<<

Technological Change in the Early Middle : The Onset of the Middle Stone at 1, , South

by

Jayne Wilkins

A thesis submitted in conformity with the requirements for the degree of Doctor of Philosophy Department of University of Toronto

© Copyright by Jayne Wilkins 2013

Technological Change in the Early Middle Pleistocene: The Onset of the Middle at Kathu Pan 1, Northern Cape,

Jayne Wilkins

Doctor of Philosophy

Department of Anthropology University of Toronto

2013 Abstract

This dissertation describes the technological behaviors represented by the ~500-thousand-- old stratum 4a assemblage from Kathu Pan 1 (KP1), Northern Cape, South Africa, and situates new evidence from this site into evolutionary context. The findings highlight the significance of the early Middle Pleistocene in Africa for understanding behavioral in later .

The stratum 4a assemblage at KP1 represents a mainly flake and -based that employed multiple strategies to produce blanks that were retouched into a variety of forms, including unifacially retouched points. Diverse core reduction strategies at KP1 suggests that KP1 hominins were flexible to the demands of local raw materials, consistent with increased degrees of ‘behavioral variability’ and adaptability.

Several lines of evidence indicate that the KP1 points were used as tips. Points from sites ~300 thousand ago (ka) and younger were often used as tips, and evidence for this behavior can now be pushed back to ~500 ka, with important implications for cognition and social behavior among early Middle Pleistocene hominins.

Raw materials in the KP1 assemblage were acquired from multiple local sources. Based on comparisons with a sample from the underlying stratum 4b assemblage, the stratum 4a assemblage does not exhibit major changes in the kinds or quality of raw material exploited; thus, the technological changes represented by the stratum 4a assemblage are not explained by changes in raw material.

ii

New evidence from KP1 poses problems for current models that link the appearance of to speciation and dispersion ~300 ka. Middle Stone Age technologies appear in the African archaeological record by ~500 ka. The new timing for the origins of Middle Stone Age technologies provides a parsimonious explanation for technological similarities between the lithic assemblages of and modern Homo sapiens, who share a common ancestor in the early Middle Pleistocene. Limits imposed by the of the African archaeological record and chronometric analyses may explain why the antiquity of these technological changes was not previously recognized.

iii

Acknowledgments

I have many people to thank for helping me with the completion of this dissertation. My supervisor, Michael Chazan, has been an excellent mentor, always available for engaged conversations about interpreting the record that often urged me to step back and appreciate differing perspectives on an issue. I also appreciate that Michael encouraged me to explore the questions and problems that interest me the most. I am grateful for the guidance I received from my committee members Max Friesen and Genevieve Dewar. The comments from my external reviewer, Alison Brooks, were extremely helpful and I am honored that she participated in my defense. I would also like to thank my additional defense participants, Susan Pfieffer and Ed Swenson for additional insight on my findings.

Peter Beaumont deserves special acknowledgement for pioneering research at Kathu Pan and for ensuring that all the materials from his excavations were carefully provenienced and curated for study. I am privileged to be able to build upon the foundation he established.

I also owe thanks to the current team investigating Kathu Pan 1, including Liora Horwitz, Naomi Porat, and Rainer Grün, whose pivotal research provided the basis for my dissertation.

David Morris at the McGregor Museum was instrumental in organizing access to the Kathu Pan 1 collection, facilitating export of the points, and helping with the logistics of field research.

Research collaborators Ben Schoville and Kyle Brown were key to the success of this project, particularly the functional analysis of the KP1 points. I would like to thank them for their research contributions, but more importantly for being excellent role models; their enthusiasm for Stone Age science is infectious. I look forward to much collaboration in the future.

The guidance I received while conducting my Master’s research was instrumental in helping me gain skills required for conducting this dissertation project, and for that I am mostly indebted to Kathy Kuman. It is also because of my first field school under the direction of Kathy Kuman and Luca Pollarolo that I decided to pursue graduate studies in the South African

iv

Stone Age. I had my very first and very memorable Palaeolithic course with Mary McDonald at the University of Calgary. Part of my success in the project to the encouragement, guidance, and inspiration I received from Mary, Kathy, and Luca during my undergrad and MA programs.

I would also like to thank Vincent Dinku, Jane Joubert, and Chantel Wilson for assisting me with my research of the collections and Koot Msawula, Mark McGranaghan, Hilary Duke, Sarah Ranlett, Anna Phillips, and Ian Watts for helping me with raw material surveys. Thank you to Ed Swenson for granting me access to his microscope for viewing my thin- sections, and to Heather Miller and Greg Braun for assisting me with the saw. My office mates and good friends Danielle Macdonald and Matthew Walls helped me work through a lot of these ideas and I will miss our ‘work-a-thons’ in the ‘power office’. Danielle has always been exceptionally supportive and I treasure the friendship we developed during my in Toronto.

This research was primarily supported by a Social Sciences and Humanities Research Council (SSHRC) Joseph-Armand Bombardier Canada Graduate Scholarship, but I am also grateful for funding from other sources, including the Ontario Graduate Scholarship program, the Department of Anthropology, and the University of Toronto.

Finally, I would like to express appreciation to Scott, for his love and support, even when distance and stress challenged us, and to my Mom, who always in me and is always there to back me. I dedicate this dissertation to my father, Raymond Wilkins, who I know would be proud.

v

Table of Contents

Acknowledgments...... iv Table of Contents ...... vi List of Tables ...... vii List of Figures ...... x

1 Introduction ...... 1

2 Hominin Evolution in the Middle Pleistocene ...... 5 2.1 The African Stone Age ...... 5 2.2 Morphological Evolution in the Middle Pleistocene ...... 8 2.3 The Significance of Technological and Morphological Change in the Middle Pleistocene ...... 14 2.4 Summary ...... 32

3 Technological Change in the Middle Pleistocene ...... 34 3.1 Chronology of the ESA-MSA ‘transition’ ...... 34 3.2 Defining and Evaluating the Fauresmith Industry ...... 45 3.3 Blade Production in the Middle Pleistocene ...... 62 3.4 MSA Points and Point Function ...... 69 3.5 Raw Material Foraging in the ESA and MSA ...... 75 3.6 Summary ...... 80

4 Kathu Pan 1 (KP1) ...... 82 4.1 Geological and ecological setting ...... 82 4.2 Previous Research at KP1 ...... 84 4.3 Lithic Collection Catalogue and Documentation ...... 91 4.4 A consideration of site formation processes ...... 98 4.5 Summary ...... 102

5 Technological Analysis of KP1 Stratum 4a Lithic Assemblage ...... 106 5.1 Methods...... 106 5.2 Results ...... 110 5.3 Discussion ...... 134

6 Functional Analysis of Stratum 4a KP1 Points ...... 142 6.1 Methods...... 143 6.2 Results ...... 158 6.3 Discussion ...... 185

7 Raw Material Analysis ...... 193 7.1 Petrographic Identification of Lithic Raw Material Types ...... 193 7.2 Stratum 4a assemblage composition ...... 198 7.3 Potential Sources of Raw Material ...... 200

vi

7.4 Raw Material Selection at KP1 ...... 216 7.5 Discussion ...... 227

8 Discussion...... 232 8.1 Summary of results ...... 232 8.2 Synthesis and Interpretation ...... 232 8.3 Future Directions for Research ...... 252 8.4 Conclusion ...... 254

