2012:55). Jawoyn Elders identified this site as belonging to the estate of the Buyhmi clan, for whom it was an important camping place for non- Jawoyn people travelling into this country for ceremony (David et al. 2011:74). Excavations began at Nawarla Gabarnmang in 2010 and have already produced many noteworthy findings, including some of the oldest dates for ground-edge axe technology and use specifically for art (David et al. 2013b; Geneste et al. 2010).

11

Figure 2 Location of the different types of sites in the Nawarla Gabarnmang site complex, the star marks the locations of the main shelter where excavations took place (Gunn et al. 2012:56). Gunn et al. (2012:56) emphasised that the site’s proximity to reliable water from an adjacent creek and the wide range of foods available at a nearby swamp make this a well- situated campsite. Pollen work is currently being carried out in this region, which should indicate how far back the contemporary environmental situation extends. There are limited palaeoenvironmental studies for northern Australia (e.g. Reeves et al. in press-b; Reeves et al. 2008; Shulmeister and Lees 1995), these provide a general view of how the local environment and resource structure may have been affected by climatic change over time.

Key periods of time represented in the Nawarla Gabarnmang Square A sequence are the initial colonisation of Sahul (45,000–30,000 BP), the Last Glacial Maximum (LGM) (21,000–15,200 BP), the terminal Pleistocene/early Holocene (14,800-8500 BP), and the late Holocene (2500–250 BP) (David et al. 2011). These dates largely overlap with those from other squares excavated at the site, which are not yet reported in press. For each of these periods broad environmental trends have been identified. These include high productivity and an active monsoon-influenced climate during initial occupation, intense aridity and variability during the LGM, climatic amelioration following the LGM and into the early Holocene, and ENSO-influenced variability in the mid- to late Holocene (Reeves et al. in press-b). These trends provide an environmental context that hunter-gatherers likely encountered in this region and allow predictions to be made regarding technological organisation over time, which are described in more detail in the next section.

12 Expectations for Technology It has been argued that variability in climate, environment, and resource availability can be proximate causes for alterations observed in technology through time (Chappell 2001; O'Connell and Allen 1995:855). Specifically in northern Australia, the predictability and magnitude of the summer monsoon has been argued to be critical to variability in hunter- gatherer adaptations over time (Kershaw 1995; Morwood and Hobbs 1995:748; Schrire 1972). Whilst technology can, and often does, play a more dynamic role beyond economics it is often outside the scope of research projects such as this to consider both the economic and social aspects of technology (cf. Porr 2000; White 2011).

45,000–30,000 cal. BP, Initial Occupation During the Pleistocene: A Period of Humid and Wet Monsoon-Influenced Conditions The initial period of site occupation occurs from ~45,000 cal. BP and continues through to ~30,000 cal. BP. Palaeoenvironmental data indicates that this was a relatively cool humid period in northern Australia (Fitzsimmons et al. in press; Reeves et al. in press-a:4) with high levels of effective precipitation (Nanson et al. 1992:793; Reeves et al. 2008:17; Ward et al. 2005:1908). An active summer monsoon has also been argued to be in effect at this time (Johnson et al. 1999:1150). It is thus reasonable to expect that conditions during this period would have been ideal for hunter-gatherers in the region (Hiscock and Wallis 2005; Morwood and Hobbs 1995).

Given that there was likely no substantial risk of failing to acquire resources, mobility can be inferred to have been largely unconstrained (Bamforth and Bleed 1997; Torrence 1989). A provisioning of places strategy, as described by Kuhn (1995:22), may have been a successful way of ensuring access to materials to make tools at areas in the landscape where resources were predictable. If so, we would expect the following in the Nawarla Gabarnmang assemblage:

1. A small percentage of exotic materials and dominance of local stone. Reasonable quality raw material is abundant at the site hence stockpiling or long-distance transportation of exotic raw materials is unlikely (Andrefsky 1994). 2. Production of relatively large flakes with abundant cortex and little evidence for preparation or previous reduction. The abundant supply of local stone means there would have been little need to invest in substantial reduction or preparation of cores for transport to the site (Beck et al. 2002; Hiscock 2009:90–91).

13 21,500–15,200 cal. BP, Last Glacial Maximum: A Period of Hyper-Aridity Following a period of likely abandonment, Nawarla Gabarnmang was reoccupied at the height of LGM aridity, estimated to fall between 22,000 and 19,000 cal. BP (Reeves et al. in press-a:7). Palaeoenvironmental evidence for the LGM indicates that this period featured extreme variability in resource and water availability (De Deckker 2001; Kershaw 1995:666), these conditions would have made it critical for Aboriginal groups to alter their subsistence and technological strategies to alleviate increased risks of not obtaining resources (Hiscock 2006:71). That this site is occupied at this time suggests that it might have offered some benefit to people as a refuge. During this period large areas of the continent have demonstrated abandonment (Hiscock 2008:60; Veth 1993) and areas where occupation continued featured substantial changes such as restricted foraging ranges (Hiscock 2008:59–60; Marwick 2002).

A provisioning of individuals strategy could have alleviated some of the stress from living in a hyper-arid LGM environment. In terms of technology, a strategy involving the use of mobile toolkits would be advantageous, as it would ensure a constant supply of tools so people could take advantage of resources on encounter (Kuhn 1995:22). If so, we would expect the following:

1. A focus on local materials related to increased occupation intensity and use of the site as a refuge, but with some evidence for exotic material use, connected to the use of mobile toolkits. 2. A reduction in flake size and an increase in flake removals and preparation, which would reflect increased reduction and maintenance of tools and cores as a result of increased mobility (Bleed 1986:745; Shott 1986:20).

14,800–8500 cal. BP, Terminal Pleistocene/Early Holocene: A Period of Amelioration and Reestablishment of the Summer Monsoon Following the LGM peak there was again a break in site occupation. Human activity is evident again by ~14,000 cal. BP. Palaeoenvironmental evidence indicates that the reduction of aridity may have been slower in northern Australia compared to the rest of the continent (Kershaw 1995:666) and it is likely that some of the aridity and instability of the LGM continued until ~14,000 cal. BP when the monsoon reinitiated (Reeves et al. in press-a; Wyrwoll and Miller 2001:8). Following the reestablishment of monsoonal conditions, evidence suggests that the terminal Pleistocene and early Holocene were very

14 stable, warm, and wet (Bowler et al. 2001:78; Nanson et al. 1993:298; Reeves et al. 2008; Ward et al. 2005:1908).

The gradual improvement in environmental conditions would have alleviated much of the resource risk associated with the LGM. It is likely that a reversion to the pre-LGM strategy of provisioning places would have been preferential following the reinstatement of monsoonal conditions. If a provisioning of places strategy was employed at this time we would expect the following:

1. A focus on locally abundant raw materials indicative of a decreased need to transport tools or cores. 2. A reducing emphasis on maintenance and extended artefact reduction, which would manifest as relatively larger flakes showing minimal evidence for prior reduction or removals.

2500–250 cal. BP, Late Holocene: A Period of Increasing ENSO-Influence, Variability and Aridity The last period of occupation at the site took place between ~2500 cal. BP and ~250 cal. BP and represents an increase in occupational intensity after lower and more sporadic degrees of occupation during the early and mid-Holocene. It is well-documented that the late Holocene was a period of increased variability and aridity largely driven by ENSO in El Niño mode (Reeves et al. in press-a:9; Shulmeister and Lees 1995). These conditions are linked to a decrease in the influence of the monsoon and in effective precipitation (Fitzsimmons et al. in press:15; Luly 1993:587; McGlone et al. 1992:446; Shulmeister and Lees 1995:14). The present configuration of resources and plentiful water in close proximity to the site was likely established but fluctuated during this period. This variability is likely to have had a substantial impact on the ways in which people organised their technology and was perhaps more influential than the preceding pre- and post-LGM occupation periods, which coincided with relative stability and high or improving productivity (Anderson et al. 2006; Haberle and David 2004).

Many researchers have proposed a risk-minimising strategy for this period, where small, easily transportable and maintainable tools were emphasised (Hiscock 1994; Veth 2005). These propositions fit within the general expectations of a provisioning of individuals strategy (Hiscock 2006:71; Kuhn 1995:22). In terms of expectations for technology this strategy is predicted to be manifest as:

15 1. An increase in exotic raw materials indicating higher mobility and increased impetus to transport materials or tools. 2. A decrease in flake size and an increase in reduction intensity related to a strategy to maintain or extend artefact use-lives (Bleed 1986:741; Hiscock 2006; Shott 1986:20).

Summary

This chapter considered the treatment and interpretation of lithic technology in Australian archaeology, discussed the current explanatory frameworks, and identified a critical gap in our knowledge concerning the technology from the oldest periods of human occupation and how they compare to those of later times. A review of archaeological research in Jawoyn lands highlighted the potential for this region to contribute to our understanding of Aboriginal occupation over a considerable period of time. Finally, it proposed a series of testable hypotheses for technological organisation and strategies given demonstrated environmental variation over time. These hypotheses will be evaluated in the following chapters.

16 CHAPTER 3: METHODS

Introduction

In this chapter I present the methods used to analyse the Nawarla Gabarnmang assemblage. I discuss contemporary approaches to debitage analysis, the rationale behind the methods utilised and protocols for their application. Following a brief description of the site and assemblage, I set out a series of technological tests to address the hypotheses proposed in Chapter Two and present the rationale for their inclusion.

Debitage Analysis

Because debitage is abundant in the archaeological record and is unlikely to have been removed from its place of creation, it has the potential to provide greater context specific information about technology at a site than rarer formal implements or cores (Shott 1994:71; 2004:211). Debitage has been argued to be an accurate proxy for activities as the composition of debitage assemblages and the form of individual flakes can be directly related to tool or core form and reduction stage (Andrefsky 2009:80). Its analysis is distinct from that of tools or cores as it can represent all reduction stages (Crabtree 1982:32; Rinehart 2008b).

Debitage is defined as any flake that has not been used or retouched (Shott 1994:70; Sullivan and Rozen 1985:755; Williams and Andrefsky 2011:871). Some analysts have imposed size thresholds for debitage or excluded larger flakes that have ‘potential’ for use (e.g. Ahler 1989; Sullivan and Rozen 1985). However, imposing size ranges on flakes in the absence of use-wear or other indications of selection for use involves making guesses about the size of ‘useful’ flakes. All flakes larger than the 2.1 mm mesh sieve used in recovery are included in the analysis. This sampling strategy provides a stronger basis for building inferences about past technology because it makes no assumptions about intentionality and is more open to variation across the entire assemblage.

Contemporary debitage analysis takes two forms: aggregate and attribute analysis (Andrefsky 2001). These are not mutually exclusive techniques, although they are often presented this way (Bradbury and Carr 2004).

17 Attribute analysis is a robust method used to examine technology and reduction stage across an assemblage, based on the premise that flake attributes and metrics can be used to infer technological origin (Andrefsky 2001:12; Carr and Bradbury 2001; Sullivan and Rozen 1985:755). Analysts using this method record attributes and metrics on individual flakes that relate to manufacturing techniques and characteristics of core form, usually inferred from archaeological and experimental assemblages of known origin (Shott 1994).

Aggregate analysis (also mass analysis) has been most commonly employed in Australian studies of debitage, often in combination with Sullivan and Rozen’s (1985) breakage typology (e.g. O'Connor 1999; Schrire 1982; Veth 1993). It involves dividing an assemblage into size classes (typically by passing it through nested sieves) and recording metrics for each class (generally counts and mass). The data should then be statistically compared to experimental assemblages to identify the technology and/or reduction stage represented (Ahler 1989:393; Andrefsky 2007). This has rarely been undertaken in Australian studies.

This study used attribute analysis as the main approach. However, a simplified form of aggregate analysis was also used for broken flakes that are not amenable to attribute analysis to provide additional lines of evidence.

Minimal analytical nodule analysis (MANA) was used to subdivide raw material types within an assemblage into discrete nodule groups (Andrefsky 2009:84; Hall 2004), based on the assumption that there is visible variation in nodule appearance between different nodules and sources (White 2012). This technique provides a minimum number of individual nodules that were brought to and/or worked at a site and can be a strong indicator of mobility (Hall 2004). Given that Nawarla Gabarnmang is itself a raw material source, MANA generates a deeper understanding of how much exotic material was brought to the site over time.

Replicative Experiments Experimental studies are well-established as a critical component of debitage analysis. Replicative experiments, whilst constrained by numerous variables, provide a way to establish what attributes or metrical distributions characterise certain reduction techniques and the data obtained from these experiments can subsequently be applied to characterise archaeological assemblages (Pelcin 1997:1108; Rinehart 2008a:64). Whilst replicative experiments are common in debitage studies generally, they have not been routinely

18 undertaken or reported in Australia. Owing to variability in raw materials and differences in technologies it is not appropriate to uncritically apply experimental findings designed for a North American context, where the majority of replicative studies have been performed (e.g. Ahler 1989; Carr and Bradbury 2001; Pelcin 1997; Sullivan and Rozen 1985). It was therefore necessary to perform experiments designed for an Australian setting in order to more accurately interpret the Nawarla Gabarnmang debitage data.

A series of experiments were conducted using high quality quartzite and from northern Australia to replicate bifacial points, unifacial points, multiplatform cores and retouched scraper types. These data were compared to a late Holocene sequence from Geleji, Wardaman Country (Clarkson 2007), to test the validity of this technique using Australian raw materials. The replication of debitage characteristics indicates that initiation types vary substantially between technologies and are a useful means of differentiating each reduction technology (Table 1). There is strong agreement between the interpretations drawn about the Geleji assemblage based on the experimental data and established archaeological knowledge for that period and region (Clarkson 2007; Matthews 2012), further emphasising the validity of this method in an Australian setting.

Table 1 Percentage of bending to hertizan initiation types across the four replicated technologies and compared to the archaeological assemblage (note that multiple experiments are pooled here). Chi square tests confirm the validity of these results and show that only the unifacial point assemblage was not significantly different from the Geleji assemblage (unifacial assemblage 1 p = 0.993 and unifacial assemblage 2 p = 0.419).

Bifacial Unifacial Multiplatform Retouched Geleji Initiation / point point core scraper Technology N % N % N % N % N % Bending 62 32 72 45 14 5 9 17 509 48 Hertzian 134 68 88 55 282 95 45 83 560 52

These experimental data (Matthews 2012) serve as a method of interpretation. They specifically demonstrate the links between key platform characteristics (such as initiation and platform type) and core form. The importance of initiation and other platform attributes has been previously demonstrated (e.g. Pelcin 1997) and is shown here to be applicable to Australian technologies and materials.

19 Procedures for Analysis

Flakes with an intact platform (i.e. all complete flakes and proximal fragments) formed the main analytical unit of study. The attributes selected were drawn from contemporary fracture mechanics and lithic analysis literature (e.g. Cotterell and Kamminga 1987; Macgregor 2005; Pelcin 1997; Rezek et al. 2011). Table 2 presents a description of each attribute group employed; these groups are robust indices of flake size and shape, flaking technique, applied force and platform characteristics.

Table 2 Attributes recorded in analysis, see Appendix 1 for expanded description of each individual attribute.

Characteristic Attributes References Flake mass; percussion length; (Andrefsky 2005a:94, 99, width; thickness; proximal width; 101; Clarkson 2007:32; Size distal width; platform width; Clarkson and David platform thickness; length x width 1995:32; Rezek et al. (overall flake size) 2011; Shott 1994:80) Relative thickness (Clarkson and David ([length+width]/thickness); 1995; Rasic and Shape Elongation (length/width) Andrefsky 2001; Sullivan 2001:200) Initiation type; bulb attributes; flake (Clarkson 2007:32; Flaking technique termination Clarkson and O'Connor and force 2006:164; Macgregor 2005) Platform type; platform preparation; (Clarkson 2007:32; platform angle Clarkson and O'Connor 2006:172; Phagan Platform description 1985:235, 273; Rezek et al. 2011; Whittaker 1994:101) Number of dorsal scars; direction (Andrefsky 2005a:109; Dorsal surface and termination of dorsal scars; Phagan 1985:279) presence of arises

MANA was performed using the same analytical unit—that is, flakes with intact platforms—as attribute analysis. The descriptive attributes recorded facilitate the division of raw materials into individual nodules. The attributes included source location, colour, patterning, texture, grain size and appearance (Hall 2004; White 2012). Figure 3 presents a simplified version of how this technique divides materials across the different descriptive categories to arrive at a number of individual nodules within the assemblage.