References ...... 256

vii

List of Tables

Table 1 African H. heidelbergensis , chronologies, and archaeological associations ...... 12 Table 2 Shea’s (in press) new paleolithic framework with Modes A-I...... 19 Table 3 Summary of published Fauresmith-designated assemblages...... 48 Table 4 Summary of MSA blade core reduction strategies during MIS 5 and 4 ()...... 65 Table 5 Types of evidence used to determine whether MSA points were used as hunting implements ...... 70 Table 6. Summary of Kathu Pan sites based on Beaumont (1990b)...... 85 Table 7 Summary of KP1 stratigraphy...... 86 Table 8 Summary of lithic contents and excavated levels by square in the KP1 lithic collection that is currently housed at the McGregor Museum, Kimberley...... 93 Table 9 Analyzed 4a assemblage...... 107 Table 10 Attributes recorded for the subsample from each technical category...... 109 Table 11 Bulb and platform analysis...... 110 Table 12 Assemblage composition of analyzed sample...... 112 Table 13 Summary of stratum 4a blade metrics for all blades and B1 blades (produced during optimal phase of ) from squares F23 and F21 (n=513)...... 116 Table 14 Summary of blade technical category frequencies for KP1 stratum 4a (n=972)...... 117 Table 15 Summary of core types, counts, and frequencies from KP1 stratum 4a, squares F23, F21, C23, and C21...... 122 Table 16 Summary of flake technical category frequencies for KP1 stratum 4a (n=972)...... 126 Table 17 Summary of stratum 4a Levallois flake metrics for all Levallois flakes from squares F23 and F21 (n=113)...... 127 Table 18 Summary of retouch pieces types, counts, and frequencies from KP1 stratum 4a squares F23, F21, C21, C23...... 130 Table 19 Percent frequency of retouched pieces manufactured on flake, blade, and B1 blade blanks in square F23...... 131 Table 20 Percent frequencies of bulb and platform attributes at KP1 ...... 132 Table 21 Summary of LCT metrics...... 134 Table 22 Comparison of KP1 blade production to published results for the Kapthurin Formation (Johnson and McBrearty, 2010) and the Amudian industry as represented at Qesem , (Barkai et al., 2009; Shimelmitz, 2009; Shimelmitz et al. 2011)...... 137 Table 23 Point sample for functional analysis...... 145 Table 24 Types of edge fractures identified on KP1 points...... 149 Table 25 Summary of diagnostic impact fracture counts and frequencies from experiments reported by Fischer et al. (1984), Pargeter (2011), and Lombard et al. (2004)...... 155 Table 26 PED frequency on unmodified (non-retouched edges) of KP1 points (n=106)...... 159 Table 27 Post-patination damage frequency on KP1 points (n=106)...... 162 Table 28 KS tests comparing post-patination damage distributions between all edges...... 165 Table 29 KS tests comparing post-patination distribution on the KP1 points to a “random” (equal probability) distribution...... 165 Table 30 KS tests comparing PED cumulative distributions on KP1 points to post-patination damage cumulative distributions on the equivalent edge...... 165

viii

Table 31 Damage frequency on experimental spear tips (n=32)...... 169 Table 32 Summary of expectations based on previous work ( et al., 2007; Schoville, 2010; Schoville and Brown, 2010) and KP1 results...... 175 Table 33 Counts and frequencies of diagnostic impact fractures on the KP1 points and point fragments (n=210)...... 177 Table 34 Counts and frequencies of diagnostic impact fractures on experimental spear tips (n=32)...... 177 Table 35 Summary of diagnostic impact fracture frequencies in the literature...... 179 Table 36 Summary of KP1 point metrics...... 181 Table 37 Experimental point characteristics and results of experimental trials...... 184 Table 38 Petrographic identification of raw material types in the KP1 assemblage...... 197 Table 39 Raw material type frequencies in the KP1 stratum 4a assemblage...... 199 Table 40 Cortex frequencies by raw material type in KP1 stratum 4a...... 199 Table 41 Banded ironstone type frequencies in KP1 stratum 4a...... 201 Table 42 Frequency of different banded ironstone structure and flaw types in the KP1 stratum 4a assemblage...... 203 Table 43 Geological formations near KP1, their location relative to KP1, and macroscopic and microscopic characteristics ...... 204 Table 44 Raw material type frequencies at six sampled secondary source location in three zones ...... 215 Table 45 Frequency of different banded ironstone structure and flaw types at each of the six secondary source sampling locations...... 218 Table 46 Comparison of percent frequency of raw material types for blades and Levallois flakes in stratum 4a assemblage at KP1...... 222 Table 47 Comparison of frequency of banded ironstone structure types between blades and Levallois flakes in stratum 4a KP1 assemblage...... 222 Table 48 Comparison of flaw type frequencies between blades and Levallois flakes in the stratum 4a assemblage of KP1...... 222 Table 49 Comparison of raw material type frequencies in stratum 4b sample compared to Stratum 4a sample...... 226 Table 50 Frequency of different banded ironstone structure and flaw types in the KP1 stratum 4b assemblage...... 226

ix

List of Figures

Figure 1 Summary of chronology for three types of Middle Pleistocene assemblages...... 46 Figure 2 Map of Fauresmith-designated sites mentioned in text...... 52 Figure 3 Map showing location of Kathu Pan and KP1...... 83 Figure 4 Stratigraphic profile of KP1 from 2004 investigations (Porat et al., 2010) showing chronometric age estimates and methods (looking north)...... 90 Figure 5 Three-dimensional reconstruction of KP1 excavations showing lithic density for 20 cm spits (kg/m3)...... 94 Figure 6 Correlation of samples, level, and stratum...... 95 Figure 7 Lithic artifact type frequencies for strata 3, 4a, and 4b...... 96 Figure 8 Size distribution of lithic artifacts...... 100 Figure 9 Example weathering states of KP1 lithic artifacts...... 102 Figure 10 Frequency of weathering states for different artifact types in stratum 4a...... 103 Figure 11 Frequency of weathering states for different levels within stratum 4a...... 104 Figure 12 Photographs of selected stratum 4a KP1 lithic artifacts...... 113 Figure 13 KP1 stratum 4a blades and proximal blade fragments from the optimal phase of debitage (BI)...... 114 Figure 14 Other blade types...... 115 Figure 15 Histogram of blade width (mm) from squares F23 and F21 of KP1 (n=511)...... 116 Figure 16 Percent frequency of blades (n=972) with cortex vs. complete flakes and proximal flake fragments (n=1544) with cortex...... 119 Figure 17 Bulb and platform attributes of KP1 blades...... 121 Figure 18 Blade cores at KP1...... 123 Figure 19 KP1 blade core organization...... 124 Figure 20 KP1 stratum 4a Levallois cores...... 125 Figure 21 Box plot of core size for KP1 stratum 4a assemblage...... 129 Figure 22 Vertical distribution of artifact types within stratum 4a, square F23 (n=1936). Dating samples (see Chapter 4) were recovered from ~80-110 cm below the top of stratum 3...... 135 Figure 23 Examples of KP1 complete retouched points...... 144 Figure 24 Examples of other KP1 point types...... 145 Figure 25 Edge damage distribution quantification methods...... 148 Figure 26 Experimental methods...... 153 Figure 27 Macrofracture type after Fischer et al. (1984)...... 155 Figure 28 Scatterplot of point length and width showing ‘optimal zone’ after Shea et al. (2001)...... 158 Figure 29 Edge damage frequency on KP1 points...... 160 Figure 30 Edge damage distribution on KP1 points, ...... 163 Figure 31 Cumulative percent frequency (using data from Figure 30B) of post-patination damage on the KP1 points compared to a “random” (equal probability) distribution...... 166 Figure 32 PED vs. post-patination damage cumulative distributions comparing equivalent edges (using data from Figure 30A and B)...... 167 Figure 33 PED ventral left vs. ventral right cumulative distributions (using data from Figure 30A)...... 169

x

Figure 34 Edge damage frequency on experimental spear tips ...... 171 Figure 35 Comparison of dorsal and ventral damage edge damage frequencies on the experimental spear tips ...... 171 Figure 36 Edge damage distribution on experimental spear tips...... 172 Figure 37 Examples of diagnostic impact fractures on KP1 points...... 176 Figure 38 Examples of basal modifications on KP1 points...... 180 Figure 39 Scatterplot of length and width of the KP1 points plotted with experimental data from Shea et al. (2001)...... 182 Figure 40 Comparison of KP1 point size to MSA point size...... 183 Figure 41 Scatterplot of length and widths of the experimental points plotted with the ‘optimal zone’ of Shea et al. (2001)...... 186 Figure 42 Banded ironstone KP1 stratum 4a archaeological samples (sectioned flake fragments) and thin sections...... 195 Figure 43 Other raw material types KP1 stratum 4a archaeological hand samples (sectioned flake fragments) and thin sections...... 196 Figure 44 Examples of banded ironstone structure types...... 202 Figure 45 Examples of banded ironstone flaw types...... 202 Figure 46 Map of primary sources of raw material in the Kathu region showing thin section sample locations...... 205 Figure 47 Geological samples and thin sections of primary and secondary sources of banded ironstone...... 206 Figure 48 Geological samples and thin sections of primary and secondary sources of other raw material types...... 207 Figure 49 Secondary sources of raw material in the Kathu region showing thin section sample locations and quantitative survey locations...... 208 Figure 50 Secondary source zones. See text for discussion...... 213 Figure 51 Methods for quantification of secondary sources of raw material...... 214 Figure 52 Comparison of raw material type frequencies in KP1 stratum 4a to secondary sources in Zone A and Zone B...... 217 Figure 53 Comparison of structure and flaw type frequencies in KP1 stratum 4a to secondary sources in Zone A and Zone B...... 219 Figure 54 Comparison of KP1 stratum 4a artifact length (complete pieces) and maximum length of cobbles at secondary sources for each raw material type...... 221 Figure 55 Definitions for direction of flaking axis with respect to banding...... 223 Figure 56 Comparison of flaking direction with respect to banding for different artifact types in KP1 stratum 4a assemblage...... 224 Figure 57 Comparison of flaking direction with respect to banding between strata 4a and 4b. . 227

xi

1

1 Introduction

This dissertation considers the significance of technological change in the early Middle Pleistocene at the of Kathu Pan 1 (KP1), Northern Cape Province, South Africa. The research forms part of the ongoing larger investigation of Stone Age archaeology in the Northern Cape directed by Michael Chazan at the University of Toronto, which also includes new investigations at the -known site of (Avery, 2007; Berna et al., 2012; Chazan, in press; Chazan et al., 2012a; Chazan et al., in press; Chazan and Horwitz, 2009; Chazan and Horwitz, in press; Chazan et al., 2008; Jacobson et al., in press; Matmon et al., 2012; Rüther et al., 2009). Hominin occupation in the Hills region of the Northern Cape began ~2 Million years ago (Chazan et al., 2012a; Chazan et al., 2008; Matmon et al., 2012) and there is a rich archaeological record spanning the Acheulean, Middle Stone Age, Later Stone Age, and the historical period.

Most of what is currently known about the Pleistocene archaeology of this part of the Northern Cape is the result of decades of extensive field research under the direction of Peter Beaumont at the McGregor Museum, Kimberley. Beaumont’s legacy is best appreciated when entering the store room at the McGregor that houses what could amount to millions of artifacts from dozens of sites. Several sites in the Northern Cape were featured on the 1990 and 2004 Southern African Association of Archaeologists conference tours, and with only a few exceptions, the ‘guidebooks’ prepared for these tours (Beaumont and Morris, 1990; Morris and Beaumont, 2004) are the only written record of Beaumont’s excavations and findings. The archaeological sites at Kathu Pan were investigated and excavated over a period of twelve years and six field seasons between 1978 and 1990. For Beaumont and others who examined the Kathu Pan and fauna, the sites there stood as a record of large-scale paleoenvironmental changes that occurred through the Pleistocene and (Beaumont, 1990b; Beaumont, 2004a; Butzer, 1984; Klein, 1988). However, prior to the study described here, the lithic collection of Kathu Pan had not been described beyond basic industrial assignments (Beaumont, 1990b; Beaumont, 2004a) and a preliminary analysis based on a sample of 217 flakes and blades from KP1 (Porat et al., 2010).

KP1 was subject to recent investigations by Chazan’s team, who re-exposed the stratigraphic profile to collect samples for chronometric dating (Porat et al., 2010). Optically stimulated

2 luminescence (OSL) and combined uranium-/electron spin resonance (U-series/ESR) analyses securely situate the stratum 4a assemblage from KP1 within the early Middle Pleistocene ~500 thousand years ago (ka). The stratum 4a assemblage at KP1 is assigned to the ‘Fauresmith Industry’ based on the co-occurrence of handaxes, Levallois , blades, and points (Beaumont, 1990b; Beaumont, 2004a; Porat et al., 2010). The Fauresmith is variably considered a Late Acheulean, early Middle Stone Age, or ‘transitional’ industry. As an early assemblage with Levallois technology, blades, and points, KP1 has the potential to shed light on the appearance of new lithic technologies in the early Middle Pleistocene, as well as their character, purpose, and relationship to lithic raw material characteristics, distribution, and abundance.

The shift from Acheulean to Middle Stone Age technologies in Africa is an important topic that requires further investigation. The origin of the Middle Stone Age is a crucial time period because of the role of the Middle Stone Age in accounts of hominin speciation and dispersion, the origins and evolution of behavioral and variability, the development of sociality and encephalization, and advances in cognitive facilities related to long-term planning, problem solving, and language (e.g. Ambrose, 2001; Ambrose, 2010; Foley and Gamble, 2009; Foley and Lahr, 2003; Foley and Lahr, 1997; Gamble et al., 2011; McBrearty and Brooks, 2000; Shea, 2011b; Wynn and Coolidge, 2011). Thus far, the details of the origins of Middle Stone Age technologies have remained ambiguous, because the archaeological record for this time period is patchy and there are few assemblages that can be dated using radiometric techniques.

My approach to the Fauresmith-designated stratum 4a assemblage at KP1 has three components; (1) the chaînes opératoires of flake and blade production, (2) the function of unifacially retouched and convergent points, and (3) the lithic raw material foraging strategies. This dissertation tests the hypotheses that new technological adaptations are represented in the ~500 thousand year old assemblage at KP1, including systematic blade production, diversity in lithic strategies, the production of unifacially retouched points and their use as hafted spear tips, and that these new technological adaptations occurred independently from differences in raw material characteristics, abundance, and distribution.

My research was conducted over four years and included three field seasons in the Northern Cape. The first field season in 2009 consisted of a pilot study of the KP1 lithic material and

3 preliminary survey for characterizing potential geological sources of raw material. In 2010, twelve weeks were spent in Kimberly analyzing the lithic assemblage at the McGregor Museum. Two employees with experience working on the McGregor museum collections assisted me with the cataloguing process. This same season, geological samples in and around the Kuruman Hills were collected for thin section. The sample of KP1 points used for the functional analysis were exported to the University of Toronto and analyzed there over a period of 6 months between September 2010 and February 2011. An experiment using a crossbow and reproductions of hafted were conducted as part of the functional analysis in July 2011 in Mossel Bay, South Africa. July-August 2011, an additional 6 weeks were spent at the McGregor Museum to complete analysis of the lithic assemblage. During this season, I also conducted the quantitative sampling of secondary sources of lithic raw material in and near the Kuruman Hills, which involved a team of three undergraduate and graduate students that I directed. The experimental spear tips points and geological thin sections were analyzed in Toronto from September 2011- January 2012.

In Chapter 2, I provide the evolutionary context for interpreting technological change in the Middle Pleistocene. Chapter 3 presents the current chronological situation for the archaeology of the Middle Pleistocene, followed by a detailed consideration of Middle and Late Pleistocene . I focus specifically on three aspects of technological behavior that are tied to the definition of the Middle Stone Age –blade production, hafted hunting , and raw material selection and transport. In Chapter 4, I present the necessary contextual background for the site of KP1, including stratigraphy, chronology, and on overview of the archaeological sequence, as well as a consideration of the general state of the lithic assemblage and its ability to inform us of technological behaviors in the early Middle Pleistocene. The following three chapters contain the bulk of my original research, which consists of a technological or chaîne opératoire analysis (Chapter 5), a functional analysis of the KP1 points (Chapter 6), and an examination and classification of the raw materials used for (Chapter 7). Chapter 8 concludes by asserting that current models linking expansion, Middle Stone Age technologies, and hafted hunting weapons to H. helmei or late ‘archaic Homo sapiens’ ~300 ka are out of sync with the new evidence from KP1. Rather, the onset of the Middle Stone Age in Africa appears to be associated with or and the increase in brain size associated with this group of hominins. The new data from KP1 have

4 important implications for how we interpret technological change and variability during the Pleistocene in the context of human origins and evolution.

5

2 Hominin Evolution in the Middle Pleistocene

The Middle Pleistocene (781 -126 ka) was witness to significant changes in hominin behavior and morphology. In Africa during this period, the archaeology shifts from the Earlier to the Middle Stone Age (~350-300 ka) and there may have been three speciation events represented by the first appearances of Homo rhodesiensis (or H. heidelbergensis, see discussion below), H. helmei, and H. sapiens. The Middle Stone Age (MSA) persists well into the Late Pleistocene until ~40 ka, and this latter period of the MSA has been the focus of relatively intense research in the past decade because of the role it plays in questions about early modern and the origins of modern human behavior. The nature of the MSA during the Middle Pleistocene remains largely unknown, in large part because of the scarcity of known archaeological sites from this period, especially on the African . The earlier half of the Middle Pleistocene (>350 ka) is generally associated with the Earlier Stone Age (ESA), even though there are only a handful of known sites in Africa dated to this period. Despite the paucity of robust archaeological research on the African Middle Pleistocene, this period has played a pivotal role in models for the behavioral and morphological evolution of the genus Homo. In this chapter, I summarize what is known from the archaeological and record of the African Middle Pleistocene and review hypotheses that have been put forth regarding the significance of the behavioral and morphological changes during this time.