20

Raw Source Texture Grain size Colour Pattern Material

Red 0.25mm (fine) Unpatterned Light red Granulite Red spots Light brown 0.5mm Exotic (medium) Pinkish white Unpatterned

White Red lines Quartzite 0.25mm (fine) Unpatterned Very pale brown

Red spots Local White Unpatterned 0.5mm Granulite (medium) Very pale brown Grey bands

White Unpatterned 1mm (coarse) Very pale brown Red mixed

Figure 3 Flow chart demonstrating how MANA divides materials using a simplified example of quartzite from the Nawarla Gabarnmang assemblage; lines that do not extend to a pattern category are unpatterned only. Expanded description of all attributes can be found in Appendix 2. 21 Broken flakes were not analysed using the above methods, but were instead divided by raw material and fragment type following Clarkson and O'Connor (2006:194), and counts and mean flake mass recorded for each group. Flaked pieces, defined as ambiguous cultural fragments of stone, were also recorded in this way. These two lines of evidence provided indices of fragmentation.

Even though this project was explicitly focused on debitage rather than retouched artefacts or cores, such objects do exist in the Nawarla Gabarnmang assemblage. These items were analysed following the methods set out in Clarkson (2007:32–39) in order to record the extent of their reduction, which included analysing core morphology (i.e. number of flake scars, rotations and metrics) and applying the index of invasiveness to retouched artefacts (Clarkson 2002). These items were also included in MANA. These data provide another line of evidence to infer reduction intensity and technological change over time.

The Nawarla Gabarnmang Assemblage

To date over a dozen squares (most 50 x 50 cm) have been excavated at Nawarla Gabarnmang. This analysis is concerned with ‘Square A’, which measures 50 x 60 cm and was excavated to a depth of 66.2 cm (Geneste et al. 2010:66). Square A was the first excavated at the site and the first sorted and ready for analysis, it was also selected for this research because of its long sequence and the opportunity it afforded to consider technology from the very earliest period of occupation (i.e. some other excavation units only contained Holocene materials). Basic artefact and raw material counts have been previously reported for Square A along with details of the stratigraphy and radiocarbon chronology (David et al. 2011).

Similar to many other rockshelter sites across Arnhem Land, the Nawarla Gabarnmang rockshelter itself is a source of raw material, with the ceiling, pillars and surrounding escarpment showing clear evidence for exploitation/modification (Delannoy et al. 2013; Taçon 1991). The quartzite present in the shelter is generally low to medium quality but there are seams of higher quality material (i.e. more silicified). The local quartzite is generally distinctive from the exotic quartzites found in the assemblage; the latter are consistently finer-grained and come in a wide variety of colours. The locations of exotic material sources relative to Nawarla Gabarnmang are currently unknown.

22 As the assemblage from Square A represents a long time span, it was necessary to subdivide it based on the radiocarbon chronology and stratigraphy (Table 3). While, such lumping may hide some critical variation over time, it is best to be cautious, while awaiting the results of detailed sedimentological analysis. Decreased intensity and more sporadic degrees of occupation of between 6000 and 10,000 years occur between each phase (except Periods 2 and 3), enabling clear divisions in most cases.

Table 3 Subdivision of Square A into analytical periods, based on dates and stratigraphic units, from David et al. (2011).

Age Range XU SU 14C Age (years BP) Calibrated Age BP (95%) 1 1 250±30 430-360, 230-170, 190-140, 20 to -10 2 1/2 3 2 142±30 290-0 4 2/3 1: Late Holocene 5 3 1560±30 1530-1380 (N = 4,397) 6 3 2264±30 2350-2290, 2260-2150 7 3 8 3 2165±30 2310-2060 9 3 3117±25 3395-3315, 3310-3260 10 3 2491±25 2720-2465 11 3/4 8616±34 9670-9520 12 3/4 8924±31 10,200-10,110, 10,090-9910 13 3/4 14 3/4 9287±30 10,580-10,380, 10,320-10,300 2: Early 15 3/4 Holocene/Terminal 9670±34 11,200-11,070, 10,960-10,860, 10,850- Pleistocene 16 4 10,800 (N = 1,319) 17 4 18 4 11,907±40 13,910-13,610 19 4 20 4 12,261±43 14,800-14,760, 14,550-13,920 21 4/5 22 4/5 23 4/5 12,599±42 15,200-14,520, 14,300-14,280 3: Terminal 24 4/5 17,722±64 21,480-20,810, 20,700-20,560 Pleistocene/LGM 25 4/5 12,746±43 15,570-14,870, 14,830-14,770 (N = 500) 26 4/5 17,833±77 and 21,550-20,930 and 21,585-21,150 27 4/5 17,931±64 28 5 30,615±309 36,260-35,820, 35,640-34,620 29 5 31,316±618 37,110-34,690 4: Pleistocene 30 5 (N = 485) 30,761±314 and 36,290-35,760, 35,710-34,720 and 31 5 31,063±595 36,710-34,650 32 5 41,680±1532 49,075-43,215 5: Pleistocene 33 5 37,842±917 44,090-41,285 Initial 34 5 41,149±1426 48,400-42,855 (N = 179) 35 5 36 5

23 Testing Technological Expectations

In Chapter Two I proposed a series of hypotheses for the Nawarla Gabarnmang assemblage. Table 4 sets out a range of tests to evaluate those expectations and link the data collected to past technological strategies and site/landscape use. A complete understanding of technological organisation is difficult to infer solely from debitage, and as such many of these tests focus on identifying flaking strategies (Carr and Bradbury 2001).

24 Table 4 Technological tests used to ascribe meaning to the data collected in analysis. A description of each attribute is presented in Appendix 1.

Test Attributes Rationale References  Number of artefacts discarded, volumetrically By using relative numbers of artefacts discarded an (Haberle and David Occupational adjusted over time inference can be made on how intensively a site may 2004:171; Hiscock intensity have been used. Although this could also be attributed to 1981; Pecora 2001; different technology types and reduction stages. Smith 2006)  Comparison of debitage to retouched artefacts A proxy for how much manufacture/maintenance took (Carr and Bradbury and cores place at the site; this is likely to be an over-estimation as 2001; Holdaway and In situ flaking many tools could have been removed for use elsewhere. Douglass 2012:25; intensity  Index of fragmentation Also influenced by reduction stage/juncture. Holdaway et al. 2004; Pecora 2001)  Index of invasiveness recorded on retouched These attributes provide a strong indication of how (Andrefsky 2009:743; artefacts extensively reduced artefacts at the site are. This is Clarkson 2008; Reduction  Reduction intensity of cores important for understanding the transport and 2010:55) intensity  Dorsal scar number and pattern maintenance of artefacts.  Percentage of cortex Raw material  Comparison of percentages of local to exotic This test is important for establishing the range and (Hall 2004; Smith diversity and materials intensity of mobility. Higher proportions of and diversity 2011) richness  Ratio of material types to artefact numbers of exotic materials linked to higher mobility. Core  Dorsal scar number and pattern This test helps determine the flaking strategies employed (Clarkson and David preparation  Platform preparation (presence and type) at the site. The attributes relate to the creation of 1995; Kessler et al. and flake  Flake termination predictable and successful flake removals and 2009:151; Macgregor removal  Consistency in flake shape characteristics of previous flake removals. 2005; Whittaker strategy 1994)  Proportions of bending to hertzian initiations These attributes assist in identifying the objective pieces (Pelcin 1997; Phagan Objective  Proportions of different platform types flaked in each period. Initiation and platform 1985; Rezek et al. piece  Bulb presence and morphology characteristics are critical for understanding the kind of 2011:1350; Root technology  Variability in platform size, flake size, and technology flakes were likely produced from. Flake 2004) and reduction metrics can be used to indicate comparative stages of stage relative thickness  Platform angle reduction.

25 The combinations of attributes, outlined above, will be essential to forming inferences about the provisioning strategies employed at Nawarla Gabarnmang and for testing the predictions outlined in Chapter Two. The follow explains how these tests allow an understanding of past flaking strategies, technological organisation, and site use.

The intensity of occupation (evidenced by artefact discard rates) can be used to infer site usage (following Haberle and David 2004; Smith 2006), which is inferred to have been influenced by variation in mobility and characteristics of site use such as duration of stay, frequency of visitation, and numbers of people visiting. Breakage indices, generated by comparing the rate of intact to broken flakes, have been linked to intensity of core reduction, which is important because it might indicate increased emphasis on obtaining maximum utility from a core (Austin 1997; Baulmer and Downum 1989; Carr and Bradbury 2001; Holdaway et al. 2004). Also important here are rates of platform preparation, which can be related to overall core reduction strategies (Whittaker 1994).

Raw material diversity can be directly related to mobility (Hall 2004; Kelly 1995). An increase in raw material richness can be inferred to relate to higher levels of mobility, as this would directly influence the range of exotic materials available and provide more opportunity for embedded procurement (Binford 1979; Clarkson 2006). Conversely, increased emphasis on local raw materials would be expected to relate to low and mostly local mobility (Andrefsky 1994).

Reduction intensity for cores and retouched artefacts is also important for inferring past levels of mobility. Higher levels of reduction are expected to relate to higher levels of mobility, as this would create impetus for extending artefact use-lives or obtaining maximum core utility (Clarkson 2008; Hiscock 2009). Overall low levels of reduction are more likely to relate to low mobility or movements based around a central place where there is less need to maximise utility (Beck et al. 2002; Kessler et al. 2009). Flake size is also linked to reduction intensity and can be influenced by transport concerns and a need for increased portability. Reduced flake size can be related to higher levels of reduction intensity and/or increased maintenance activities and is expected to be correlated to increased mobility and the use of mobile toolkits (Hiscock 2006; Sullivan 2001).

26 Descriptive platform attributes—such as platform, initiation and bulb types—combine to characterise the form of the core or objective piece that the flakes were created from (see Replicative Experiments section). Flaking strategy, when understood in relation to reduction extent and raw material use, can be used to infer overall technological strategies (Flenniken and White 1985; Root 2004). Associated with platform description, is the morphology of the dorsal surface, which can be used to indicate the number of prior removals and to infer reduction extent (Andrefsky 2005b; Carr and Bradbury 2001).

Statistical Analyses

Statistical analyses provide a means of testing the significance and strength of differences observed in the data, specifically the probability that any differences between samples could result from the vagaries of sampling rather than real differences between populations (Drennan 2009:151). The two types of analysis used in this project were t-tests and chi-square tests. T- tests evaluate the probability that two samples derive from the same population. T-tests are well-suited to continuous datasets such as metrical attributes recorded on flakes and cores (Drennan 2009:153). T-tests are very robust and resistant to effects of skewed datasets. Similarly, the chi-square test looks at the probability that two samples could be as different as they are if they came from an identical population. This test is well-suited to categorical data such as the recorded descriptive attributes (Drennan 2009:182). The results of each test are presented following the conventions set out by Drennan (2009:155, 187).

Summary

The rationale behind the methodology and methods employed in this study has been presented to demonstrate how this project addresses the research question and aims. Justifications for the selection of debitage attribute analysis and MANA were presented to establish that this is a robust approach to understanding variability in the Nawarla Gabarnmang lithic assemblage. This chapter also provided details about the Nawarla Gabarnmang excavation, dating and stratigraphy, as well as the statistical methods employed. The critical component of this chapter was the presentation of a series of technological tests and their rationale, as these tests will allow evaluation of the technological predictions proposed in Chapter Two.

27 CHAPTER 4: RESULTS

Introduction

This chapter presents results for each technological test outlined in Chapter Three. It concludes with a summary of each test, highlighting key trends in the data pertinent to addressing the research question and aims of this project.

Occupational Intensity

The intensity with which a site was occupied in the past is often measured by the number of artefacts and other cultural material discarded over time. As there is a paucity of other materials at Nawarla Gabarnmang, only stone artefacts can be used to infer occupational intensity. Figure 4 shows that per litre of sediment there is an initial increase in artefacts discarded in Periods 5 and 4, followed by a decline during the LGM, then sustained increases during the terminal Pleistocene/early Holocene (Periods 3 and 2), with a final decline in the last few spits.

Figure 4 Volumetrically adjusted artefact discard rate for Square A. (Refer to Appendix 3 for raw data.)

28 In Situ Flaking Intensity

The Square A assemblage is principally composed of debitage (98.9%). Table 5 presents counts and percentages for broad artefact categories. As described in Chapter Three, only complete/proximal flakes, retouched artefacts and cores were analysed in detail, equating to approximately one-quarter of all artefacts (25.2%).

Table 5 Counts and percentages of the four main artefact types for the entire Square A assemblage (see Appendix 3).

Complete/Proximal Flakes Retouched Cores Broken Total

Total number 1706 59 15 5272 7052 of artefacts % of Total 24.19 0.84 0.21 74.76

Retouched artefacts and cores comprise only 1.1% of the entire assemblage, and 4.2% of the total analysed assemblage (Figure 5). This indicates a high level of in situ flaking (i.e. creation of broken and complete/proximal flakes), but low levels of implement manufacture and/or discard at the site.

Figure 5 Breakdown of the analysed Square A assemblage, comparing the frequency of cores, debitage, retouched artefacts and flake pieces (see Appendix 3 for raw data).

29 Breakage rates are important proxies for intensity of stone working. Figure 6 indicates that there are consistently higher numbers of broken artefacts compared to those complete or retaining a platform, though this ratio changes through time. This may reflect breakage relating to intensive manufacture (e.g. through end shock and cone splits), or less likely trampling, given the position of the square in an area unlikely to have seen heavy traffic (David et al. 2013a:Figure 2). Figure 7 illustrates that there are numerous flaked pieces throughout the assemblage, but they are always outnumbered by identifiable pieces and that their representation in the assemblage decreases through time.

Figure 6 Ratio of complete/proximal to broken flakes. Figure 7 Ratio of identifiable (i.e. complete, proximal, and broken flakes) to flake pieces.

Raw Material Richness

MANA indicated that there were 111 analytical nodules represented in the assemblage from Square A (refer to Appendix 5 for description of each). Table 6 presents the number of different nodules associated with each raw material type and Figure 8 presents the distribution of each material type graphically. Quartzite is the only material that has both local and exotic variants, and local quartzite dominates all periods. Rare material types such as chalcedony, chert, silcrete and volcanics fluctuate in presence throughout the sequence. An interesting

30 trend illustrated in Figure 8 is the spike in between spits 20 and 18 (Period 2) coinciding with a sharp decrease in quartzite. Cortex is rare throughout the Nawarla Gabarnmang assemblage; however, the rounded cortex present on several quartz flakes suggests that this material was sourced from pebbles rather than outcropping quartz. Unfortunately the scarcity of cortex present on the other exotic material types does not permit a similar understanding of their likely sources.

Table 6 Number of analytical nodules for each raw material in the Square A assemblage and indication of which periods they occur in.

Raw Material # AN Local/Exotic 1 2 3 4 5 Chalcedony 6 Exotic x x

Chert 24 Exotic x x x x Quartz 2 Exotic x x x x Quartzite 50 Local/Exotic x x x x x Silcrete 21 Exotic x x x x Volcanic 8 Exotic x x x

Figure 8 Percentage of each main raw material type across all spits. All complete/proximal flakes, broken flakes, retouched artefacts, and cores are included here.

31 Figure 9 presents the percentage of local to exotic raw material for each time period. There is a clear increase in the proportion of exotic raw material discarded in Period 1 and Period 2. Raw material richness, expressed as a ratio of total artefact numbers to number of analytical nodules, is presented in Figure 10. This graph is interesting as Period 5 (with only one raw material type) is most similar to Period 2, which has the highest number of exotic material types. This suggests that sample size strongly influences the proportions of exotic to local material shown in Figure 9, particularly the increase in Period 2 (see Appendix 6 for raw data).

Figure 9 Comparison of local to exotic raw material. Sample sizes: Period 1 N Exotic=445 N Local=751; Period 2 N Exotic=125 N Local=157; Period 3 N Exotic=26 N Local=101; Period 4 N Exotic=23 N Local=95; Period 5 N Exotic=0 N Local=41.

Figure 10 Raw material richness based on MANA, total number of artefacts divided by number of analytical nodules.

MANA gives a strong indication of how much exotic material is present in each period. Figure 11 presents a ratio of exotic flakes to exotic nodules demonstrating a substantial increase in the number of exotic artefacts through time. Overall, the raw material data indicate that there is a substantial increase in the range and proportion of exotic material types through time, with Period 1 showing the most extensive increase in exotic materials and overall raw material richness of any period. Period 2 also shows a substantial increase in exotic material, as show in Figure 9 and Figure 11, but Figure 10 suggests caution in its interpretation as increased sample size appears to be an issue.

32

Figure 11 Ratio of exotic flakes relative to the number of exotic nodule groups present in each period.