2.1 The African Stone Age

2.1.1 Overview

The African Stone Age represents nearly 2.6 million years of hominin evolution. It is generally divided into three main periods: the ESA, MSA, and Later Stone Age (LSA). The ESA of Africa has two components, the and the Acheulean. The Oldowan represents the earliest known evidence for the manufacture of stone beginning 2.6 Ma and is characterized by an expedient technology focused on flake production (Ambrose, 2001). The first appearance of the Oldowan is roughly coeval with the first appearance of the genus Homo.

The Acheulean is defined by the presence of large bifacial tools, such as handaxes, cleavers, and picks (Clark, 1970; Goodwin and Van Riet Lowe, 1929). Its earliest occurrence is dated to 1.8 Ma in East Africa (Asfaw et al., 1992; Lepre et al., 2011) and could be as old as 1.7-1.4 Ma in

6

South Africa (Chazan et al., 2008; Gibbon et al., 2009; Kuman and Clarke, 2000). The first appearance of the Acheulean is roughly coeval with the first appearance of H. erectus (Lepre et al., 2011). A group of hominins sometimes designated as H. heidelbergensis, or H. rhodesiensis in Africa, are generally considered responsible for later Acheulean assemblages (e.g. Foley and Gamble, 2009).

The Middle Stone (MSA) is characterized by the absence of large bifaces, an emphasis on Levallois technology, and the presence of points (Goodwin and Van Riet Lowe, 1929). There is general consensus that the earliest MSA dates to ~300 ka (McBrearty and Brooks, 2000). The MSA is also associated with many developments characteristic of modern human behavior (Henshilwood and Marean, 2003; Marean and Assefa, 2005; McBrearty and Brooks, 2000), including the use of symbolic resources (e.g., d'Errico et al., 2001; Henshilwood et al., 2004; Henshilwood et al., 2009; Henshilwood et al., 2001) and new hunting and subsistence technologies (e.g., Faith, 2008; Marean et al., 2007; Villa et al., 2009b). The earlier MSA is generally attributed to a group of hominins that are variably described as late archaic Homo sapiens, or H. helmei (Foley and Lahr, 2003; Foley and Lahr, 1997; McBrearty and Brooks, 2000). By ~195-150 ka, anatomically modern human fossils are known from East Africa (Clark et al., 2003; McDougall et al., 2005; White et al., 2003) and modern H. sapiens are responsible for the later MSA.

The LSA is characterized by microlithic technologies, dates to ~40 ka, and represents modern hunter-gather populations (Deacon and Deacon, 1999; McBrearty and Brooks, 2000; Villa et al., 2012). Non-lithic artifacts and symbolic items such as polished bone tools, including , of shell and ostrich eggshell, and engraved bone and wood items are also abundant in LSA contexts, as well as items such as weights and grindstones.

Two regionally-limited industries, the and the Fauresmith, are associated with the ESA- MSA transition. Sangoan assemblages are generally characterized by the presence of ‘crude’ handaxes, core-, picks, and high frequencies of denticulates and notches (Kuman et al., 2005b; McBrearty, 1987; McBrearty, 1988; McBrearty, 1991, 1992). It has been suggested that Sangoan assemblages represent an adaptation to woodland environments based on the observation that the majority of Sangoan sites lie within the present high rainfall zone of tropical

7

Africa, and the high proportion of heavy-duty tools, denticulates and notches, which may have been employed in the manufacture of finished products in wood and bark (Clark, 1959b). However, there are Sangoan occurrences outside this zone (McBrearty, 1987, 1991; Sampson, 1974b; Van Peer et al., 2003) and Sangoan-like industries have also been described in South Africa (Kempson, 2007; Kuman et al., 2005a). The Fauresmith, known mainly from archaeological contexts south of the Limpopo River, is supposed to differ from the Sangoan in consisting of small, finely-made handaxes, with less emphasis on heavy-duty tools, denticulates, and notches, and reflect a local adaptation to drier environmental conditions (Clark, 1959b).

In addition to the characteristic types of large cutting tools (handaxes, cleavers, picks, core-axes) used to define Sangoan and Fauresmith assemblages, these types of assemblages also contain prepared core, or Levallois technology. Levallois technology involves the careful preparation of cores so that larger, thin flakes and blades of predetermined form can be extracted (Boëda, 1995; Chazan, 1997). There are antecedents of Levallois technology in the ESA. For example, the Industry in involves the shaping and preparation of very large cores so that large flakes could be removed, which were then shaped into handaxes or cleavers (McNabb, 2001). Other strategies of using large prepared cores to produce blanks for handaxe and manufacture are known in many parts of the Old World (Sharon, 2009).

2.1.2 The African Stone Age in Old World context

The tripartite scheme for the African Stone Age was modeled after the European Paleolithic sequence - the , the , and the - which has a much longer of research (McBrearty and Brooks, 2000). The distinct nomenclature for the African Stone Age was established to emphasize the reality that there are fundamental and important differences between the African and European sequences, but there is value in merging the two sequences to address some questions about hominin behavioral evolution.

Clark’s (1969) five technological Modes (1-5) have been used for nearly four decades as a simplified system for describing major trends in flaked stone technology at a global scale, and have become widely used in paleoanthropology. The sequence was based largely on the archaeological record of Western Europe, though Clark recognized that there were regions where the same cultural succession did not apply. Many paleoanthropologists have used or at least considered Clark’s Mode system when describing archaeological assemblages, situating

8 archaeological assemblages within the large-scale patterns of hominin morphological evolution, and considering early hominin dispersal across and beyond the Old World. The Modes are defined as follows:

 Mode 1 consists of pebble cores and flaked tools and includes the Oldowan industry.

 Mode 2 assemblages contain large bifacial cutting tools such as handaxes, cleavers, and picks and includes the Acheulean industry.

 Mode 3 assemblages are characterized by flake tools struck from prepared cores and includes the MSA and the Middle Paleolithic.

 Mode 4 assemblages are defined by presence of punch-struck blades from prismatic cores. The Upper Paleolithic is a typical Mode 4 industry.

 Mode 5 assemblages are distinguished by retouched and complex compound tools and includes Later Stone Age, late Upper Paleolithic and Epipaleolithic assemblages.

Clark viewed these modes as a cumulative sequence, reflecting the acquisition of more and more strategies over time, and assemblages were classified within the system based on their most “derived” component. There are problems with simplifying lithic variability to this extent, but in some cases Clark’s modes do serve as convenient framework for establishing large-scale temporal and spatial patterns in lithic technology and correlating these patterns to the fossil record. Some of these correlations are considered further in Section 2.3, after first discussing relevant changes in hominin morphology as evidence by the Middle Pleistocene fossil record.

2.2 Morphological Evolution in the Middle Pleistocene

2.2.1 Homo heidelbergensis sensu lato

The fossil record for the early Middle Pleistocene in Africa is extremely scant, but many of the specimens that have been recovered show a mixture of primitive features shared with the earlier (1.8-1.0 Ma) and derived features shared with modern Homo sapiens. There is debate on how to classify these specimens, with some researchers using the label “early archaic H. sapiens” and others Homo heidelbergensis. In Africa, this group of hominins is represented at

9 at least 5 localities, and perhaps as many as 10, depending on various definitions and designations. The earliest specimens date to at least 600 ka and the youngest to less than ~400 ka.

Taxonomy for early Middle Pleistocene hominins is contentious and the discussion about how best to classify these contemporaneous hominins in Europe, Africa, and Asia is ongoing. The biggest divide between researchers is whether the term H. heidelbergensis should refer to only European specimens or to African specimens as well (Mounier et al., 2009; Stringer, 2012). There is some evidence that supports the idea that H. heidelbergensis in Europe is a distinct group ancestral only to Neanderthals and should not be lumped with Middle Pleistocene hominins from Africa. In this view, Middle Pleistocene African specimens are more appropriately designated H. rhodesiensis, the species name given to the Kabwe (previously called Broken Hill 1) cranium discovered in Zambia. H. rhodesiensis is sometimes considered ancestral to modern humans, but not Neanderthals, and H. heidelbergensis as ancestral to Neanderthals, but not modern humans. Support for this view comes from derived features claimed for the European H. heidelbergensis fossils, especially those from Sima de los Huesos, Spain (Rosas, 2001).

Alternatively, there is the possibility that African Middle Pleistocene specimens share enough traits with the European specimens to be grouped with them and the H. heidelbergensis designation is appropriately attributed to specimens in Africa as well as Europe (Rightmire, 2001, 2004; Stringer, 2012). Proponents of this view generally regard H. heidelbergensis as the last common ancestor of both Neanderthals and modern humans. Stringer (2012) argues that the Sima de los Huesos specimens should actually be grouped with the H. neanderthalensis clade, an argument that is supported by a recent dental analysis (Martinón-Torres et al., 2012). Doing so would clear up much of the dispute about what features should be used to identify H. heidelbergensis.

The morphological specifics of the debate are beyond the scope of this research and are mainly an issue for the European record. For our purposes, it does not really matter whether the African Middle Pleistocene specimens are called H. heidelbergensis or H. rhodesiensis or whether they represent a species distinct from the European hominins at that time. I will use the term H. heidelbergensis sensu lato(s.l.) to describe certain Middle Pleistocene African fossils (Table 1),

10 and I assume these specimens represent or at least closely approximate a species or clade that is ancestral to H. sapiens. Potentially, this species or clade may also be ancestral to H. neanderthalensis.

The fossil record in the early Middle Pleistocene documents what has been characterized as an “encephalization event” (Gamble, 2010:23). et al. (1997) show that for the period between ~1.8 Ma and 600 ka there is no marked increase in the brain size relative to body size, i.e., encephalization, and these hominins are about one third as encephalized as modern humans. In contrast, during the next ~500 ka, encephalization comes within 10% of modern human values. Likewise, Rightmire (2001, 2004, 2008, 2012) argues H. heidelbergensis s.l. fossils differ morphologically from H. erectus in having significantly larger , and represent an early Middle Pleistocene speciation event.

As a group, Middle Pleistocene hominins have a mean cranial capacity of 1206 cm3, significantly greater than the 952 cm3 average for H. erectus (Rightmire, 2012). Even accounting for the fact that H. heidelbergensis s.l. had greater body size, H. heidelbergensis s.l. as a group exhibits a greater relative brain volume and enchephalization quotient1 (Rightmire, 2004). There is a strong pattern for an increase in absolute brain size within H. erectus over time, but H. heidelbergensis s.l. specimens have absolute brain sizes above that predicted by the Homo erectus trend, and there is no apparent chronological trend in brain size for H. heidelbergensis s.l. (Rightmire, 2004). Furthermore, there are morphological changes in the H. heidelbergensis s.l. cranium (longer parietals, rounded occipital, and differences in the basicranium) that are independent of brain size (Rightmire, 2012). It must be recognized, however , that Rightmire (2004, 2012) includes the Sima de los Huesos specimens and Omo II specimens (usually attributed to late archaic H. sapiens or H. helmei) in his analyses and it is reasonable to question whether the same patterns would hold true if they were excluded. Furthermore, the African specimens in his analysis – Bodo and Broken Hill – actually have relative brain size and enchephalization quotients within the range of H. erectus (see also Conroy et al., 2000).

1 The encephalization quotient is a relative measure of brain size with respect to body mass. Rightmire (2004) uses orbital height to estimate body mass.

11

Chronological data for African H. heidelbergensis specimens are rare, but suggest that the species first appears in Africa before 600 ka. One of the most complete cranial specimens is Bodo, recovered from in the , . A tuff in a nearby geological unit was dated to 640 ± 30 ka using the Ar/Ar method (Clark et al., 1994). Faunal and archaeological evidence support a correlation between the hominin-bearing sediments and the dated geological unit and this date is generally accepted as reliable. The chronological situation for Broken Hill (Kabwe) is less secure, however. The contextual details of its 1921 discovery are unknown. Associated faunal remains at a nearby locality show some similarities with Beds III-IV, suggesting an age range of approximately 700-400 ka. However, more recent paleomagnetic analyses indicate that the top of Bed IV could be at least 780 ka (McBrearty and Brooks, 2000, and references therein). If the correlations between the cranium, the fauna, and Olduvai Gorge Beds III-IV are accepted, than the Broken Hill cranium could be older than 780 ka. However, there are good reasons to question the connections. The same problem plagues the Elandsfontien specimen, which has also been linked to a faunal assemblage with affinities to Olduvai Beds III and IV (Klein, 1988). Chronological assessment of the Ndutu cranium gives a very broad time range. The specimen could be ~600-500 based on correlation with dated tuff (amino acid racemization), but could be as young as 200 ka depending on which tuff is correlated, with paleomagnetic data indicating a range from 990-370 ka (McBrearty and Brooks 2000, and references therein). Other African specimens are consistent with a Middle Pleistocene chronology (Table 1). The Salé specimen gives us the most recent bounding maximum age at 455 ka. In sum, the strictest range for H. heidelbergensis in Africa seems to be ~600-400 ka, but the actual duration that this species can probably be extended on either side of this time range.

2.2.2 H. helmei

The Florisbad cranium (Dreyer, 1935), recovered from Florisbad spring in central South Africa, is often considered the type specimen of H. helmei and is dated by ESR to 259 ± 35 ka (Grün et al., 1996). Like H. heidelbergensis, H. helmei is a problematic taxon. It has not been formally defined as a species. Some researchers prefer to lump the Florisbad cranium with Homo sapiens, while others emphasize differences in brow ridge development and angulation of the occipital bone compared to ‘anatomically modern humans’ and recognize it as a separate taxon. Other researchers lump H. helmei with H. rhodesiensis (Wood and Leakey, 2011). In a gradational scheme for the genus Homo, McBrearty and Brooks (2000) assign the Florisbad cranium to

12

Table 1 African H. heidelbergensis fossils, chronologies, and archaeological associations Location Preservation Cranial characteristics Cranial Age Archaeology Other notes References capacity (cm3) Broken Hill Zambia, Adult cranium archaic traits: projecting face, 1,280 thought to be 700-400 ka handaxes, points, haplotype of H. (Clark, 1959a, (Kabwe) East Africa and cranial, heavy brows, cranial base is based on associated fauna (cf. prepared cores, rhodesiensis and references dental, and less flexed, derived traits: Olduvai Beds III-IV), but recent Sangoan or MSA therein; postcranial vertical nose analyses correlate the top of designation?, McBrearty and remains of 3 or Bed IV with the Matuyama unknown which Brooks, 2000; more individuals Bruhnes boundary (780-1.33 archaeological Rightmire, Ma) or base of Masek Beds stratum is associated 2009), with Jaramillo subchron (1.07- with fossils 1.33)

Bodo Middle Adult cranium, archaic traits: large face, 1,250 ~600 ka (640 ± 30 ka) based on assemblages (Clark et al., Awash, parietal heavy brows, post-orbital Ar/Ar of tuff in correlated unit designated as 1994) Ethiopia, fragment, and constriction and low forehead, (correlation based on faunal Oldowan and East Africa distal humerus , derived traits: and archaeological evidence) Acheulean, simple vertical nose, anatomy of flakes and cores, palate, other characteristics of some locations have cranial bones handaxes

Elandsfontien South Africa Adult calvaria, archaic traits: large brow n/a potentially associated fauna at nearby Cutting 10 (Herries, 2011; (Saldanha, mandible ridges, sloped forehead, comparable to Olduvai Gorge site, Acheulean Klein, 1978; Hopefield) fragment derived traits: parietal bossing Beds III and IV handaxes ("finely- Klein, 1988; and vertical occipital made") McBrearty and (consistent with increased Brooks, 2000; brain size) Rightmire, 2009; Singer and Wymer, 1968) Ndutu , Adult cranium derived traits: convex side 1,100 could be ~600-500 based on non-diagnostic probably female (Clarke, 1990; East Africa walls, characteristics of correlation with dated tuff artifacts, similar based on more McBrearty and temporomandibular joint (AAR), could be as young as sediments nearby gracile browridge Brooks, 2000; 200 ka depending on which yield Acheulean and other features Mturi, 1976; tuff is correlated, assemblages ('later' Rightmire, 1983) paleomagnetic data indicates Acheulean?) range from 370-990 ka Kapthurin , East Two adult n/a n/a 510–512 ka Ar/Ar handaxes, points, fossils show affinities (Deino and Africa mandibles, and blades, Levallois (K3 with H. erectus, or McBrearty, postcranials Member) Homo ind. sp. 2002; Johnson and McBrearty, 2010; McBrearty and Brooks, 2000)