Volcanic materials are rare and were most commonly used in the manufacture of ground-edge axes. As illustrated in Figure 8, these are present only in the pre- and post-LGM phases. One of the most exciting discoveries from Nawarla Gabarnmang thus far was a ground-edge axe flake from Square A dating to ~35,000 BP (Geneste et al. 2010). Additional flakes are also present, mostly in the post-LGM phases (Table 7). As a special flake category axe flakes are excluded from the debitage results presented below.

Table 7 Axe flakes from Square A; note that the flake from Spit 30 was previously described by Geneste et al. 2010.

Spit Time Period Raw Material Complete Broken Type 1 1 Volcanic Broken Marginal 3 1 Quartzite Complete

4 1 Volcanic Complete

6 1 Volcanic Broken Distal 9 1 Volcanic Broken Medial 19 2 Quartzite Broken Medial 30 4 Volcanic Complete

33 Reduction Intensity

There are 59 retouched artefacts in the Square A assemblage, comprising 0.8% of the total assemblage. Table 8 presents counts for the different retouched artefact categories, revealing that over half (69.5%) of all retouched artefacts are retouched flakes, which would traditionally be called ‘scrapers’. Owing to the general paucity of retouched artefacts and lack of formal types, traditional typological categories are not employed here.

Table 8 Frequency of different retouched artefact types across the different occupation periods and their percentage of the total retouched assemblage.

Bifacial Point Retouched Flake Unifacial Point

Period # % of total # % of total # % of total 1 12 20.3 27 45.7 5 8.5 2 1 1.7 1 1.7

3 6 10.2

4 6 10.2

5 1 1.7

Total 13 22 41 69.5 5 8.5

The index of invasiveness is a robust measure of reduction intensity (Clarkson 2002). Figures 12 and 13 present the index of invasiveness for retouched artefacts and points. Retouched artefacts suggest an increase in reduction extent later in time (note that the first point occurs in spit 11, Period 2), although sample size, as presented in Table 8, is too small to allow statistical testing. Points are consistently more extensively reduced, emphasising the overall low levels of reduction for retouched flakes at this site. In keeping with the predictions from Chapter Two, Period 1 shows the highest and most consistent evidence for intensive reduction, this can be connected to use of mobile toolkits as part of a provisioning of individuals strategy.

34

Figures 12 and 13 Index of invasiveness results for all retouched flakes (left) and all points (right) in Square A.

The Square A assemblage contains only 15 cores. Cores are only common in the LGM and late Holocene periods and the most common are multiplatform, direct percussion cores (Table 9). Core morphology shows a trend of decreasing size through time, although sample sizes are too small to allow statistical testing. Although in Period 1 this decrease is not accompanied by an increase in the extent of reduction (i.e. increased number of flake scars and rotations) as would be expected. This could be related to small nodule use or to smaller cores not preserving evidence for prior flaking, both possibilities are in keeping with the prediction for a provisioning of individuals strategy in Period 1. The high extent of reduction displayed in Period 3 is consistent with the provisioning of individuals strategy predicted for that period. While the increased core mass in Periods 2 and 4 is in line with predictions for a place provisioning strategy with expected decreases in intensive reduction.

Table 9 The extent of core reduction for each time period containing cores, including the mean mass (g), number of scars >15mm, number of core rotations, frequency of cores and the raw material types used. Time Mean Mean Mean Number Percentag Raw Material Types Period Mass Scars Rotation of Cores e of Total Chert (1), Exotic Quartzite 1 2.68 0.67 2.67 9 60 (2), Local Quartzite (6) 2 56.53 4 3 1 6.7 Local Quartzite (1) Local Quartzite (3), Quartz 3 16.57 2.67 4.33 4 26.7 (1) 4 121.55 4 4 1 6.7 Local Quartzite (1)

35 Dorsal scars can be used to identify reduction intensity on flakes. Figure 14 presents the mean number of dorsal scars per period. Period 4 and 5 are very similar (4-5, t = -1.033, df = 154, p = 0.303), as are Period 2 and 3 (2-3, t = -0.349, df = 396, p = 0.727). An increased range makes Period 1 distinctive. Dorsal scar directionality is another way to understand reduction intensity and is related to extent of core rotation. Figure 15 presents a ratio of multiple to single dorsal scar directions. Strikingly, the pre-LGM periods stand out by having more flakes with multidirectional scars for every flake with a unidirectional scar pattern. This is a statistically real trend, Period 4 and 5 are distinctive from all other periods but are statistically similar to each other (4-5, χ2 = 0.031, p = 0.860). However, the decreases in dorsal scar number and directionality in Periods 1 and 2 could be related to a decrease in flake size (see Figures 21 and 22) reducing the likelihood that multiple dorsal scars would be identified (cf. Andrefsky 2005a:752).

Figure 14 Box plot showing dorsal scar numbers. Sample size: Period 1 N=1152; Period 2 N=278; Period 3 N=120; Period 4 N=114; Period 5 N=42.

Figure 15 Ratio of multiple to single dorsal scar directions. Sample size: Period 1 N=990; Period 2 N=246; Period 3 N=95; Period 4 N=98; Period 5 N=38.

36 Core Preparation and Flake Removal Strategy

The dorsal scar numbers presented above showed an increase in range over time but a lower mean number following the LGM, which followed a pattern of decreasing multidirectional scar patterns (see Figure 14 and Figure 15). Rates of platform preparation are important for understanding core preparation strategies and were predicted to relate to investment in maintenance or flake predictability. As shown in Figure 16 the final three periods (1, 2 and 3) are very similar in the amount of platform preparation they exhibit (1-2, χ2 = 0.339, p = 0.560; 1-3, χ2 = 0.008, p = 0.929; 2-3, χ2 = 0.187, p = 0.665). The pre-LGM Pleistocene periods are statistically distinctive from all other periods as well as each other (4-5, χ2 = 13.4, p < 0.0005). The reasonably high rates of platform preparation in Periods 1 and 3, fits with the predictions for increased investment in maintainability as part of a provisioning of individuals strategy for these periods. Periods 4 and 5 show an interesting change with a statistically significant increase in Period 4, this was not predicted.

Figure 16 Percentage of different platform preparation types. ‘Other’ refers to platforms that were facetted, ground or a featured a combination of techniques, these are presented as one category as they represented such a small proportion of the assemblage. Sample sizes: Period 1 N=1151; Period 2 N=278; Period 3 N=120; Period 4 N=114; Period 5 N=42.

37 Objective Piece and Reduction Stage

Figure 17 presents proportions of different initiation types, which can indicate objective piece form (as demonstrated by the replicative experiment data presented in Chapter Three). Periods 4 and 5 are statistically similar to each other and to Period 2 in their high proportion of bending initiations (4-5, χ2 = 0.163, p = 0.686; 2-4, χ2 = 2.070, p = 0.150; 2-5, χ2 = 2.414, p = 0.120). All other periods are significantly different when compared to the pre-LGM periods (1-4, χ2 = 4.117, p = 0.042; 1-5, χ2 = 3.554, p = 0.059; 3-4, χ2 = 5.067, p = 0.024; 3-5, χ2 = 4.793, p = 0.029). Period 3 is distinct with a high proportion of hertzian initiations and Period 1 similarly for a relatively even mix of attributes; these periods were expected to be distinctive due to their high risk environmental contexts. Periods 2, 4 and 5 are similar with high rates of bending initiations; this is in keeping with the prediction for similarities in flaking strategies between these periods given their relatively similar environmental contexts.

Figure 17 Percentages of different initiation types. Sample sizes: Period 1 N=1152; Period 2 N=278; Period 3 N=120; Period 4 N=114; Period 5 N=42.

Platform types can be indicative of reduction, with cortical platforms related to initial stages, single platforms to early stages of reduction and multiple platforms indicating more extensive core rotation and reduction. Replicative experiments (see Chapter 3) have demonstrated that

38 this can also be an important attribute for recognising objective piece form. Figure 18 shows proportions of different platform types. Proportions of single and multiple platform types, being the most common, were statistically compared. This analysis indicated that Period 1 is very similar to Period 2 (χ2 = 0.885, p > 0.347) but is statistically different from every other period (1-3, χ2 = 6.466, p > 0.011; 1-4, χ2 = 5.497, p > 0.019; 1-5, χ2 = 4.806, p > 0.028). All other periods showed non-significant relationships. The relative increases in multiple, crushed and other platform types in Period 3 suggests more intensively reduced core forms in keeping with the predicted provisioning of individuals strategy.

Figure 18 Percentage of different platform types; ‘other’ refers to focalised and mixed platform types. Sample sizes: Period 1 N=932; Period 2 N=214; Period 3 N=88; Period 4 N=82; Period 5 N=38.

Bulb attributes can be used to indicate the force applied to detach a flake and hence the flaking strategy or technique used. The ratio of flakes without a bulb to those with bulbs is presented in Figure 19; higher values indicate more flakes without bulbs. Period 3 stands out and is statistically distinctive from the Holocene periods (3-1, χ2 = 3.989, p = 0.046; 3-2, χ2 = 5.205, p = 0.023), but similar to the pre-LGM periods (3-4, χ2 = 1.233, p = 0.267; 3-5, χ2 = 0.579, p = 0.447). Figure 20 presents the proportion of different bulb types, flakes with diffuse bulbs and those without any bulb are most common. Statistically the difference evident between Period 3 and all other periods is significant (3-1, χ2 = 15.249, p < 0.0005; 3-2, χ2 = 4.024, p = 0.045; 3-

39 4, χ2 = 5.220, p = 0.022). The sample size of Period 5 is too small to be included in statistical tests. Overall, these results strongly suggest that Period 3 is distinctive from all other periods in terms of applied force and flaking strategy. This finding is in keeping with a predicted change in technological strategy in this period in response to the substantial environmental changes during the LGM.

Figure 19 Ratio of flakes without bulbs to those with bulbs. Sample sizes: Period 1 N=1152; Period 2 N=278; Period 3 N=120; Period 4 N=114; Period 5 N=42.

Figure 20 Percentage of different bulb types. Sample sizes: Period 1 N=592; Period 2 N=144; Period 3 N=72; Period 4 N=63; Period 5 N=17.

Platform size can be a good measure of core size and reduction extent, this metric shows a strong decrease over time (Figure 21). The pre-LGM periods are statistically very similar (4-5, t = 0.201, df = 178, p > 0.841), and Period 4 is significantly different from the Holocene periods (4-1, t = -2.085, df = 1334, p > 0.037; 4-2, t = -2.089, df = 408, p > 0.037). Period 3 is statistically similar to the Holocene periods (3-1, t = 0.282, df = 1343, p > 0.778; 3-2, t = - 0.325, df = 417, p > 0.745), which is clearly shown in Figure 21. The sample size of Period 5 is too small for statistical testing, but appears similar to Period 4. Overall, the statistically significant increase in Periods 4 and 5 is in keeping with the predicted provisioning of places strategy for these periods, in which a lower emphasis on reduction was expected.

40

Figure 21 Mean platform size. Sample sizes: Period 1 N=1215; Period 2 N=289; Period 3 N=130; Period 4 N=121; Period 5 N=59. Flake mass is a powerful metric for comparing differences in flake size and was predicted to relate to the extent of reduction, with decreases in size linked to higher mobility. The most obvious change in Figure 22 is the increase in Period 3 and substantial decrease in Period 1. Period 1 is significantly different when compared to all other periods, except Period 5 as sample size is an issue (1-2, t = -2.557, df = 1480, p > 0.011; 1-3, t = -3.277, df = 1321, p > 0.001; 1-4, t = -2.488, df = 1310, p > 0.013). The substantial decrease in Period 1 is in keeping with the proposed provisioning of individuals strategy. Whilst the increase in Period 3 could relate to an increased reliance on local materials, as predicted in Chapter Two.

Figure 22 Mean flake mass. Sample sizes: Period 1 N=1194; Period 2 N=288; Period 3 N=129; Period 4 N=118; Period 5 N=58.

41 Relative thickness is based on the assumption that as reduction proceeds flakes will become thinner relative to their overall size (Sullivan 2001); this data is presented in Figure 23. Statistically the late and early Holocene periods are very similar (1-2, t = 0.169, df = 567, p > 0.866) and the LGM period has a weak non-significant relationship with the preceding and following periods suggesting some similarity between them (3-2, t = 1.614, df = 163, p > 0.108; 3-4, t = 1.444, df = 89, p > 0.154). All other periods are significantly different. Relative thickness increases through time, which is interesting given that platform size generally decreases (Figure 21) and flake mass rises and then falls (Figure 22). This suggests that this metric is more closely related to technology type rather than purely reduction extent as it does not track the same changes as the other metrics presented.

Figure 23 Mean relative thickness. Relative thickness is calculated as (Length+Width)/Thickness. Sample sizes: Period 1 N=454; Period 2 N=115; Period 3 N=50; Period 4 N=41; Period 5 N=31.

Summary of Results

This section summarises the key findings from the data presented. Non-conclusive results from the tests above are provided in Appendix 4 and artefact illustrations featuring key types and attributes identified for each period are provided in Appendix 7.

42 During Period 5 artefact discard rates are low. This period has the lowest ratio of complete to broken flakes (i.e. many more broken) and the highest rate of flaked pieces. Only local quartzite is present in these levels. There is a single minimally retouched flake. Other measures of reduction intensity are more informative, with a high mean number of dorsal scars (2) and high ratio of multidirectional dorsal scars. This period exhibits a significantly lower proportion of platform preparation to other periods and features the highest proportion of bending initiations, statistically similar to Periods 2 and 4. Flake metrics indicate large platform sizes and relatively high flake mass, but thinnest flakes relative to their overall size.

In Period 4 artefact discard rates are do not fluctuate dramatically and there is a reasonably high rate of complete flakes and flaked pieces. Local material dominates this assemblage but a reasonably wide range of exotic nodules were also identified in small quantities. There are a small number of retouched flakes, which are minimally reduced. Other measures of reduction intensity show a contrasting trend with a high mean number of dorsal scars (2) and a high ratio of multidirectional dorsal scars. This period records the highest rate of platform preparation and is statistically different from all other periods. There is a high rate of bending initiations and this period is statistically similar to Period 2 and 5 in regards to this attribute. Flake metrics demonstrate the largest mean platform sizes and a high flake mass but relatively thin flakes similar to Period 5.

Period 3 features decreasing artefact discard rates, a high rate of broken artefacts and flaked pieces. Local raw material is most common but there are a reasonably wide range of exotic nodules. There are a small number of minimally reduced retouched flakes and a high number of relatively large and intensively reduced cores. Statistically similar to Period 2 is the low mean number of dorsal scars (1) and low ratio of multidirectional scars. This period displays a statistically similar proportion of platform preparation as Periods 1 and 2. Platform attributes indicate the highest proportion of hertzian initiations of any period, a trend significantly different from the pre-LGM periods. This period also has the highest proportion of multiple platform types and flakes with hertizan bulbs, and the latter are statistically significant from all other periods. In terms of flake metrics, we find a low mean platform size but the highest mean flake mass. Flakes in this period are relatively thick.

43 In Period 2 artefact discard rates increase. There are reasonably high rates of broken artefacts and low ratio of flaked pieces to identifiable flakes. There is a wide range of raw materials represented; the highest proportion of exotic material but raw material richness indicates that a higher sample size does influence this. This period has a low number of retouched artefacts and a single core. Flakes have a low mean number of dorsal scars (1) and low ratio of multidirectional dorsal scars. Similar to Period 1, this period has reasonably high proportions of overhang removal. Platform attributes indicate a high proportion of bending initiated flakes and single platform types. Flakes from this period have the smallest average platform sizes and relatively low mean mass. Similar to Period 1, flakes in this period are relatively thick.

Period 1 features a peak in artefact discard, a comparatively high ratio of complete to broken flakes and the lowest ratio of flaked pieces. This period is the richest in terms of raw material use. There are increased levels of reduction evident on retouched artefacts in this period. Cores are comparatively smaller but are generally not intensively reduced. Flakes in this period show the widest range of dorsal scars but overall have a low mean number (1) and low ratio of multidirectional dorsal scars. Core preparation is high with ~40% of flakes exhibiting overhang removal. Platform characteristics show relatively even numbers of bending and hertzian initiations and statistically distinctive platform types with the highest proportion of single platforms. In terms of metrics, flakes exhibit small platform sizes and the lowest mean mass, which is statistically significant from all other periods. However, flakes are comparatively thick and similar in this respect to Period 2.

Summary

This chapter presented the results from the analysis of lithic artefacts from Square A of Nawarla Gabarnmang. Overall, the two pre-LGM periods were most alike and also shared certain characteristics with the terminal Pleistocene/early Holocene period. The LGM and late Holocene periods are the most distinctive for a number of attributes. The nature of changes and their technological significance are discussed further in the following chapter.

44 CHAPTER 5: DISCUSSION

Introduction

This chapter discusses the project’s findings in light of the research question and aims. Specifically I engage with the predictions made in Chapter Two for technological organisation and provisioning strategies over time and evaluate these following the technological tests outlined in Chapter Three and the results presented in Chapter Four. These predictions form a way to directly address the research question, how might have changing environmental contexts influenced the technological strategies people used over time at Nawarla Gabarnmang? Following this discussion, I directly engage the research problem and draw conclusions as to how this project has contributed to solving it. I close by reflecting on the success and significance of this study and highlighting directions for future research.

Testing Predictions at Nawarla Gabarnmang

45,000–30,000 cal. BP: Initial Occupation pre-LGM Pleistocene (Periods 4 and 5)

The two analytical periods used between 45,000 and 30,000 cal. BP are technologically very similar and are discussed here together. The palaeoenvironmental data suggests that initial occupation occurred during a period of high productivity and seasonality influenced by an active summer monsoon system (Johnson et al. 1999:1150; Reeves et al. in press-a:4). Following Kuhn (1995:22) in Chapter Two a provisioning of places strategy was predicted for this period anticipating a low investment in artefact reduction and a reliance on local material.

The rates at which artefacts are discarded is employed as a proxy for how intensively people used the site through time, i.e. the more it was used the more likely that artefacts would be discarded (Smith 2006)—although there is some debate over the accuracy of this approach (Hiscock 1981; Pecora 2001). As revealed in Chapter Four the low artefact discard rates of these periods suggest that Nawarla Gabarnmang was not intensively occupied at this time. The demonstrated high proportions of flaked pieces could relate to shatter during intensive core reduction (Austin 1997; Baulmer and Downum 1989:112) or alternatively to the modification of the shelter itself—another important activity conducted at this site (Delannoy et al. 2013).

45 The overwhelming use of local material during these periods suggests that the site was valued as raw material source. However, in Period 4 exotic material becomes more abundant confirming that people were indeed mobile but that there was not a strong emphasis on the transport of materials. Overall, the emphasis on local material suggests that there was little impetus to transport stone over long distances even though people were highly mobile.

The reliance on locally available material has implications for how we understand human movements across the landscape, as well as the technological nature of the assemblage. Local raw material was most likely directly sourced from the ceiling or pillars of the site or the surrounding escarpment. The lack of cores in Periods 4 and 5 is therefore not unexpected, as the primary material could be removed as large flakes rather than individually formed nodules. Although conversely a strategy of sourcing large flakes for subsequent use could be expected to manifest as a predominately retouched assemblage, which is not the case for this period and as such will require further attention. The paucity of flakes with cortex is also testament to the nature of the local material; artefacts generally exhibit a weathered exterior consistent with their removal from the ceiling, pillars or escarpment. Overall the immediately local supply makes it difficult to evaluate this assemblage based on traditional assumptions of reduction intensity, such as by percentage of cortex and ratios of cores to flakes (Andrefsky 2005a; Sullivan 2001:194).

These levels contain extremely low numbers of retouched artefacts and cores, which all display low levels of reduction. However, flake dorsal scar patterns suggest that these artefact classes do not accurately represent reduction levels across the entire assemblage. The high mean number of dorsal scars and significantly higher ratio of multidirectional scar patterns suggest that some effort was expended obtaining more utility from cores, perhaps as a way to shape and predict future flake removals. Another key line of evidence to support this suggestion is the proportion of artefacts showing platform preparation. Interestingly, Period 5 records the lowest levels of this activity but Period 4 the highest, which would indicate an adjustment to flaking strategies between the two periods. In Period 4 it does appear that there was some investment in improving the predictability of flake removals (following Whittaker 1994:101), and this could be interpreted as a modification to improve the existing strategy.

46 The most interesting finding from these periods is the nature of the tool or core forms suggested by the debitage attributes. The dominance of bending initiations, reasonably low platform angles (73–65°) and relatively thin flakes suggests that the technology does not fit the traditional conception of of a ‘core tool and scraper tradition’ (Bowler et al. 1970:52). Instead, it appears that the flaking strategy focused on the production of flakes from low-edge angled pieces (see also Andrefsky 2005a:724; Patterson 1990) as would be expected if the shelter itself was being flaked. These pieces could be retouched flakes or thin cores and will be the subject of further investigation as the isolation of these flakes and examination of their platforms and dorsal surfaces may help understand the technology they were produced from better; this would include whether they display use-wear and could have been produced from utilised tools as ‘use-flakes’ (cf. Dortch 2004:89).

In summary, the results for the initial periods are consistent with the prediction for a provisioning of places strategy at Nawarla Gabarnmang. Rather than transporting extensive amounts of raw material across the landscape the locally abundant material was the focus—we could say that the site itself was already provisioned with material and so no extra effort was expended to transport stone. The debitage data suggest predominately low levels of reduction, which is in keeping with the expected low need to transport materials (Beck et al. 2002; Hiscock 2009:90–91). There is also evidence for the consistent production of relatively large thin flakes with bending initiations, a flaking strategy in keeping with the production of flakes with low-angled profiles, a finding not previously identified for this time period in Australia.

21,500–15,200 cal. BP: Last Glacial Maximum (Period 3)

From ~30,000 cal. BP onwards Nawarla Gabarnmang appears to have been abandoned or was at least occupied sporadically; as this period marks the initial onset of glaciation, this might be part of the reason people lessened or ceased their use of the site (Reeves et al. in press-a:7). Nawarla Gabarnmang was reoccupied during the peak of the LGM around 21,500 cal. BP. This suggests that the immediate area was not as severely affected by aridity as other areas of the surrounding landscape and that it may have served as a ‘refuge’ (Veth 1993).

Extreme variability in resource and water availability is strongly suggested by the palaeoenvironmental evidence for the LGM (De Deckker 2001; Kershaw 1995:666), placing

47 more impetus on groups to alter their subsistence and technological strategies to alleviate increased risks (Hiscock 2006:71; Veth 1993). As such, in Chapter Two I predicted a provisioning of individuals strategy, which would ensure that individuals had necessary tools with them at all times and would thus be able to take advantage of resources when encountered. This strategy is advantageous when there are increased risks associated with not obtaining resources (Hiscock 1994; Kuhn 1995).

Artefact discard rates are highest at the start of Period 3 and decrease through time (i.e. as you move up the sequence artefact numbers subsequently decrease). If discard rates are taken to represent the intensity of human use of the site then this trend suggests that occupation was most intensive during the peak of the LGM (when this period starts) and decreased through time with climate amelioration, perhaps when other areas of the landscape became more conducive to higher levels of residential mobility. Similarly to the pre-LGM periods, local raw material dominates but there is a comparatively higher ratio of exotic nodules, meaning the increase in exotic material in this period is related to an increase in different exotic sources/nodules being used rather than simply due to a larger sample size. This suggests that material transfer was not particularly important or intensive at this time, yet mobility across the landscape was reasonably wide-ranging (cf. Clarkson 2006:186; Smith 2006:405).

Retouched artefacts in Period 3 are minimally reduced, however, the comparatively large numbers of cores (26.7% of all cores in the assemblage) exhibit the highest mean number of flake scars and rotations. This evidence, coupled with high ratios of broken artefacts, suggests intensive levels of core reduction. This is further supported by a statistically significant decrease in platform sizes and an increase in multiple, crushed and other platform types, characteristics shown to be indicative of higher numbers of removals and increased reduction (Andrefsky 2005a:750).

However, other evidence complicates the above interpretation, suggesting a mixture of strategies rather than solely intensive core reduction. Dorsal scar numbers and directionality are more in keeping with low reduction intensity levels, while the high mean flake mass and increase in flake relative thickness suggest that technological strategies were not entirely focused on intensive core reduction. The high proportion of hertzian initiations and bulbs indicate that greater force was applied to detach flakes, supported by the increase in flake

48 metrics. This can be inferred to relate to the creation of large flakes that had more reduction/use potential for transport across a variable landscape (Rasic and Andrefsky 2001:64), a suggestion supported by presence of relatively larger cores.

Overall, the prediction of a provisioning of individuals strategy is met, albeit with some deviations. There is a clear focus on local materials and more intensive use of the site during the LGM peak, which suggests that the ease of access to quality material at Nawarla Gabarnmang, along with the other advantages, e.g. access to reliable water and resources, may be why this site was occupied at that time. Occupational intensity decreased through time, this trend appears to be correlated to climatic amelioration. There is some evidence for intensive core reduction, which is in keeping with a strategy to realise as much potential from cores as possible; however, there is a secondary/mixed technological focus with the production of large, thick flakes being relatively common. These artefacts would have had potential for extended use and maintenance as part of mobile toolkits (Kuhn 1994; Lycett and Eren 2013; Rasic and Andrefsky 2001). A clearer understanding of this mixture could be achieved by analysing the local and exotic materials separately as they may have been employed in distinct ways in the overall technological system. As there is not a lot of evidence suggesting tool maintenance was conducted during Period 3 it is more likely that the site functioned as a raw material source with a focus on production for transport.

14,800–8500 cal. BP: Terminal Pleistocene/Early Holocene (Period 2)

Palaeoenvironmental data suggests that post-LGM amelioration was slower in northern Australia than elsewhere (Kershaw 1995:666), and that the monsoon did not return until around 14,000 cal. BP marking the start of identifiable improvements in stability and environmental productivity (Reeves et al. 2008; Wyrwoll and Miller 2001). Given this context, in Chapter Two a provisioning of places strategy was predicted with lessened need to transport materials and pursue risk-minimisation strategies such as increasing reduction intensity.

In direct contrast to the preceding period across Period 2 artefact discard rates increased, a phenomenon potentially explained as increased intensity of site use through time and linked to increasing local resource availability with climatic amelioration, which may have drawn

49 people to the site. The expectations for this period align with the pre-LGM periods in terms of human responses to productive environmental contexts. One significant difference is the rate at which exotic raw materials were used and discarded. This period exhibits the highest rates of exotic material discard, however, a measure of raw material richness suggests that this is linked to increased sample size rather than an overwhelming emphasis on exotic materials. This suggests that rather than reverting to the pre-LGM strategy, i.e. focusing principally on local materials, people at this time altered their provisioning strategies and technological organisation in a novel manner. The evidence suggests wide-ranging mobility, which would have allowed more patches (and material sources) to be encountered, and some impetus to transport raw material with mobile foragers; though it is important to note that local lithic materials still comprise a large part of the assemblage and so use of local material continues to be important. If the overall strategy is understood as a provisioning of places, this suggests that mobility was predictable and that people had less need to intensively curate a mobile toolkit but engaged in material transport perhaps to obtain higher quality material or for other non-economic reasons (cf. Brumm 2010; Taçon 1991).

The second technological signature expected with a provisioning of places strategy relates to overall low levels of reduction; this finds more support in Period 2. Very low numbers of retouched artefacts, a single large core, a low mean number of dorsal scars, and a low ratio of flakes with multidirectional scar patterns all suggest that minimal time was invested in heavily reducing artefacts at the site. Rather, the dominance of small-sized flakes is highly suggestive of maintenance activities, an inference supported by the replicative experiments presented in Chapter Three (see also Andrefsky 2005a:743; Baulmer and Downum 1989; Patterson 1990). This is in keeping with an altered use of the site as a place to maintain artefacts.

Debitage attributes for Period 2 suggest an interesting mix of tool and core forms. The large proportion of bending initiations and high rates of platform preparation statistically aligns Period 2 with the pre-LGM periods, suggesting again that flakes were being created from pieces with relatively low-angled profiles and that some effort was put into preparing cores for successful flake removals (following Andrefsky 2005a:724). However, the decreased size of flakes (indicated by mean mass) suggests that whilst core form was likely similar to the pre- LGM periods, core size had significantly decreased—this also places some restriction on the

50 interpretation of small-sized flakes relating only to maintenance made above, which will require further investigation (cf. Pecora 2001). Flakes from Period 2 have the smallest average platform size and, along with evidence of abundant platform preparation, suggest a strategy to constrict platform area and increase the predictability of flake removals.

During this period I argue that the technological strategy used was similar in some respects to the pre-LGM periods but that it was not a complete reversion, which strongly suggests that the choice of different strategies was not structured by environmental context alone. The flexibility identified in core forms, and inferred flaking strategies, supports the expectation of a provisioning of places strategy (cf. Clarkson 2006:176). Overall, we see a lack of restriction on movement (which was wide-ranging and likely predictable), minimal evidence for intensive reduction, a focus on maintenance activities and continued use of the site as a raw material source.

2500–250 cal. BP: Late Holocene (Period 1)

Nawarla Gabarnmang appears to have been occupied sporadically from ~9500 cal. BP for a period of ~6000 years, potentially related to increasing variability and drier conditions in the mid-Holocene driven by ENSO (Reeves et al. in press-a:9); however, given that the site was used at the peak of the LGM lowered occupation during the mid-Holocene is puzzling—more regional research is needed to understand this. The on-set of ENSO conditions and fluctuating aridity would have imposed certain restrictions on hunter-gatherer behaviour during the late Holocene, particularly related to subsistence and technology (Veth et al. 2011a) and in Chapter Two, predictions were made about how technology may have been organised at this time to off-set such risks. A provisioning of individuals strategy was argued to have been most likely, with strategies to support higher mobility and minimise risks (Hiscock 2006:71).

This period features the highest rates of artefact discard of any in this site, although discard decreased in the most recent levels. Whilst exploitation of local material and maintenance of tools continued (see below) there were also increasing numbers of retouched artefacts and cores being left at the site. This could be interpreted as longer periods or frequency of visitations and/or a change to how technological activities were organised and conducted at the site.

51 Related to this change in organisation are the rates at which cores and retouched artefacts were reduced. Core sizes (and also average platform sizes) decrease in this period, but the extent of reduction evident on those artefacts (evidenced by the numbers of scars and rotations) does not suggest that this is a result of dramatically increased reduction. However, this evidence along with the low mean number of dorsal scars and low rate of multidirectional scar patterns, needs to be qualified as it is unlikely that small-sized flakes and cores, characteristic of this period, would necessarily preserve evidence of extensive reduction (cf. Andrefsky 2005a:752) Although the high rates of complete flakes and low rates of flaked pieces, do support an inference that core reduction intensity was not extremely high (cf. Carr and Bradbury 2001). Retouched artefacts indicate higher rates of reduction. If the retouched artefact data is linked with the significant decrease in mean flake mass, this strongly suggests a focus on maintenance activities at the site rather than intensive production (following Andrefsky 2005a:743; Baulmer and Downum 1989; Bleed 1986; Patterson 1990).

The core forms suggested by the debitage platform characteristics generally reflect the range of retouched artefacts present, i.e. bifacial and unifacial points and a range of retouched flake forms. These are in keeping with an emphasis on mobile toolkits.

The high rate of exotic materials in Period 1 supports this interpretation, suggesting wide- ranging mobility and high rates of material transport. The exotic materials employed are generally of higher quality than the locally available quartzite and are well suited to a maintainable toolkit as they are easier to maintain and repair as they fracture more predictably (MacDonald 2008:224–225; Rasic and Andrefsky 2001:64). Further, if procurement of raw materials was embedded within subsistence activities as Binford (1979) proposed, then the wide range of exotic nodules suggests that numerous different patches across the landscape were in use.

The prediction for a provisioning of individuals strategy is met for Period 1. Core reduction and an emphasis on maintenance both in this assemblage. The rates and range of exotic material discarded are strongly suggestive of high mobility and the evidence suggesting a focus on maintenance is in keeping with this. Overall, the technological strategies of this period can be interpreted as a way to ensure mobile toolkits were always ready for use as a

52 way to deal with resource related risks and a need to have tools immediately available, in line with suggestions by Clarkson (2006:178) and Hiscock (2009:90).

Summary

The major occupation phases at Nawarla Gabarnmang are separated by periods of abandonment or decreased occupation. Upon reoccupation, technological strategies employed at the site tended to feature use of the site as a raw material source, a refuge/optimal camping location in times of environmental stress, or a place to maintain and/or produce artefacts. There is no evidence for directional change in technological complexity in the flaking strategies identified. From the initial period of occupation I have confidently inferred that organisational strategies were deliberate ways to take advantage of the benefits this site offered and to meet the changing needs of mobile hunter-gatherers influenced by climatic variation through time. Whilst changes in technology over time definitely occur there is no suggestion that the different time periods can be equated to different traditions. Rather it seems more likely that the changes identified over time are adjustments in technological organisation and strategies in response to changing environmental and cultural contexts. A summary of the main findings and interpretations is presented in Table 10.