13

?Thomas , Subadult partial n/a n/a Middle Pleistocene based on none (McBrearty and Quarry mandible, associated fauna Brooks, 2000and cranial, and references maxillary therein ) fragments ?Salé Morrocco, two adult archaic traits: thick walls, n/a 389-455 ka, ESR on associated none pathology may (Hublin, 1991as North Africa mandibles, and sagittal keel, derived traits: bovid teeth explain small brain cited in postcranials characteristics of squamous size and some McBrearty and temporal and parietal vault derived Brooks 2000) characteristics, Rightmire 2004 attributes to H. erectus

?Tighenif Oran, three mandibles, n/a n/a Middle Pleistocene based on Acheulean (Geraads et al., (Ternifine) Algeria, parietal fragment associated fauna, 1986) North Africa paleomagnetism suggests age of more than 600 ka

?Berg Aukas Namibia, Femoral n/a n/a undated none (Grine et al., Southern fragment 1995) Africa

14

“Group 2” with specimens like Ngaloba, Guomde, Omo II, and Porc Epic. These ‘Group 2” specimens are morphologically intermediate between ‘Group 1’, which includes H. rhodesiensis and/or H. heidelbergensis, and ‘Group 3’ (anatomically modern Homo sapiens). Brauer (1997) uses the term “late archaic Homo sapiens” to describe these intermediate specimens and distinguish them from H. heidelbergensis s.l. which is described as “early archaic Homo sapiens”. Foley and Lahr (2003; 1997) also extend the taxon H. helmei to some Eurasian specimens based solely on contemporaneity with the African specimens.

Besides the Florisbad cranium, there are a few other H. helmei specimens that have been dated to the middle Middle Pleistocene. The Guomde remains (including KNM-ER 3884) from , Kenya, are dated to between 270 and 300 ka based on U-series analyses (Brauer et al., 1997). The Ngaloba cranium from Tanzania gave an ages estimate of ~200 ka based on amino acid racemization of ostrich eggshell, and correlation with units dated by 40Ar/39Ar, but other interpretations of the stratigraphy and correlations are possible, and the specimen could date to >200 ka (Table 1 in McBrearty and Brooks, 2000). Other H. helmei specimens give either minimum ages estimates of ~130 ka or less, or are dated to the Late Pleistocene (McBrearty and Brooks, 2000). When there are associated archaeological assemblages with H. helmei specimens, they are MSA assemblages (McBrearty and Brooks, 2000).

2.3 The Significance of Technological and Morphological Change in the Middle Pleistocene

The origins of Mode 3 technology, represented by the Middle Paleolithic in Europe and the MSA in Africa, reflects a shift away from the manufacture of large core tools like handaxes, cleavers, and picks, toward an emphasis on smaller flake tools. Flake tools were also manufactured in the Acheulean, and large core tools are still used in the MSA and Middle Paleolithic. However, core reduction strategies in Mode 3 assemblages involve more preparation and the resulting flakes are often larger, with a more regular form and thinner profile. Some of these flakes were hafted onto handles, forming parts of composite tools.

Multiple interpretations have been put forth for the how, what, and why of technological change in the Middle Pleistocene, and I present these multiple interpretations below. They represent diverse ideas about why technological change occurred, what the implications were for how we

15 interpret hominin capabilities and behavior, and how the evolution of Homo was affected. Some of these ideas are largely informed by the archaeological record, some more so by the morphological record, and others by psychology, sociobiology, and modern hunter-gatherer . They are not conflicting models and to label them as such would be disingenuous. At this point, the archaeological record for the Middle Pleistocene is not robust enough to test conflicting models, even if they did exist. We are still in the model building phase for the origins of Mode 3 technology and hominin evolution in the early Middle Pleistocene.

2.3.1 Hominin speciation and dispersal

The appearance and distribution of Mode 3 has been connected to hominin speciation events. Foley and Lahr (1997) transform Clark's modes into a phylogenetic framework, linking the major archaeological changes represented by each Mode to biological species and their local distributions. They argue that the archaeological evidence can be interpreted in a phylogenetic sense and used to track hominin speciation and dispersals, as well as population-specific behavioral traits. They suggest a “Mode 3 Hypothesis”, which correlates the origins of Mode 3 to the middle Middle Pleistocene ~300 ka. The hominin associated with the development of Mode 3 is probably H. helmei – a taxon represented in Africa by specimens such as Florisbad. Shortly after developing in Africa, Mode 3 technologies spread to Europe and the hominin taxon associated with Mode 3 diverges into two biological lineages - Neanderthals in Europe and modern H. sapiens in Africa. These two species used similar technologies. Shared technological capacities between Neanderthals and modern humans were inherited, with each species following their own evolutionary trajectory, and with different evolutionary trajectories in different regions of the world. The implication of this model could be that the origins of Mode 4 and Mode 5 technologies could reflect regional stylization of stone technologies more so than they do cognitive advances. Foley and Lahr (1997) mention that there could be a link between the appearance of Mode 3, an increase in brain size, and an increase in social group size (Aiello and Dunbar, 1993). Flexibility offered by Mode 3 technology may have been a response to encephalization and increased sociality. In this model, H. heidelbergensis in Africa and Europe is associated with Mode 2 technology. According to Foley and Lahr (1997), the abandonment of Mode 2 once Mode 3 technology was established was fairly rapid in Africa, and soon after its initial appearance, Mode 3 technology dispersed into .

16

In their 2003 article, Foley and Lahr again suggest that Mode 3 technologies, originating in Africa ~300 ka may have spread to Europe with H. helmei, perhaps the most recent ancestor of Neanderthals and modern Homo sapiens. They point out that, while fossil evidence indicates an early divergence of the Neanderthal and Homo sapiens lines in the early or middle Middle Pleistocene, the archaeological evidence suggests a more recent shared ancestor. Based on timing, they argue that H. heidelbergensis is not correlated with any major technological change. Mode 3 is important because it marks the beginning of human levels of variation; there are a diversity of flakes that can be shaped into a diversity of tool forms. In earlier modes we see faithful imitation (i.e. handaxes), suggesting a limited ability to innovate and make modifications. They argue that there is a strong contrast in behavioral flexibility between hominins before and including H. heidelbergensis and those after.

2.3.2 Modern human behavior

One of the hottest topics in the past couple of decades has been the origins and evolution of modern human behavior (for reviews see Henshilwood and Marean, 2003; McBrearty and Brooks, 2000; Shea, 2011b; Wadley, 2003; Willoughby, 2007). McBrearty and Brooks (2000) highlight the substantial archaeological evidence for modern human behavior that appears alongside Mode 3 technology, long before its first appearance in Europe, where it is associated with Mode 4/5 technologies. The behaviors associated with modern human behavior that are represented in the archaeological record of the MSA include the use of symbolic resources such as , beads, and incised objects, regional diversification of artifact , and long- distance exchange of lithic raw material.

The view of McBrearty and Brooks (2000) on the potential origin and evolution of is generally characterized as the ‘gradualist’ model (Henshilwood and Marean, 2003) and contrasts the Late Pleistocene ‘revolutionary’ model (e.g. Binford, 1985; Klein, 2001) that holds that the shift to modernity was a rapid event dating to ~50 ka. McBrearty and Brooks (2000) suggest that the origins of Mode 3 about 300 ka reflects a significant behavioral shift leading to modernity, because Mode 3 technology is associated with both H. helmei (or ‘Group 2’ hominins in a gradational scheme) and H. sapiens. This implies increased cognitive abilities with the appearance of H. helmei ~300 ka, and behavioral similarities and a close phylogenetic relationship between H. helmei and H. sapiens. During the course of the MSA, the behaviors we

17 associate with behavioral modernity accumulate, representing the expansion of a shared body of knowledge, and the application of novel solutions, rather than biological or cultural ‘revolutions’.

McBrearty and Brooks (2000) point out that the temporal and morphological details of the Middle Pleistocene hominins remain to be worked out, but there may have been multiple species present in Africa during this period, which would help explain the diversity of assemblage types. They do not provide a mechanism for explaining technological similarities between Neanderthals and early MSA hominins.