53 Table 10 Summary of key findings, interpretations and inferred provisioning strategies for Nawarla Gabarnmang. Environmental Provisioning Period Key Findings Interpretation Complications Context Strategy • High • Focus on Low and/or Place Low sample productivity local raw predictable provisioning size makes it • Active material mobility with difficult to 5 monsoon • Reduction of little need to assess system low-edge transport tools statistical angled cores or stone. reliability of findings • High • Principally Low and/or Place Increase in productivity local raw predictable provisioning exotic • Active material used mobility with materials and monsoon • Reduction of little need to internal 4 system low-edge transport tools modification angled cores or stone. to flaking strategy needs further investigation. • Intensive • Primarily Strategy to Individual Mixture of aridity local material extract as provisioning technology • Highly in use but much potential types and/or unpredictable increasing from cores as possible rates of exotic possible and to reduction 3 • Intensive produce large stages needs core reduction flakes for further • Production transport, investigation. of larger while relying flakes on local stone resources. •Ameliorating • ~Even rates Flexibility and Place Small-sized • Increasing of local and lack of provisioning flakes may not productivity exotic material restriction on relate solely to • Active • Low mobility, a maintenance 2 monsoon reduction focus on activities. system intensity and maintenance production of activities at small-sized the site. flakes • ENSO- • High exotic High mobility Individual Some influenced material use and use of provisioning attributes (e.g. aridity and • Intensive curated mobile dorsal scar 1 variability artefact toolkits. numbers and • High risk maintenance patterns) need and probable further core reduction investigation.

54 Readdressing the Research Problem

In Chapter One I introduced concerns over how lithic technology from the earliest periods of occupation in Australia is represented. The question was posed as to whether examination of the Nawarla Gabarnmang lithic assemblage with contemporary methods and perspectives might support the traditional view of Pleistocene technology as ‘simple’, or whether it would suggest that this view is out-dated and in need of revision. The section to follow specifically addresses this issue in light of the interpretations offered above.

Research Approaches Key debates in Australian lithic studies have been principally based on the analysis of retouched artefacts and cores, even vey recently (e.g. Hiscock and Attenbrow 2011), despite these items only comprising ~5% of assemblages and much less than that at many sites, such as Nawarla Gabarnmang. Yet, Holdaway and Douglass (2012) have argued that in many contexts cores and retouched artefacts were not more valued or important to Aboriginal people than unretouched flakes or debitage, and that the continued bias towards the analysis of cores and tools is more reflective of research approaches than it is informative about past human behaviours. It is worth considering what might be missing from our understanding of Australian lithic technology and how it changes (or perhaps does not change) if the remaining 95% was given as much attention (Holdaway 1995).

Debitage is intimately linked to tool production and maintenance, and as such its analysis should provide evidence of these activities throughout a reduction sequence (Andrefsky 2005a:753; Rinehart 2008a). Debitage analysis potentially allows an understanding of technology in sites and time periods where retouched artefacts and cores are scarce and also provides direct evidence for lithic activities carried out at a single site, which can be extended to understand flaking strategies and technological organisation more broadly (Rinehart 2008b; Shott 2004).

There are a relatively small number of retouched artefacts and cores in the Nawarla Gabarnmang assemblage, given this an analytical emphasis on debitage is well justified. This approach allowed recognition of ongoing changes throughout time. I argue that the identified changes in technological strategies were likely a response to shifting needs due to variation in

55 environmental conditions, resource structures, and likely changing social conditions (although the latter are beyond the scope of this thesis). The identification of change throughout the sequence makes the establishment of a linear trajectory impossible and arguments of an early simplistic tradition leading directly into a more complex one are not substantiated. Rather, there is a constant rearrangement of context-dependent technological strategies.

Simple vs Complex Technology A key issue to contend with is what the identification of ‘simple’ lithic technology means for broader understandings of society at any moment in time. Balme and O'Connor (in press) emphasised that the simple lithic technology of the early Sahulian record does not accurately reflect society at that time; with more research such technologies will likely come to be seen to reflect remarkable levels of adaptability and variation across Sahul. Simple technologies, i.e. those requiring less planning or made expediently, can be used or combined in complex ways (Holdaway and Douglass 2012; Moore et al. 2009) or be part of a broader technological system where complexity manifests in other materials.

Following any conventional definition the Nawarla Gabarnmang lithic assemblage does not display evidence of complex technology during the initial occupation periods. However, the identified flaking strategies are at odds with traditionally defined notions of a simplistic early tradition (cf. Bowler et al. 1970). Indeed, previously unidentified artefact forms have been recognised from the debitage recovered from the site, in the form of low-angled cores and/or retouched flakes, such objective pieces are quite unlike those expected of the conservative core tool and scraper tradition. Another key finding of this study is that change is apparent throughout the entire occupation sequence, rather than primarily during the mid- to late Holocene. Of particular interest are the similarities or shared strategies of the pre- and post- LGM periods, a feature suggestive of a partial reversion to the earlier strategy, which would directly challenge traditional notions of unilinear movement towards complexity (Przywolnik 2005).

Axes and Organic Technology

Ground-edge axe technology is unique to northern Sahul during the Pleistocene, and its existence has been vital to arguments against claims for exclusively ‘simple’ Pleistocene lithic

56 technology. The existence of axes in the Pleistocene was first demonstrated by Schrire (1982) and their significance and antiquity in northern Sahul has since been supported by numerous other researchers (e.g. Denham et al. 2009; Geneste et al. 2012; Morwood and Trezise 1989; Veth et al. 2011b). Ground-edge axe technology has been argued to represent not only a highly efficient tool (Dickson 1981; Groube et al. 1986) but also a form of technology that could be imbued with significant social and/or symbolic meanings (Brumm 2010; Geneste et al. 2012). This technology type is present at Nawarla Gabarnmang in both the pre- and post- LGM occupation periods; determining whether the absence of this technology during the LGM is real or not will require analysis of the lithic assemblages from other squares. If so, it could have interesting consequences as it could be argued to represent a change not only technology but perhaps also in the social structure that technology was embedded in.

Organic technology is present and implicated from the very earliest periods of human activity in Australia. The very act of travelling across Wallacea to Sahul would have involved sophisticated watercraft and organic technology (Balme 2013; O'Connell et al. 2010). It has been suggested that the earliest forms of stone tools in Australia, traditionally core tool technology and heavy scrapers, were used in the manufacture of organic technology and that most retouched artefacts in early assemblages were connected with woodworking (Balme and O'Connor in press; Holdaway and Stern 2004:79). The evidence for ground-edge axe technology and its connection to the economic activities of woodworking and the extraction of arboreal resources (Dickson 1981)—not negating their potential role as important social objects— is a viable link to the unpreserved organic technologies of the people who used Nawarla Gabarnmang (cf. Balme and O'Connor in press). Thus, while technological strategies concerning lithic material at this site are not considered to be ‘complex’ (cf. Moore 2013), it is important to recognise that lithic technology exists within the broader context of technology generally and understanding technology in this way reveals more of the intricacies of Aboriginal technology and the ways it was employed to meet changing economic and social needs.

Addressing the Research Problem Stern (2009) suggested that the lithic industries of Pleistocene Sahul should be excluded from assessments of behavioural modernity owing to their simplicity, and other researchers have

57 also noted the disparity between the lithic record and the early evidence for complex symbolic behaviour and communication (e.g. Balme 2013; Balme and O'Connor in press; Hiscock and Attenbrow 2003; O'Connell and Allen 2012). Holdaway and Douglass (2012) suggested that in certain contexts lithic technology was expedient and not attributed much importance by Aboriginal people, perhaps with a practical focus on other materials or forms of technology (e.g. Meehan and Jones 2005). In other contexts lithic technology, particularly the material itself, has been shown to be highly significant and imbued with considerable cultural meanings (Bradley 2008; Brumm 2010; Meehan and Jones 2005; Taçon 1991). It is thus probable that the lack of complexity in lithic technology during the earliest periods of occupation is a product of the environmental context and social structures of that period rather than reflecting previous notions of cultural stasis or uniformity.

The appearance of points in the early Holocene, with an increased abundance of these artefact types in the late Holocene, is the only identified change in technological ‘complexity’ in this assemblage. In this case complexity is indicated by increased investment and hierarchical structuring of flaking activities (cf. Moore 2013), often interpreted as a risk-minimising behaviour (Hiscock 2006). This change may reflect the ‘radical changes’ in Aboriginal society during the late Holocene (David 2002:210), which influenced or perhaps created a new context in which technology was understood, made and used.

The absence of variation in Australian lithic assemblages before the mid- to late Holocene is most likely a product of approaches to analysis rather than actual lack of change (Balme and O'Connor in press; Holdaway 1995:759). The changes apparent in the Nawarla Gabarnmang lithic assemblage through time can be argued to reflect responses to changing in economic needs such as provisioning an individual with a mobile toolkit or having materials available at key resource locales in a landscape. While these changes may not always manifest as more complex technology they are indicative of the way that technology was organised and embedded in the daily lives of hunter-gatherers. From the earliest period of occupation changes in technology over time can be viewed as strategic ways to adapt to changing environmental and social contexts.

While there have been vast improvements in the way lithic technology is studied in Australia, our technological understanding of the earliest assemblages has lagged behind research into

58 the later changes in Aboriginal lithic technology and other aspects of behaviour and social complexity in the earliest periods of occupation (e.g. Balme 2013; Mulvaney 2013; Veth et al. 2011b). Perhaps as Australian research agendas begin to change, more attention is paid to entire assemblages, and understanding of lithic technology as embedded in broader subsistence and land-use strategies grows, we should expect a more diverse picture of lithic technology to emerge, though not necessarily one with evidence for more complex technology.

Conclusion

This project set out to investigate how changing environmental contexts influenced the technological strategies people used over time at Nawarla Gabarnmang. Through the application of debitage attribute analysis and MANA this project has identified variation throughout the occupation sequence at Nawarla Gabarnmang and by using technological organisation as a theoretical framework has sought to explain these changes as shifting technological strategies in response to variation in environmental conditions and resource structure. Critically the identification of change through time is contra to unilinear models of cultural change and emphasises the dynamic adaptability of hunter-gatherers (cf. Balme and O'Connor in press; Przywolnik 2005; Ulm 2006).

The theoretical framework of technological organisation was used in this project to understand how technology was embedded in the daily lives of hunter-gatherers and how it was connected to broader subsistence and land-use strategies. Provisioning strategies were employed as a way to understand how people might have met their technological needs under variable environmental conditions. Mobility has been evoked as a key limiting factor in this study, and has been shown to be useful for interpreting raw material variability, the intensity of core reduction and performance of maintenance activities at this site.

A key aspect of this study was the emphasis placed on debitage, which was given equal analytical footing to the retouched artefacts and cores in the assemblage, thereby allowing a more detailed understanding of the flaking strategies people used through time. Given the low numbers of cores and retouched artefacts in the Nawarla Gabarnmang assemblage this approach was critical to coming to any understanding of lithic technology over time and,

59 significantly, enabled the identification of a previously unappreciated form of technology in the pre-LGM periods (i.e. the reduction of flakes or cores with low-angled profiles). Whilst the technology types identified were generally simple, there is tantalising evidence for complexity in ground-stone technology and, by inference, organic technology. Overall, this project has identified flexibility and changing emphasis on different provisioning and technological strategies as ways to meet people’s changing needs in different contexts across time.

Future Research

It is essential to continue revaluating the assumptions that have underpinned our understanding of Australian lithic technologies. One important issue that requires broader engagement is the need to analyse entire assemblages in detail rather than focusing on retouched artefacts and cores. This holds great potential for investigating sites (or specific periods) where those objects are scarce and for obtaining a more complete understanding of the technological strategies employed by past peoples. A key aspect that was not directly engaged with in this project is the influence and changes in social structure and dynamics through time, which needs to be incorporated into future research on this topic.

As this project is a preliminary investigation into the technology from Nawarla Gabarnmang, and more broadly Jawoyn Country, the inferences made here need to be tested and verified against further evidence from this site and others across the landscape. Additionally, a key limitation of the predictions and inferences drawn here is the lack of localised palaeoenvironmental data. Establishing a palaeoenvironmental record is part of the ongoing research in this region and will be of immense value to future research.

By applying contemporary methods of lithic analysis to one of the oldest sites in the Australian record this study has provided an important revaluation of traditional views of Pleistocene technology and contributed to a growing understanding of the intricacies of the earliest occupation of Australia. The emerging image of hunter-gatherer technology is of flexibility and constant change through time, rather than adherence to any strict technological traditions, in direct contrast to previous notions of linear trajectories and increasing complexity across time.

60 REFERENCES

Ahler, S.A. 1989 Mass analysis of flaking debris: Studying the forest rather than the trees. In D.O. Henry and G. Odell (eds), Alternative Approaches to Lithic Analysis, pp.85–118. Archaeological Papers of the American Anthropological Association 1. Washington: American Anthropological Association. Anderson, A., M.K. Gagan and J. Shulmeister 2006 Mid-Holocene cultural dynamics and climatic change in the western Pacific. In D.H. Sandweiss and K.A. Maasch (eds), Climate Change and Cultural Dynamics: A Global Perspective on Holocene Transitions, pp.265–296. Academic Press: San Diego. Andrefsky, W. 1994 Raw material availability and the organisation of technology. American Antiquity 59(1):21–34. Andrefsky, W. 2001 Emerging directions in debitage analysis. In W. Andrefsky (ed.), Lithic Debitage: Context, Form, Meaning, pp.2–14. Salt Lake City: University of Utah Press. Andrefsky, W. 2005a Lithic studies. In H. Maschner and C. Chippindale (eds), Handbook of Archaeological Methods, pp.713–770. Walnut Creek: Altamira Press. Andrefsky, W. 2005b Lithics: Macroscopic Approaches to Analysis. Cambridge: Cambridge University Press. Andrefsky, W. 2007 The application and misapplication of mass analysis in lithic debitage studies. Journal of Archaeological Science 34:392–402. Andrefsky, W. 2008 Lithic Technology: Measures of Production, Use and Curation. Cambridge: Cambridge University Press. Andrefsky, W. 2009 The analysis of procurement, production and maintenance. Journal of Archaeological Research 17:65–103. Austin, R.J. 1997 Technological characterisation of lithic waste-flake assemblages: Multivariate analysis of experimental and archaeological data. Lithic Technology 24:53–68. Balme, J. 2013 Of boats and string: The maritime colonisation of Australia. Quaternary International 285:68–75. Balme, J. and S. O'Connor in press Early modern humans in Island Southeast Asia and Sahul: Adaptive and creative societies with simple lithic industries. In R. Dennell and M. Porr (eds), Southern Asia, Australia and the Search for Human Origins. Cambridge: Cambridge University Press. Bamforth, D.B. and P. Bleed 1997 Technology, flaked stone technology, and risk. In C. Barton and G. Clark (eds), Rediscovering Darwin: Evolutionary Theory in Archaeology, pp.109–140. Washington: American Anthropological Association. Baulmer, M.F. and C.E. Downum 1989 Between micro and macro: A study in the interpretation of small-sized lithic debitage. In D.S. Amick and R.P. Mauldin (eds), Experiments in Lithic Technology, pp.101–116. BAR International Series 528. Oxford: Archaeopress. Beck, C., A. Taylor, G. Jones, C. Fadem, C. Cook and S. Millward 2002 Rocks are heavy: Transport costs and Paleoarchaic quarry behavior in the Great Basin. Journal of Anthropological Archaeology 21:481–507. Binford, L. 1979 Organisation and formation processes: Looking at curated technologies. Journal of Anthropological Research 35(3):255–273.

61 Bleed, P. 1986 The optimal design of hunting weapons: Maintainability or reliability. American Antiquity 51:737–747. Bowdler, S. and S. O'Connor 1991 The dating of the Australian Small Tool Tradition, with new evidence from the Kimberley, WA. Australian Aboriginal Studies 1991(1):53-62. Bowler, J.M., R. Jones, H. Allen and A.G. Thorne 1970 Pleistocene human remains from Australia: A living site and human from Lake Mungo, Western New South Wales. World Archaeology 2(1):39–60. Bowler, J.M., K.-H. Wyrwoll and Y. Lu 2001 Variations of the northwest Australian summer monsoon over the last 300,000 years: The paleohydrological record of the Gregory (Mulan) Lakes system. Quaternary International 83–85:63–80. Bradbury, A.P. and P.J. Carr 2004 Combining aggregate and individual methods of flake debris analysis: Aggregate trend analysis. North American Archaeologist 25(1):65–90. Bradley, J. 2008 When a stone tool is a dingo: Country and relatedness in Australian Aboriginal notions of landscape. In D. Bruno and J. Thomas (eds), Handbook of Landscape Archaeology, pp.633–637. Walnut Creek: Left Coast Press. Brumm, A. 2010 'The Falling Sky’: Symbolic and cosmological associations of the Mt William greenstone axe quarry, central Victoria, Australia. Cambridge Archaeological Journal 20:179–196. Carr, P.J. and A.P. Bradbury 2001 Flake debris analysis, levels of production and the organisation of technology. In W. Andrefsky (ed.), Lithic Debitage: Context, Form, Meaning, pp.126–146. Salt Lake City: University of Utah Press. Chappell, J. 2001 Climate before agriculture. In A. Anderson, I. Lilley and S. O'Connor (eds), Histories of Old Ages: Essays in Honour of Rhys Jones, pp.171–183. Canberra: Pandanus Books. 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. 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. St Lucia: Aboriginal and Torres Strait Islander Studies Unit, The University of Queensland. Clarkson, C. 2007 Lithics in the Land of the Lightning Brothers: The Archaeology of Wardaman Country, Northern Territory. Terra Australis 25. Canberra: ANU E Press. Clarkson, C. 2008 Changing reduction intensity, settlement, and subsistence in Wardaman Country, northern Australia. In W. Andrefsky (ed.), Lithic Technology: Measures of Stone Tool Reduction, Use and Curation, pp.286–313. Cambridge: Cambridge University Press. Clarkson, C. 2010 Regional diversity within the core technology of the techno-complex. In S.J. Lycett and P.R. Chauhan (eds), New Perspectives on Old Stones: Analytical Approaches to Technologies, pp.43–59. New York: Springer. Clarkson, C. and B. David 1995 The antiquity of blades and points revisited: Investigating the emergence of systematic production southwest of Arnhem Land, northern Australia. The Artefact 18:22–44. Clarkson, C. and S. O'Connor 2006 An introduction to stone analysis. In J. Balme and A. Paterson (eds), Archaeology in Practice: A Student Guide to Archaeological Analyses, pp.159–206. Malden: Blackwell Publishing.

62 Clarkson, C. and L.A. Wallis 2003 The search for El Nino/Southern Oscillation in archaeological sites: Recent phytolith analysis at Jugali-ya rockshelter, Wardaman Country, Australia. In D.M. Hart and L.A. Wallis (eds), Phytolith and starch research in the Australian-Pacific-Asian regions: The state of the art, pp.137-152. Terra Australis 19. Canberra: Pandanus Books. Cole, S. 2009 Technological efficiency as an adaptive behavior among Paleolithic hunter- gatherers: Evidence from La-Cote, Caminade East and Le Flageolet I, France. In B. Adams and B. Blades (eds), Lithic Materials and Paleolithic Societies, pp.127–143. West Sussex: Blackwell Publishing. Cotterell, B. and J. Kamminga 1987 The formation of flakes. American Antiquity 52:675–708. Crabtree, D.E. 1982 An Introduction to Flintworking. Pocatello: Idaho State Museum. David, B. 2002 Landscapes, Rock-Art and the Dreaming: An Archaeology of Preunderstanding. London: Leicester University Press. David, B., B. Barker, F. Petchey, J.-J. Delannoy, J.-M. Geneste, C. Rowe, M. Eccleston, L. Lamb and R.L. Whear 2013a A 28,000 year old excavated painted rock from Nawarla Gabarnmang, northern Australia. Journal of Archaeological Science 40:2493–2501. David, B., J.-M. Geneste, F. Petchey, J.-J. Delannoy, B. Barker and M. Eccleston 2013b How old are Australia's pictographs? A review of rock art dating. Journal of Archaeological Science 40:3–10. David, B., J.-M. Geneste, R.L. Whear, J.-J. Delannoy, M. Katherine, R.G. Gunn, C. Clarkson, H. Plisson, P. Lee, F. Petchey, C. Rowe, B. Barker, L. Lamb, W. Miller, S. Hoerle, D. James, E. Boche, K. Aplin, I.J. McNiven, T. Richards, A. Fairbairn and J. Matthews 2011 Nawarla Gabarnmang, a 45,180±910 cal BP site in Jawoyn Country, southwest Arnhem Land plateau. Australian Archaeology 73:73–77. Davidson, I. 2013 Peopling the last new worlds: The first colonisation of Sahul and the Americas. Quaternary International 285:1–29. De Deckker, P. 2001 Late Quaternary cyclic aridity in tropical Australia. Palaeogeography, Palaeoclimatology, Palaeoecology 170:1–9. Delannoy, J.-J., B. David, J.-M. Geneste, M. Katherine, B. Barker, R.L. Whear and B. Gunn 2013 The social construction of and rockshelters: (France) and Nawarla Gabarnmang (Australia). Antiquity 87(335):12–29. Denham, T., R. Fullagar and L.M. Head 2009 Plant exploitation on Sahul: From colonisation to the emergence of regional specialisation during the Holocene. Quaternary International 202:29–40. Dickson, F.P. 1981 Australian Stone Hatchets: A Study in Design and Dynamics. Sydney: Academic Press. Dortch, C.E. 1977 Early and late stone industrial phases in Western Australia. In R.V.S. Wright (ed.), Stone Tools as Cultural Markers: Change, Evolution and Complexity, pp.104–132. Canberra: Australian Institute of Aboriginal Studies. Dortch, C.E. 1984 Devil's Lair: A Study in Prehistory. Perth: Western Australian Museum. Dortch, J. 2004 Palaeo-Environemental Change and the Persistence of Human Occupation in South-Western Australian Forests. BAR International Series 1288. Oxford: Archaeopress. Drennan, R.D. 2009 Statistics for Archaeologists: A Common Sense Approach. New York: Springer. Fitzsimmons, K.E., T.J. Cohen, P.P. Hesse, J.D. Jansen, G.C. Nanson, J.-H. May, T.T. Barrows, D. Haberlah, A. Hilgers, T. Kelly, J. Larsen, J. Lomax and P. Treble in press

63 Late Quaternary palaeoenvironmental change in the Australian drylands. Quaternary Science Reviews. Flenniken, J.J. and J.P. White 1985 Australian flaked stone tools: A technological perspective. Records of the South Australian Museum 36:131–151. Frankel, D. 1993 Pleistocene chronological structures and explanations: A challenge. In M.A. Smith, M. Spriggs and B. Fankhauser (eds), Sahul in Review: Pleistocene Archaeology in Australia, New Guinea and Island Melanesia, pp.24–33. Occasional Papers in Prehistory 24. Canberra: Department of Prehistory, Research School of Pacific Studies, The Australian National University. Frankel, D. 1995 The Australian transition: Real and perceived boundaries. Antiquity 69:649– 655. Geneste, J.-M., B. David, H. Plisson, C. Clarkson, J.-J. Delannoy, F. Petchey and R.L. Whear 2010 Earliest evidence for ground-edge axes: 35,400±410 cal BP from Jawoyn Country, Arnhem Land. Australian Archaeology 71:66–69. Geneste, J.-M., B. David, H. Plisson, J.-J. Delannoy and F. Petchey 2012 The origins of ground-edge axes: New findings from Nawarla Gabarnmang, Arnhem Land (Australia) and global implications for the evolution of fully modern humans. Cambridge Archaeological Journal 22(1):1–17. Gould, R. 1969 Puntutjarpa rockshelter: A reply to messrs. Glover and Lampert. Archaeology and Physical Anthropology in Oceania 4:229–237. Groube, L., J. Chappell, J. Muke and D. Price 1986 A 40,000 year-old human occupation site at Huon Peninsula, Papua New Guinea. Nature 324:453–455. Gunn, R.G., R.L. Whear and L.C. Douglas 2012 Dating the present at Nawarla Gabarnmang: Time and function in the art of a major Jawoyn rock art and occupation site in western Arnhem Land. Australian Archaeology 75:55–65. Haberle, S.G. and B. David 2004 Climates of change: Human dimensions of Holocene environmental change in low latitudes of the PEPII transect. Quaternary International 118–119:165–179. Hall, C. 2004 Evaluating prehistoric hunter-gatherer mobility, land use, and technological organisation strategies using minimal analytical nodule analysis. In C. Hall and M. Larson (eds), Aggregate Analysis in Chipped Stone, pp.139–155. Salt Lake City: University of Utah Press. Hiscock, P. 1981 Comments on the Use of Chipped Stone Artefacts as a Measure of 'Intensity of Site Usage'. Australian Archaeology 13:30–34. Hiscock, P. 1994 Technological responses to risk in Holocene Australia. Journal of World Prehistory 8:267–292. 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. Hiscock, P. 2007 Looking the other way. A materialist/technological approach to classifying tools and implements, cores and retouched flakes. In S. McPherron (ed.), Tools versus Cores? Alternative Approaches to Stone Tool Analysis, pp.198–222. Newcastle: Cambridge Scholars Publishing. Hiscock, P. 2008 Archaeology of Ancient Australia. London: Routledge. Hiscock, P. 2009 Reduction, recycling, and raw material procurement in western Arnhem Land, Australia. In B. Blades and B. Adams (eds), Lithic Materials and Paleolithic Societies, pp.78–93. Oxford: Wiley-Blackwell.

64 Hiscock, P. 2012 Comment: Dancing on pins: Tension between clever theory and material records in Australian archaeology. Australian Archaeology 74:25–26. Hiscock, P. and V. Attenbrow 2003 Early Australian implement variation: A reduction model. Journal of Archaeological Science 30:239–249. Hiscock, P. and V. Attenbrow 2011 Technology and technological change in eastern Australian, the example of Capertee 3. In B. Marwick and A. Mackay (eds), Keeping Your Edge: Recent Approaches to the Organisation of Stone Artefact Technology, pp.21–32. Oxford: Archaeopress. Hiscock, P. and C. Clarkson 2000 Analysing Australian stone artefacts: An agenda for the twenty-first century. Australian Archaeology 50:98–108. Hiscock, P. and L.A. Wallis 2005 Pleistocene settlement of deserts from an Australian perspective. In P. Veth, M. Smith and P. Hiscock (eds), Desert Peoples: Archaeological Perspectives, pp.34–57. Malden: Blackwell Publishing. Holdaway, S. 1995 Stone artefacts and the transition. Antiquity 69:784–797. Holdaway, S. and M. Douglass 2012 A twenty-first century archaeology of stone artifacts. Journal of Archaeological Method and Theory 19(1):101–131. Holdaway, S., J. Shiner and P. Fanning 2004 Hunter-gatherers and the archaeology of discard behavior: An analysis of surface stone artifacts from Sturt National Park, western New South Wales, Australia. Asian Perspectives 43(1):34–72. Holdaway, S. and N. Stern 2004 A Record in Stone: The Study of Australia's Flaked Stone Artefacts. Canberra: Aboriginal Studies Press. Johnson, B.J., G.H. Miller, M.L. Fogel, J.W. Magee, M.K. Gagan and A.R. Chivas 1999 65,000 years of vegetation change in Central Australia and the Australian summer monsoon. Science 284:1150–1152. Jones, R. and I. Johnson 1985 Deaf Adder Gorge: Lindner Site, Nauwalabila I. In R. Jones (ed.), Archaeological Research in Kakadu National Park, pp.165–227. Canberra: Australian National Parks and Wildlife Service. Kamminga, J. and H. Allen 1973 Report of the Archaeological Survey. Alligator Rivers Environmental Fact-Finding Study. Kelly, R. 1992 Mobility/sedentism: Concepts, archaeological measures and effects. Annual Review of Anthropology 21:43–66. Kelly, R. 1995 The Foraging Spectrum: Diversity in Hunter-Gatherer Lifeways. Washington: Smithsonian Institution Press. Kershaw, A.P. 1995 Environmental change in Greater Australia. Antiquity 69:656–675. Kessler, R., C. Beck and G. Jones 2009 Trash: The structure of Great Basin Paleoarchaic debitage assemblages in western North American. In B. Adams and B. Blades (eds), Lithic Materials and Paleolithic Societies, pp.144–159. West Sussex: Blackwell Publishing. Kuhn, S. 1994 A formal approach to the design and assembly of mobile toolkits. American Antiquity 59(3):426–442. Kuhn, S. 1995 Lithic Technology. Princeton: Princeton University Press. Langley, M.C., C. Clarkson and S. Ulm 2011 From small holes to grand narratives: The impact of taphonomy and sample size on the modernity debate in Australia and New Guinea. Journal of 61(2):197–208. Law, W.B., D.N. Cropper and F. Petchey 2010 Djadjiling Rockshelter: 35,000 14C years of Aboriginal occupation in the Pilbara, Western Australia. Australian Archaeology 70:68–71.

65 Luly, J.G. 1993 Holocene palaeoenvironments near Lake Tyrrell, semi-arid northwestern Victoria, Australia. Journal of Biogeography 20(6):587–598. Lycett, S.J. and M.I. Eren 2013 Levallois economics: An examination of ‘waste’ production in experimentally produced Levallois reduction sequences. Journal of Archaeological Science 40:2384–2392. MacDonald, D. 2008 The role of lithic raw material availability and quality in determining tool kit size, tool function, and degree of retouch: A case study from Skink Rockshelter (46NI445), West Virginia. In W. Andrefsky (ed.), Lithic Technology: Measures of Production, Use and Curation, pp.216–232. Cambridge: Cambridge University Press. Macgregor, O. 2005 Aburpt terminations and stone artefact reduction potential. In C. Clarkson and L. Lamb (eds), Lithics 'Down Under': Australian Perspectives on Stone Artefact Reduction, Use and Classification, pp.95–108. BAR International Series 1408. Oxford: Archaeopress. Macintosh, N.W.G. 1951 Archaeology of Tandandjal cave, southwest Arnhem Land. Oceania 21:178–204. 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. Marwick, B. 2002 Milly's cave: Evidence for human occupation of the inland Pilbara during the Last Glacial Maximum. Tempus 7:21–33. Matthews, J. 2012 Why little rocks Rock: Debitage attribute analysis at Geleji, a late Holocene rockshelter in Wardaman Country, N.T. Unpublished poster presented at the Australian Archaeological Association Annual Conference, Wollongong, Australia. Dec 12, 2012. Available online . McCarthy, F. 1951 Stone implements from Tandandjal Cave - an appendix. Oceania 21:205– 213. McCarthy, F. 1964 The archaeology of the Capertee Valley, NSW. Records of the Australian Museum 26(6):197–246. McGlone, M.S., A.P. Kershaw and V. Markgrad 1992 El Nino/Southern Osillation climatic variability in Australasian and South American palaeoenvironmental records. In H.F. Diaz and V. Markgraf (eds), El Nino: Historical and Paleoclimatic Aspects of the Southern Oscillation, pp.435–462. Cambridge: Cambridge University Press. Meehan, B. and R. Jones 2005 Stone tool use in a land without stone: Ethnographic notes from the Gu-jingarliya. In I. Macfarlane, M.-J. Mountain and R. Paton (eds), Many Exchanges: Archaeology, History, Community and the Work of Isabel McBryde, pp.147–170. Aboriginal History Monograph 11. Canberra: Aboriginal History. Moore, M.W. 2013 Simple stone flaking in Australasia: Patterns and implications. Quaternary International 285:140–149. Moore, M.W., T. Sutikna, Jatmiko, M.J. Morwood and A. Brumm 2009 Continuities in stone flaking technology at Liang Bua, Flores, Indonesia. Journal of Human Evolution 57:503–526. Morwood, M.J. and D.R. Hobbs 1995 Themes in the prehistory of tropical Australia. Antiquity 69:747–768.