2.3.3 Behavioral variability and adaptation

Middle Pleistocene lithic technology has also been associated with ‘human-like’ capacities for behavioral adaptation to varied environments. As originally conceived, Mode 3 was situated within a temporal trajectory between Mode 2 and Mode 4 based on the Western European archaeological record, but there are problems with conceiving of lithic technological variability solely as lineal progressive change (Shea, 2011b). Lithic artifacts are the interface between hominins and their physical environment and are part of the behavioral adaptation they employ to extract resources, thus we can expect a large part of lithic artifact variation to result from variability in the local environment. Shea (2011b) argues that by 230 ka in East Africa, elements from Modes 1-4 are all present. Based on this observation, Middle Pleistocene hominins ~200 ka exhibit as much behavioral variability as Late Pleistocene hominins. The focus on behavioral variability is an explicit shift away from the arguably flawed concept of modern human behavior. From this perspective, the significance of Mode 3 is that it is part of a flexible and diverse technological repertoire employed by humans. With respect to lithic technology, there may be no diachronic pattern of change from 230 ka onward. Instead, characteristics seem to come and go, suggesting that variability in Middle and Late Pleistocene lithic technology reflects behavioral adaptations to local resource availabilities rather than cognitive advances. Shea (2011b) also points out that many aspects of the Neanderthal archaeological record, also characterized by Modes 1-4, are consistent with a shared common ancestor with a flexible and diverse toolkit that responds to local resource availabilities in a similar fashion. Therefore, one would expect behavioral convergences between Neanderthals and modern humans reflecting adaptation.

18

In a more recent article, Shea (in press) emphasizes the limitations of the Mode concept. Clark's Modes stand in the way of more nuanced descriptions of prehistoric lithic variability, reflecting a progressive and cumulative model of technological change, and ignore potentially significant variation in the retention and loss of certain characteristics because each mode is defined solely by its most 'derived' characteristic. It matters whether certain strategies are maintained or discarded in light of new technologies. For example, we would expect generalist strategies to continue using multiple types of reduction and tools, whereas specialist strategies may focus on one manner of reduction and abandon older strategies.

Shea (in press) introduces a new framework, Modes A-I (Table 2) and applies this framework to the Levantine paleolithic to demonstrate how it picks up new patterns and results in new questions. For instance, bifacial hierarchical cores with a preferential removal (preferential Levallois, Boeda 1995) designated by Shea as Mode F1 are present in the archaeological record by 800 ka. Recurrent bifacial hierarchical cores with preferential removals (F2) only appear after ~250 ka. Also about this time, unifacial hierarchical cores (G1) become commonplace. Large Cutting Tools (LCTs, Mode E1) are fairly consistently present right up 65 ka. Many of these assemblages are typically described as and designated to Clark's Mode 3, despite the presence of LCTs, which in the absence of bifacial hierarchical cores would lead to a Clark's Mode 2 designation. Shea (in press) emphasizes that this period ~250 ka represents a shift to smaller flake tools over the use of heavy core tools, and this may be related to increasing the cutting edge to volume in the context of more technologically-mediated adaptive strategies. Technological behaviors ~250 ka reflect greater time and investment represented by and core preparation and more efficient use of stone resources. For Shea, the addition of Modes F2 and G1 to the toolkits of previous , which some could argue would define the origins of Mode 3, probably reflects adaptation to ecological conditions.

Shea’s (in press; 2011b) emphasis on behavioral variability and adaptation is in line with Potts’ (1998) argument for variability selection through the Plio-Pleistocene. Variability selection is a concept that links adaptive change to environmental variability. It refers to the process by which traits that increase adaptability to diverse environments and conditions out compete traits that confer an advantage in one specific type of environment under specific conditions. Variability selection results in complex structures or behaviors that are designed to respond to novel and unpredictable adaptive settings. Environmental variability has increased over the last 6 Ma. Potts

19

Table 2 Shea’s (in press) new paleolithic framework with Modes A-I. A Stone percussors B Bipolar cores C Pebble cores/non-hierarchical cores D Retouched flakes D1. Retouched flake-tools D2. Backed/truncated flakes D3. Burins D4.Retouched microliths E Elongated core tools E1. Large cutting tools E2. Thinned bifaces E3. Bifacial core tools with retouched proximal concavities (BRPC) E4. Celts F Bifacial hierarchical cores (BHC) F1. BHC -Preferential F2. BHC -Recurrent G Unifacial hierarchical cores G1. Platform cores G2. Blade cores H Edge-ground tools I Groundstone tools

(1998) argues that, based on istopes in deep-sea cores, environmental variability has been greatest in the last 6 Ma among the last 27 Ma, and in the last 2 Ma among the last 6 Ma. Based on pollen from the Tenaghi Philippon peat bog in , variability has been even greater in the last ~700 ka among the last 2 Ma. Furthermore, there has been a gradual increase in global climatic variability over time. The range of delta 18O values in the last ~700 ka spans the entire range known for the last 6 Ma. Paleoontological evidence also demonstrates that variability selection affected many faunal species in East Africa between 900-600 ka. Species that persisted past this period had wider dietary niches and smaller body sizes than species that went extinct. Potts (1998) argues that variability selection in hominins helps explain the mosaic of bipedal and arboreal adaptations in early hominins, who were adapted to both closed and wooded environments, and more importantly for the discussion here, variability selection helps explain encephalization, which permits flexible behavioral responses to new environments. Roughly coeval with the greatest increases in environmental variability, the largest changes in hominin encephalization appear to have occurred ~600 ka, as discussed above.

20

2.3.4 Cognition

Large brains are a part of what make humans different from extant . Large brains serve multiple purposes and many explanations for the evolutionary development of large brains have been put forth, many of which are related to technological changes observed in the Middle Pleistocene. Problem solving and long term planning are important components of the human . 'Working memory' refers to a set of abilities that permits the mind to hold in attention and process task-relevant information in the face of interference. It is primarily a network within the prefrontal cortex, but is also interlinked to much of the neocortex. Individuals with damage to the frontal lobes are unable to carry out complex, purposeful, and goal-directed actions. A significant aspect of working memory is response inhibition, i.e., delayed gratification. Low-level working memory is a trait shared with nonhuman primates, so Wynn and Coolidge (2011) used the phrase enhanced working memory to describe modern human capacities.

Traces in the archaeological record for working memory are difficult to detect, largely because technology relies on procedural cognition and long-term motor memory rather than working memory. Wynn and Coolidge suggest that snares and traps for animal capture, which present action towards a future goal, are good indicators of enhanced working memory. Indirect evidence for using traps by 70 ka in the MSA may be provided by the extensive duiker remains at , South Africa (Wadley, 2010). Multi-day hafting procedures also indicate enhanced working memory. Wynn and Coolidge (2011) argue that the cognitive implications of simple spears involving a shaft, point, and a single binding material are different than hafted tools that would have taken multiple days to manufacture. The latter are closer to the 'reliable' side of the maintainable-reliable continuum of tool kits (Bleed, 1986). Reliable tools are overdesigned to ensure function and reduce failure. Remnants of this kind of behavior indicates response inhibition and contingency planning. Residues used at Sibudu Cave >70 ka contained multiple ingredients, i.e., beeswax, , and resin, and the making of compound adhesives such as this could reflect enhanced working memory (Wadley et al., 2009). Other archaeological evidence that indicates enhanced working memory is the externalization of information, represented by incised stone and bone, and personal ornaments, with evidence in South Africa ~100 ka (e.g. d'Errico et al., 2001; Henshilwood et al., 2004; Henshilwood et al., 2009; Henshilwood et al., 2001). In this view, there is no evidence for enhanced working memory associated with the origins of Mode 3 technology, but developments do appear over the course of

21 the MSA. Hafting may be associated with a development in working memory, but the simpler strategies ~300 ka and associated with Neanderthals, do not in themselves indicate modern capacities according to Wynn and Coolidge (2011).