66 Morwood, M.J. and P.J. Trezise 1989 Edge-ground axes in Pleistocene greater Australia: New evidence from SE Cape York Peninsula. Queensland Archaeological Reseach 6:77–90. Mountford, C.P. 1958 Aboriginal cave paintings at Sleisbeck northern Australia. Records of the South Australian Museum 13(2):147–155. Mulvaney, D.J. 1975 The Prehistory of Australia. (2nd). Ringwood: Penguin Books. Mulvaney, D.J. and J. Kamminga 1999 Prehistory of Australia. St Leonards: Allen and Unwin. Mulvaney, K. 2013 Iconic imagery: Pleistocene rock art development across northern Australia. Quaternary International 285:99–110. Nanson, G.C., T.J. East and R. Roberts 1993 Quaternary stratigraphy, geochronology and evolution of the Magela Creek catchment in the monsoon tropics of northern Australia. Sedimentary Geology 83:277–302. Nanson, G.C., D. Price and S. Short 1992 Wetting and drying of Australia over the past 300 ka. Geology 20:791–794. Nelson, M. 1991 The study of technological organisation. Archaeological Method and Theory 3:57–100. O'Connell, J.F. and J. Allen 1995 Human reactions to the Pleistocene-Holocene transition in Greater Australia: A summary. Antiquity 69:855–862. O'Connell, J.F. and J. Allen 2012 The restaurant at the end of the universe: Modelling the colonisation of Sahul. Australian Archaeology 74:5–17. O'Connell, J.F., J. Allen and K. Hawkes 2010 Pleistocene Sahul and the origins of seafaring. In A. Anderson, J.H. Barrett and K.V. Boyle (eds), The Global Origins and Development of Seafaring, pp.57–68. Cambridge: McDonald Institute for Archaeological Research. O'Connor, S. 1999 30,000 years of Aboriginal occupation: Kimberley, North West Australia. Terra Australis 14. Canberra: ANH Publications and The Centre for Archaeological Research, The Australian National University. Patterson, L.W. 1990 Characteristics of bifacial reduction flake-size distribution. American Antiquity 55(3):550–558. 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. Pecora, A.M. 2001 Chipped stone tool production strategies and lithic debris patterns. In W. Andrefsky (ed.), Lithic Debitage: Context, Form, Meaning, pp.173–190. Salt Lake City: University of Utah Press. Pelcin, A.W. 1997 The formation of flakes: The role of platform thickness and exterior platform angle in the production of flake initiations and terminations. Journal of Archaeological Science 24:1107–1113. Phagan, C. 1985 Lithic technology: Flake analysis. In R. MacNeish, A. Nelken-Terner, C. Phagan and R. Vierra (eds), Prehistory of the Ayacucho Basin, Peru, pp.233–281. Michigan: The University of Michigan Press. Porr, M. 2000 Signs of the times: A difference approach towards the origins of Lower Palaeolithic handaxes. Archaeological Review from Cambridge 17(1):19–32. Przywolnik, K. 2005 Long-term transitions in hunter-gatherers of coastal north western Australia. In P. Veth, M. Smith and P. Hiscock (eds), Desert Peoples: Archaeological Perspectives, pp.177–205. Malden: Blackwell Publishing. Rasic, J. and W. Andrefsky 2001 Alaskan blade cores and specialised components of mobile toolkits: Assessing design parameters and toolkit organisation through debitage

67 analysis. In W. Andrefsky (ed.), Lithic Debitage: Context, Form, Meaning, pp.61–79. Salt Lake City: University of Utah Press. Reeves, J.M., T.T. 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 and O.I. Members in press-a 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. Reeves, J.M., H.C. Bostock, L.K. Ayliffe, T.T. Barrows, P. De Deckker, L.S. Devriendt, G.B. Dunbar, R.N. Drysdale, K. Fitzsimmons, S. Lewis, H.V. McGregor, S.D. Mooney, P. Moss, G.C. Nanson, A. Purcell and S. van der Kaars in press-b Palaeoenvironmental change in tropical Australasia over the last 30,000 years a synthesis by the OZ- INTIMATE group. Quaternary Science Reviews. Reeves, J.M., A.R. Chivas, A. García, S. Holt, M.J.J. Couapel, B.G. Jones, D.I. Cendón and D. Fink 2008 The sedimentary record of palaeoenvironments and sea-level change in the Gulf of Carpentaria, Australia, through the last glacial cycle. Quaternary International 183(1):3–22. Rezek, Z., S. Lin, R. Iovita and H. Dibble 2011 The relative effects of core surface morpology on flake shape and other attributes. Journal of Archaeological Science 38(6):1349– 1359. Rinehart, N.R. 2008a Debitage analysis: Lithic reduction or lithic production. In C. Reith (ed.), Current Approaches to the Analysis and Interpretation of Small Lithic Sites in the Northeast, pp.63–73. New York: New York State Museum. Rinehart, N.R. 2008b Moving beyond the reduction stage in debitage analysis, with a little help from the pot sherd. North American Archaeologist 29(3–4):383–390. Root, M. 2004 Technological analysis of flake debris and the limitations of size-grade techniques. In C. Hall and M.L. Larson (eds), Aggregate Analysis in Chipped Stone, pp.65–94. Salt Lake City: University of Utah Press. Schrire, C. 1972 Ethno-archaeological models and subsistence behaviour in Arnhem Land. In D.L. Clarke (ed.), Models in Archaeology, pp.653–670. London: Methuen. Schrire, C. 1982 The Alligator Rivers: Prehistory and Ecology in Western Arnhem Land. Terra Australis 7. Canberra: Department of Prehistory, Research School of Pacific Studies, The Australian National University. Shott, M. 1986 Technological organization and settlement mobility: An ethnographic examination. Journal of Anthropological Research 42:15–51. Shott, M. 1994 Size and form in the analysis of flake debris: Review and recent approaches. Journal of Archaeological Method and Theory 1(1):69–110. Shott, M. 2004 Aggregate methods and the future of debris analysis. In C. Hall and M. Larson (eds), Aggregate Analysis in Chipped Stone, pp.211–228. Salt Lake City: University of Utah Press. Shulmeister, J. and B.G. Lees 1995 Pollen evidence from tropical Australia for the onset of an ENSO-dominated climate at c. 4000 BP. The Holocene 5:10–18. Slack, M. 2007 Between the Desert and the Gulf: Evolutionary Anthropology and Aboriginal Prehistory in the Riversleigh/Lawn Hill Region, Northern Australia. Unpublished PhD thesis, School of Philosophical and Historical Inquiry, The University of Sydney, Sydney,

68 Smith, G.M. 2011 Shifting stones and changing homes: Using toolstone ratios to consider relative occupation span in the northwestern Great Basin. Journal of Archaeological Science 38(2):461–469. Smith, M. 2006 Characterizing late Pleistocene and Holocene stone artefact assemblages from Puritjarra rockshelter: A long sequence from the Australian Desert. Records of the Australian Museum 58:371–410. Stern, N. 2009 The archaeological signature of behavioural modernity: A perspective from the southern periphery of the modern human range. In J.J. Shea and D. Lieberman (eds), Transitions in Prehistory: Essays in Honor of Ofer Bar-Yosef, pp.253–283. Oxford: Oxbow Books. Stiner, M.C. and S. Kuhn 1992 Subsistence, technology, and adaptive variation in Middle Paleolithic Italy. American Anthropologist 94(2):306–339. Sullivan, A.P. 2001 Holmes's Principle and beyond: The case for renewing Americanist debitage analysis. In W. Andrefsky (ed.), Lithic Debitage: Context, Form, Meaning, pp.192–206. Salt Lake City: University of Utah Press. Sullivan, A.P. and K.C. Rozen 1985 Debitage analysis and archaeological interpretation. American Antiquity 50(4):755–779. Surovell, T. 2009 Toward a Behavioral Ecology of Lithic Technology. Tucson: University of Arizona Press. Taçon, P. 1991 The power of stone: Symbolic aspects of stone use and tool development in western Arnhern Land, Australia. Antiquity 65:192–207. Tindale, N.B. 1957 Cultural succession in south-eastern Australian from late Pleistocene to the present. Records of the South Australian Museum 13(1):1–52. Torrence, R. 1989 Time, Energy and Stone Tools. Cambridge: Cambridge University Press. Ulm, S. 2006 Coastal Themes: An Archaeology of the Southern Curtis Coast, Queensland. Terra Australis 24. Canberra:: ANU EPress. Ulm, S. 2013 ‘Complexity’ and the Australian continental narrative: Themes in the archaeology of Holocene Australia. Quaternary International 285:182–192. Veth, P. 1993 Islands in the Interior: The Dynamics of Prehistoric Adaptations within the Arid Zone of Australia. Archaeological Series 3. Ann Arbor: International Monographs in Prehistory. Veth, P. 2005 Cycles of aridity and human mobility: Risk-minimization amongst late Pleistocene foragers of the western desert, Australia. In P. Veth, M. Smith and P. Hiscock (eds), Desert Peoples: Archaeological Perspectives, pp.100–115. Oxford: Blackwell Publishing. Veth, P., P. Hiscock and A. Williams 2011a Are tulas and ENSO linked in Australia? Australian Archaeology 72:7–14. Veth, P., M. Smith, J.M. Bowler, K.E. Fitzsimmons, A. Williams and P. Hiscock 2009 Excavations at Parnkupirti, Lake Gregory, Great Sandy Desert: OSL ages from occupation before the Last Glacial Maximum. Australian Archaeology 69:1–11. Veth, P., N. Stern, J. McDonald, J. Balme and I. Davidson 2011b The role of information exchange in the colonisation of Sahul. In R. Whallon, W.A. Lovis and R.K. Hitchcock (eds), Information and its Role in Hunter-Gatherer Bands, pp.203–220. Ideas, Debates and Perspectives 5. Los Angeles: Cotsen Institute of Archaeology Press. Ward, I., G.C. Nanson, L.M. Head, R. Fullagar, D. Price and D. Fink 2005 Late Quaternary landscape evolution in the Keep River region, northwestern Australia. Quaternary Science Reviews 24:1906–1922.

69 White, B. 2012 Minimum analytical nodules and lithic acitivities at site W2, Hunter Valley, New South Wales. Australian Archaeology 75:25–36. White, P.J. 2011 Backed artefacts: Useful socially and operationally. Australian Archaeology 72:67–75. Whittaker, J. 1994 Flintknapping: Making and Understanding Stone Tools. Austin: University of Texas Press. Williams, J. and W. Andrefsky 2011 Debitage variability among multiple flint knappers. Journal of Archaeological Science 38:865–872. Wyrwoll, K.-H. and G.H. Miller 2001 Initiation of the Australian summer monsoon 14,000 years ago. Quaternary International 83–85:119–128.

70 APPENDIX 1: DESCRIPTION OF DEBITAGE ATTRIBUTES

Description of the attributes used in debitage attribute analysis and the rationale behind their selection, only flakes with a platform were recorded using this method.

Attribute Explanation Rationale Reference Variable size attribute, recorded to nearest two A consistently accurate measurement that (Shott 1994:80) Weight decimal places. provides information on flake size. Percussion length, measured from striking Provides a replicable measure of overall size (Andrefsky 2005a:99) Length platform to distal end. and shape when combined with width. Measured halfway along the flake perpendicular Provides a replicable measure of overall size (Andrefsky 2005a:99) Width to length measurement. and shape when combined with length. Taken at the intersection of the length and width Provides a replicable measure of overall size (Andrefsky Thickness measurements from the dorsal and ventral faces. and shape when combined with other 2005a:101) measurements. Widest point taken at proximal and distal ends of When combined with other size variables (Clarkson and David Proximal and the flake. can indicate flake shape e.g. elongate, 1995:32) Distal Width converging, or diverging. Width measurement is taken across platform Characterisation of platform size is (Andrefsky 2005a:94; Platform width from one margin to the other. Thickness is important for understanding core Clarkson 2007:32; and thickness measured from the ring crack or center of morphology, possible platform constriction, Rezek et al. 2011) platform to the dorsal edge. and applied force. Recorded as bending or hertizan; or if the Strong indicator of the technique used e.g. a (Clarkson 2007:32; Initiation type platform is too crushed from initiation force the bending initiation commonly form on low- Clarkson and initiation is termed to be crushed. edge angled cores. O'Connor 2006:164) The main types recorded are; cortical, crushed, Indicative of the stage of reduction, (Clarkson 2007:32; dihedral, focalized, multiple conchoidal and technique and force used e.g. crushed Phagan 1985:235) Platform type single conchoidal (although combinations of the platforms can show excess force used on a above are possible e.g. multiple focalized). small platform, focalized platforms can indicate core preparation, and multiple

71 conchoidal platforms can show that previous flakes have been removed from the core platform surface. The types recorded are; overhang removal, The preparing of platforms provides a way (Whittaker 1994:101) grinding, faceting. In regards to overhang for the knapper to control the force imparted Platform removal I differentiate between minimal (<3 on the core or flake. preparation scars) and extensive (3 or more scars) to separate possible accidental damage and intentional preparation. The types recorded are; sheared, diffuse, This attribute can be indicative of force (Clarkson and Bulb attributes hertizan, or none. imparted and stone working technique. O'Connor 2006:172; Phagan 1985:273) The types recorded are; step, feather, hinge, This attribute can be indicative of force (Clarkson 2007:32; Flake termination outrépassé. imparted, stone working technique, and Macgregor 2005) success in flake removal. Dorsal scars were counted. This attribute indicates number of previous (Andrefsky Number of flake removals and can potentially show the 2005a:109) Dorsal Scars extent of reduction. The direction of previous removals is termed as These attributes are thought to vary with (Phagan 1985:279) follows; proximal, distal, left, and right; in any technology and reduction stage i.e. the more combination. If scars are coming from 3 intensively reduced the greater number of Direction and directions it is termed weakly radial and from >3 directions flake scars come from. Termination of directions it is radial. Terminations are recorded Dorsal Scars as for flake termination above and if a termination is not visible it is recorded as outline only. The angle of intersection between the platform This can be an important indicator of core (Andrefsky 2005a:91; and ventral face of the flake, recorded at the mid- surface morphology, but consistency in this 2009:74; Rezek et al. Platform angle point of the platform. measurement is questionable and whilst this 2011) was undertaken it is not considered to be wholly accurate.

72 APPENDIX 2: DESCRIPTION OF MANA ATTRIBUTES

Description of the attributes used for minimal analytical nodule analysis and the rationale behind their selection, these attributes are designed to separate the material types into minimal nodule groupings and seeks to highlight variation

Attribute Description Rationale The main types present were; chalcedony, chert, quartz Raw material has the potential to inform us about transport, Raw Material (crystal and milky), quartzite (local and exotic), silcrete, land-use and raw material provisioning. sandstone, and volcanic. Identification of the material as being local or exotic. This provides an indication of where the lithic materials used at the site are coming from; of critical importance is the Exotic/Local separation of local and exotic quartzite as the local variety likely originates from the shelter itself. The colour that covers the majority of the artefact, this Allows the consistent description and separation of different Primary colour was identified under a magnifying lamp using Munsell material types. colour descriptions for consistency. Secondary The colour of any features within the material, if there is Allows the consistent description and separation of different colour/s more than one they are listed in order of abundance. material types. Describes any patterning within the rock, if the material Allows the consistent description and separation of different has only one colour and no patterning it is termed as material types. Pattern being ‘unpatterned’. Examples of patterning description include; banded, fine lines, spots and lines, swirled/mixed. Yes/No – describes the presence or absence of grains This supports the designation of texture (see below) and Visible grains in the material. allows better separation of material types. 0.25 mm or 0.5 mm or 1 mm, based on the average Helps to separate artefacts of the same material type based on grain size visible. This was estimated and measured how coarse or fine the grains present are, particularly useful Grain size with calipers if ambiguous. for different varieties of quartzite and other granulite materials. Grain description Grains are described as being ‘fused’, when there is no Helps to separate artefacts of granulite materials and is a good

73 separation between the visible grains and ‘separated’ if indicator of differences in material formation. there is visible separation, this was determined with use of a hand-lens at 10x magnification. Most common texture types; Macrocrystalline (e.g. Texture is a key descriptive element for characterizing any quartz); Microcrystalline (e.g. /chalcedony); stone material (Hall 2004) and these are the most common Texture Granulite (e.g. quartzite, some silcretes); and texture descriptions drawn from geology. Porphyritic (silcretes with grain size variability). Types of weathering; This attribute describes whether the object has experienced Lightly = some surfaces are smoothed and sharp angles weathering processes in its history, and can separate identical Weathering and edges smoothed. materials that have experienced post-production degradation Heavily = all surfaces are smoothed, no sharp angles or or change. edges are present, rounded edges. Checked with a hand lens to confirm that the artefact is Similarly to weathering, staining can separate identical not simply dirty, stain must cover >10% of the artefact. materials; this attribute is connected to the nature of the Staining Most common form is black sediment stain; other types sediment the artefact was recovered from. are described in the comments section. Estimated in 10% increments, if there is no cortex this is Cortex is a strong indicator of artefact reduction, with the Cortex % expressed at 0%. presence of cortex often used to show primary or initial stages of reduction (e.g. Beck et al. 2002). Describes whether cortex is present on the dorsal or This attribute helps to further identify reduction stage, e.g. Cortex Location platform surface, or both. could be used to indicate whether the has been previously reduced or not. The types recorded are; rounded, flat/angular, or This attribute can be used to separate different nodules of the Cortex irregular. If the 'cortex' is more likely to be an old same material from each other. Description weathered surface, this is noted in the comments section of the database.

74 APPENDIX 3: SQUARE A ARTEFACT COUNTS

Total artefact counts for the Square A assemblage, broken down into the four main categories and volumetrically adjusted.

# # Raw Vol. Adj. Spit Volume # Cores # Broken Comp/Prox Retouched Total Total 1 4.5 63 6 4 98 171 38 2 5 82 2 1 116 201 40.2 3 6 100 8 199 307 51.16667

4 5 80 5 1 296 382 76.4 5 7.5 214 8 1 643 866 115.4667 6 7.5 140 3 489 632 84.26667

7 6.5 114 416 530 81.53846

8 6.5 177 5 1 403 586 90.15385 9 8 108 4 1 290 403 50.375 10 6.5 74 3 242 319 49.07692

11 6 65 1 174 240 40

12 5.5 43 1 157 201 36.54545

13 8 31 188 219 27.375

14 4 27 84 111 27.75

15 8.5 33 127 160 18.82353

16 5 6 31 37 7.4

17 7.5 21 100 121 16.13333

18 4 20 35 55 13.75

19 5 12 52 64 12.8

20 7 20 1 1 89 111 15.85714 21 6.5 24 76 100 15.38462

22 4.5 17 1 48 66 14.66667

23 5 10 1 65 76 15.2

24 6 17 57 74 12.33333

25 4 10 1 1 69 81 20.25 26 6.5 18 1 2 93 114 17.53846 27 4.5 24 1 1 92 118 26.22222 28 5 30 2 107 139 27.8

29 4 20 1 74 95 23.75

30 4.5 30 3 1 80 114 25.33333 31 4 34 103 137 34.25

32 4.5 16 66 82 18.22222

75 33 4 13 1 62 76 19

34 3 2 20 22 7.333333

35 1.5 2 9 11 7.333333

36 2 9 22 31 15.5

Total 1706 59 15 5272 7052

Breakdown of the Square A assemblage across different artefact types relative to the total artefact count for each spit.

Other Total Bifacial Unifacial Axe Retouched Bipolar Spit Core ground artefac Point Point Flake Flake Flake stone ts 1 4 1 3 1 3 2 201

2 1 2 248

3 2 1 8 379

4 1 1 1 4 101

5 1 4 2 2 253

6 1 1 2 173

7 135

8 1 1 4 205

9 1 1 1 1 2 127

10 2 1 87

11 1 76

12 1 53

13 42

14 1 30

15 41

16 1 11

17 2 30

18 24

19 1 3 14

20 1 1 28

21 24

22 1 1 19

23 1 13

24 20

25 1 1 17

26 2 1 26

27 1 1 118

76 28 2 199

29 1 126

30 1 1 3 149

31 147

32 117

33 1 107

34 40

35 17

36 46

Total 15 14 6 7 42 2 8 3443

77 APPENDIX 4: FURTHER RESULTS

Reduction Intensity

Figure 1 compares the average percentages of cortex. Sample size is extremely small with only 29 artefacts in the assemblage exhibiting any cortex.

Figure 1 Mean cortex percentage. Core preparation and flake removal strategy

Figure 2 presents percentages of different termination types. There is general statistical similarity between all periods. Period 2 is statistical distinctive in terms of proportion of feather to aberrant terminations with the lowest rates of aberrant terminations (1-2, χ2 = 3.694, p > 0.055; 2-3, χ2 = 7.830, p > 0.005; 2-4, χ2 = 12.920, p > 0.0005).

Figure 2 Percentage of different termination types. Sample sizes: Period 1 N=484; Period 2 N=122; Period 3 N=55; Period 4 N=40; Period 5 N=23.

78 Figure 3 presents mean elongation values for the different periods of occupation. Statistically all periods are very similar and flakes are generally parallel sided.

Figure 3 Bar graph showing mean elongation values. Sample sizes: Period 1 N=403; Period 2 N=110; Period 3 N=48; Period 4 N=32; Period 5 N=16. Objective piece and reduction stage

Figure 4 presents mean platform angle. Period 5 is significantly different from all other periods (5-1, t = -3.615, df = 988, p > 0.0005; 5-2, t = -6.984, df = 268, p > 0.0005; 5-3, t = - 5.780, df = 133, p > 0.0005; 5-4, t = -3.761, df = 131, p > 0.0005); Period 1 and 4 are very similar (t = 1.347, df = 1045, p > 0.178); and Period 2 and 3 are very similar (t = -0.137, df = 327, p > 0.890).

Figure 4 Mean exterior platform angle. Sample size: Period 1 N=952; Period 2 N=232; Period 3 N=97; Period 4 N=95; Period 5 N=38.

79 APPENDIX 5: DESCRIPTION OF MANA GROUPINGS

Detailed description for each MAN group; including the key attributes that divided them and a count of how many artefacts are in each group. Originally analysis of the MANA data resulted in an expanded range of nodule groups utilizing more of the attributes recording during MANA. However, the security of those groupings was not assured and thus the more conservative estimate of nodules, presented here, was used in the analysis and interpretations. Original groups that have been combined to form this final table are noted as ‘combined’ in the colour description.

MAN Source Colour Description Texture Count Chal1 Exotic Light pink combined Microcrystalline 2 Chal2 Exotic Weak red unpatterned Microcrystalline 2 Chal3 Exotic Red with clear sections Microcrystalline 1 Chal4 Exotic Very pale brown unpatterned Microcrystalline 1 Chal5 Exotic Pink unpatterned Microcrystalline 1 Light red with dark reddish brown Chal6 Exotic Microcrystalline 1 sections mixed Che1 Exotic Brown unpatterned Microcrystalline 2 Che2 Exotic Dark greyish brown unpatterned Microcrystalline 1 Che3 Exotic Dark reddish brown unpatterned Microcrystalline 1 Che4 Exotic Dark yellowish brown unpatterned Microcrystalline 2 Che5 Exotic Dusky red combined Microcrystalline 11 Che6 Exotic Gray combined Microcrystalline 2 Che7 Exotic Light brown combined Microcrystalline 8 Che8 Exotic Light gray unpatterned Microcrystalline 2 Che9 Exotic Light red combined Microcrystalline 18 Che10 Exotic Light reddish brown combined Microcrystalline 3 Che11 Exotic Pale red combined Microcrystalline 10 Che12 Exotic Pale yellow with pale red mixed Microcrystalline 1 Che13 Exotic Pink combined Microcrystalline 6 Che14 Exotic Pinkish white combined Microcrystalline 2 Che15 Exotic Red combined Microcrystalline 14 Che16 Exotic Reddish black unpatterned Microcrystalline 2 Che17 Exotic Reddish brown unpatterned Microcrystalline 2 Che18 Exotic Reddish yellow combined Microcrystalline 2 Che19 Exotic Very dark gray unpatterned Microcrystalline 1 Che20 Exotic Very pale brown combined Microcrystalline 23 Che21 Exotic Weak red combined Microcrystalline 36

80 Che22 Exotic White combined Microcrystalline 8 Che23 Exotic Yellow unpatterned Microcrystalline 4 Che24 Exotic Yellowish brown combined Microcrystalline 3 Pet1 Exotic Pale brown with red lines Foliated 1 Qtz1 Exotic White milky quartz Macrocrystalline 13 Qtz2 Exotic Clear crystal quartz Macrocrystalline 11 Granulite with fused Qtze1 Local Brown unpatterned 1 grains ~0.25mm Granulite with fused Qtze2 Exotic Clear with pink mixed 1 grains ~0.25mm Granulite with fused Qtze3 Exotic Dark greyish brown unpatterned 1 grains ~0.25mm Dark yellowish brown with black Granulite with fused Qtze4 Exotic 1 spots grains ~0.5mm Granulite with fused Qtze5 Exotic Dusky red unpatterned 18 grains ~0.25mm Granulite with fused Qtze6 Exotic Dusky red unpatterned 1 grains ~0.5mm Granulite with fused Qtze7 Local Gray unpatterned 3 grains ~0.5mm Granulite with fused Qtze8 Exotic Gray unpatterned 4 grains ~0.25mm Granulite with fused Qtze9 Exotic Gray combined 4 grains ~0.25mm Granulite with fused Qtze10 Local Gray with red and white mixed 1 grains ~0.5mm Granulite with fused Qtze11 Local Light brown combined 6 grains ~0.25mm Granulite with fused Qtze12 Local Light brown unpatterned 1 grains ~0.5mm Granulite with fused Qtze13 Local Light gray combined 11 grains ~0.25mm Granulite with fused Qtze14 Exotic Light gray unpatterned 1 grains ~0.5mm Granulite with fused Qtze15 Exotic Light pink combined 3 grains ~0.25mm Granulite with fused Qtze16 Exotic Light red combined 34 grains ~0.25mm Granulite with fused Qtze17 Exotic Light red combined 10 grains ~0.5mm Granulite with fused Qtze18 Exotic Light reddish brown unpatterned 12 grains ~0.25mm Granulite with fused Qtze19 Exotic Light reddish brown unpatterned 2 grains ~0.5mm Qtze20 Local Light yellowish combined Granulite with fused 6

81 grains ~0.25mm Granulite with fused Qtze21 Local Pale brown unpatterned 9 grains ~0.25mm Granulite with fused Qtze22 Local Pale brown unpatterned 1 grains ~0.5mm Granulite with fused Qtze23 Exotic Pale red combined 27 grains ~0.25mm Granulite with fused Qtze24 Local Pale yellow unpatterned 1 grains ~0.5mm Granulite with fused Qtze25 Exotic Pink unpatterned 8 grains ~0.5mm Granulite with fused Qtze26 Exotic Pink combined 18 grains ~0.25mm Granulite with fused Qtze27 Exotic Pinkish white unpatterned 4 grains ~0.5mm Granulite with fused Qtze28 Exotic Pinkish white combined 33 grains ~0.25mm Granulite with fused Qtze29 Exotic Red with white and black spots 1 grains ~0.5mm Granulite with fused Qtze30 Exotic Red combined 26 grains ~0.25mm Granulite with fused Qtze31 Exotic Reddish brown combined 7 grains ~0.5mm Granulite with fused Qtze32 Exotic Reddish brown combined 18 grains ~0.25mm Granulite with fused Qtze33 Exotic Reddish gray combined 4 grains ~0.25mm Granulite with fused Qtze34 Exotic Reddish yellow combined 8 grains ~0.25mm Granulite with fused Qtze35 Local Stained 21 grains ~0.25mm Granulite with fused Qtze36 Local Stained 2 grains ~0.5mm Granulite with fused Qtze37 Exotic Strong brown with weak red mixed 1 grains ~0.25mm Granulite with fused Qtze38 Exotic Very dark gray unpatterned 3 grains ~0.25mm Granulite with fused Qtze39 Local Very pale brown combined 10 grains ~1mm Granulite with fused Qtze40 Local Very pale brown combined 387 grains ~0.25mm Granulite with fused Qtze41 Local Very pale brown unpatterned 162 grains ~0.5mm Granulite with fused Qtze42 Local Very pale yellow unpatterned 1 grains ~0.5mm Qtze43 Exotic Weak red combined Granulite with fused 2

82 grains ~0.5mm Granulite with fused Qtze44 Exotic Weak red combined 58 grains ~0.25mm Granulite with fused Qtze45 Local White combined 152 grains ~0.5mm Granulite with fused Qtze46 Local White unpatterned 11 grains ~1mm Granulite with fused Qtze47 Local White combined 383 grains ~0.25mm Granulite with fused Qtze48 Exotic Yellow unpatterned 1 grains ~0.5mm Granulite with fused Qtze49 Exotic Yellow combined 14 grains ~0.25mm Granulite with fused Qtze50 Exotic Yellowish brown combined 2 grains ~0.25mm Granulite with Sil1 Exotic Light brown unpatterned separated grains 2 ~0.5mm Granulite with Sil2 Exotic Light gray combined separated grains 3 ~0.25mm Granulite with Sil3 Exotic Light gray with white bands separated grains 1 ~0.5mm Porphyritic with Sil4 Exotic Light pink combined separated grains 2 ~0.25mm Granulite with Sil5 Exotic Light red unpatterned separated grains 2 ~0.5mm Granulite with Sil6 Exotic Light red with white sections separated grains 1 ~0.25mm Porphyritic with Sil7 Exotic Light reddish brown separated grains 2 ~0.5mm Granulite with Sil8 Exotic Pale red unpatterned separated grains 3 ~0.5mm Porphyritic with Sil9 Exotic Pale yellow unpatterned separated grains 1 ~1mm Granulite with Sil10 Exotic Pink combined separated grains 2 ~0.5mm

83 Granulite with Sil11 Exotic Pinkish white unpatterned separated grains 8 ~0.25mm Granulite with Sil12 Exotic Pinkish white unpatterned separated grains 10 ~0.5mm Granulite with Sil13 Exotic Red unpatterned separated grains 1 ~0.25mm Granulite with Sil14 Exotic Red with black bands separated grains 1 ~0.5mm Granulite with Sil15 Exotic Reddish yellow unpatterned separated grains 1 ~0.25mm Granulite with SIl16 Exotic Very pale brown unpatterned separated grains 1 ~0.5mm Granulite with Sil17 Exotic Very pale yellow unpatterned separated grains 3 ~0.25mm Granulite with Sil18 Exotic Weak red unpatterned separated grains 1 ~0.25mm Granulite with Sil19 Exotic Weak red combined separated grains 2 ~0.5mm Granulite with Sil20 Exotic White combined separated grains 5 ~0.25mm Granulite with Sil21 Exotic White combined separated grains 5 ~0.5mm Vol1 Exotic Dark greyish brown 1

Vol2 Exotic Light gray 1

Vol3 Exotic Light greenish gray 2

Vol4 Exotic Light olive brown 2

Vol5 Exotic Pale brown 1

Vol6 Exotic Stained 1

Vol7 Exotic Very dark gray 3

Vol8 Exotic Very pale brown 2

84 APPENDIX 6: ANALYTICAL NODULES PER SPIT

Breakdown of the number of analytical nodules identified for each excavation unit and the percentage of local to exotic materials for each.

# # # # Richness # Local # Exotic % % Spit Local Exotic MAN Artefact Ratio Artefacts Artefacts Local Exotic AN AN 1 32 76 0.42 9 49 23 27 64.47 35.53 2 21 81 0.26 6 57 16 24 70.37 29.63 3 32 107 0.3 6 69 26 38 64.49 35.51 4 16 87 0.18 3 59 13 28 67.82 32.18 5 40 220 0.18 9 135 31 85 61.36 38.64 6 33 143 0.23 6 88 27 55 61.54 38.46 7 26 110 0.24 5 69 21 41 62.73 37.27 8 41 182 0.23 7 110 34 72 60.44 39.56 9 31 113 0.27 8 71 23 42 62.83 37.17 10 28 77 0.36 6 44 22 33 57.14 42.86 11 28 67 0.42 5 29 23 38 43.28 56.72 12 17 44 0.39 4 22 13 22 50 50 13 18 31 0.58 4 13 14 18 41.94 58.06 14 17 27 0.63 4 13 13 14 48.15 51.85 15 13 32 0.41 5 24 8 8 75 25 16 3 6 0.5 3 6 100 0.00

17 12 19 0.63 5 12 7 7 63.16 36.84 18 10 30 0.33 5 14 5 6 46.67 20.00 19 9 13 0.69 3 7 6 6 53.85 46.15 20 6 23 0.26 2 17 4 6 73.91 26.09 21 7 24 0.29 3 20 4 4 83.33 16.67 22 5 18 0.28 2 15 3 3 83.33 16.67 23 8 11 0.73 2 5 6 6 45.45 54.55 24 6 18 0.33 2 12 4 4 66.67 22.22 25 4 11 0.36 2 9 2 2 81.82 18.18 26 6 21 0.29 3 17 3 4 80.95 19.05 27 5 26 0.19 3 23 2 3 88.46 11.54 28 11 32 0.34 4 24 7 8 75.00 25.00 29 7 21 0.33 4 18 3 3 85.71 14.29 30 6 31 0.19 2 26 4 5 83.87 16.13 31 9 34 0.26 3 27 6 7 79.41 20.59 32 2 15 0.13 2 15 100 0.00

33 2 14 0.14 2 14 100 0.00

85 34 2 2 1 2 2 100 0.00

35 2 2 1 2 2 100 0.00

36 2 8 0.25 2 8 100 0.00

86 APPENDIX 7: ARTEFACT ILLUSTRATIONS

87

88 89 90

91