Ambrose (2010) situates the origins of Mode 3 and composite tool technology within the framework of constructive memory – the uniquely human capacity for imagining future scenarios and planning for them. This type of memory is linked to an area of the anterior/medial area of the brain called the frontopolar prefrontal cortex, which is larger in humans than in . This part of the brain is only activated when subjects have to keep in mind a main goal, while performing concurrent subgoals (Koechlin et al., 1999). Ambrose identifies the origins of composite tools ~300 ka as one key event in the evolution of constructive memory in humans. Composite tool manufacture requires constructive memory. It involves the collection and preparation of several kinds of components composed of different raw materials obtained at different times and places. Final assembly occurs later. Some materials may be kept for maintenance and repair. There are homologous neurological pathways for speech and manual artifact assembly, implying that composite artifact manufacture and grammatical language coevolved (Greenfield, 1991). Both speech and composite tool manufacture involve hierarchical sequencing of different components and fine-motor control (Ambrose, 2001).

2.3.5 Social networks and population expansion

Mode 3 technologies have also been interpreted in light of the expansion of human social networks and increases in population. Foley and Gamble (2009) identify five major transitions, with transitions 3 and 4 occurring in the Middle Pleistocene, and of interest here. Transition 3 (800-700 ka) is associated with what they call Mode 2B, characterized by finely-made symmetrical handaxes, and based on evidence from Gesher Benot Ya’aqov, the origins of controlled-fire use and . They suggest that the social implications of fire and cooking may have been a greater emphasis on bonding between individual males and females, strengthening smaller family units within the larger community structure. Transition 4 (400-300 ka) is associated with Mode 3 technologies, ‘projectiles’ (presumably referring to hafted spears, not darts or bows and ), and the abandonment of the handaxe. This period is correlated with H. helmei, and the divergence of the Neanderthal and human lineages, as well as encephalization and the regionalization of lithic production styles. Foley and Gamble (2009)

22 suggest that new technologies resulted in more reliable resources in a greater range of environments and consequently, population increase and demographic packing. Demographic packing may explain the increase in brain size (the social brain hypothesis posits that larger brains function to cope with larger group sizes, Dunbar, 1998; Dunbar, 2003, and see below) and may also explain diversification at the regional level, because material culture began to communicate social identity. The social result of these changes was a supracommunity structure with ‘exploded fission-fussion’ dynamics, within which large social networks could be maintained even when small communities were separated from each other for long periods of time.

2.3.6 The social brain hypothesis

The social brain hypothesis posits the large size of the human brain serves to solve social problems rather than ecological problems (Aiello and Dunbar, 1993; Dunbar, 1998; Dunbar, 2003). This idea is supported by a correlation between neocortex size and group size in primates and other . In modern humans, group size is ~150 on average and is considered the number of individuals with which one has a personal relationship, or the number of people of whom one feels one can ask a favor (Dunbar, 2003).

Gamble et al. (2011) integrate the social brain hypothesis with the archaeological record of . The terms ‘materials’ and ‘emotions’ are used to describe the resources used by hominins to solve social problems and reinforce social bonds. ‘Materials’ describe physical resources – the use of which has better archaeological visibility, while ‘emotions’ seem to reflect the less tangible resources (ie. theory of mind, language). They identify three major ‘movements’ in the Pleistocene, 2.6-1.6 Ma, 1.5-0.4 Ma, 300-25 ka. The last two movements are of interest here.

1.5 Ma-400 ka: This movement is described has having two subparts, with the second subpart (600-400 ka) documenting most of the change. The first part spans the time period associated with Acheulean assemblages in Africa. During this period brain size increases and group sizes expand from about 50 to about 100. Materially, hominins during this period were armed with handaxes and other stone tools manufactured from local raw materials. These tools show more symmetry than earlier tools with important cognitive implications (Hodgson, 2009), but there is little change in how materials are used over time. After 600 ka, the material resources

23 used by hominins to mediate social relations included bone and wood tools, control of fire and , and dietary changes associated with hunting large game. ‘Emotional’ resources include language, , and third-party punishment. Gamble et al. (2011) suggest that the absence of material indicators of change in the earlier part of this period (before 600 ka) despite the increase in brain size implies that immaterial ‘emotional’ resources may have played a more important role than material resources.

300-25 ka: This movement is associated with the many developments in the MSA that indicate extended social relations. Materials that gain importance in this period include shellfish, beads, ochre, and composite artifacts and hunting. Some materials are exchanged beyond 100 km, suggesting that and exchange networks existed across the landscape. Composite artifacts require the same kind of recursive analogical reasoning that underpins complex language. The types of ‘emotional’ resources employed by humans during this movement also proliferate. Brain size at this time is similar to contemporary humans, indicating that social groups were large, language and other symbolic systems of communication became full-blown, and humans maintained large social networks beyond their immediate communities. During this movement, both material and ‘emotion’ is important for reinforcing social bonds and the pace of change quickens.

In this model put forth by Gamble et al. (2011), Mode 3 technologies and composite artifacts explicitly serve as a marker for the cognitive and linguistic capacities of early humans. Also, stone tool technology, in general, was a resource for creating, maintaining, and manipulating social relations, and solving the problems associated with living in large groups, and this could be part of the significance of Mode 3 technology (Gamble, 1999). Gamble et al. (2011) do not extrapolate on this potential role for Mode 3 and composite technology, but if analogies are drawn from their treatment of -eating ~2.5 Ma, one could argue that group dynamics linked to cooperation, sharing, and reciprocity would have been impacted by a more reliable access to large game.

Ambrose (2010) also uses the social brain hypothesis to understand technological change in the later MSA ~70 ka. He argues that evidence for large-scale social networks in the later MSA takes the form of long-distance exchange of lithic raw material. The maintenance of long-distance social relationships may have required innovations in speech that engender trust and reciprocity

24 and enhancements to constructive memory for engaging in future-tense speech. In other words, changes to the structure of the brain during this transition 70 ka may have resulted from the need to live in large and complex social groups.

2.3.7 Cooperation and ultrasociality

Hill et al. (2009) argue that two things account for human uniqueness, reliance on social learning and cooperation between nonkin, and that these two traits evolved among later Homo. have socially-learned traits, but they have different mechanisms of social learning. There is no accumulation of learned traits, and the content of what is transmitted is different. Also, animal culture does not exhibit socially-learned regulations of behavior with third-party punishment, i.e., what we call norms, nor do they use symbolic reinforcement of a specific rule system, i.e., what we call ethnicity. In contrast, humans have norms, ethnicity, and are ultrasocial. Humans cooperate in large social groups composed partially or completely of nonkin (Hill et al., 2011).

Human cooperation is intimately tied to our hunter-gatherer ancestry. Early humans at some point developed a dependence on high-quality resources that included meat from large game. Chimpanzees also eat meat, but that meat comes in small packages and is not shared the same way that humans share meat. Hunting large game changes the reliability of food resources and its potential for distribution, with important life-history implications for humans, including the need to provision children, longer childhood, lower adult mortality, and later senescence (Kaplan et al., 2000). These life-history characteristics co-evolved with another characteristic of humans – cooperative breeding.

Cooperative breeding is a pattern in which individuals other than genetic parents engage in behaviors that increase juvenile survival and/or the survival and fertility of reproductive adults. Food provisioning is one way that humans participate in cooperative breeding. Among the Ache and Hiwi hunter-gathers of , for example, about 30-45% of adults have no dependent offspring, yet they intentionally acquire more food than they consume (Hill and Hurtado, 2009).

Hill et al. (2009) suggest the following temporal sequence for the evolution of human social learning and cooperation:

25

1. During the , hominins became bipedal, developed more dexterous , and amplified home ranges across various environments 2. In the Middle Pleistocene hunting provided a reliable source of high-quality food, and a slower life history developed - juveniles were more dependent, adult mortality was reduced - and cooperative breeding helped to buffer the costs of slower biological development .With cooperative breeding, prosocial emotions, enhanced theory of mind, and shared intentionality also evolved. 3. During the Late Pleistocene, humans developed a capacity for cumulative cultural change, resulting from the interaction of imitation and shared intentionality. This required large populations, because social networks allow for increased observation of rare innovations that are unlikely to be discovered by independent learning. When people reside together, they have frequent opportunities to observe innovations, evaluate their success, and imitate traits judged most successful or most common in large populations. 4. By the end of the Late Pleistocene, cumulate culture change and prosocial emotions led to language, norms, ethnicity, and extensive nonkin cooperation.

The second and third parts of the sequence are most relevant to the discussion here. According to the model put forth by (Hill et al., 2009), at some point in the archaeological record of the Middle Pleistocene we would expect archaeological evidence for an increased reliance on hunting and high quality food resources such as large game, as well as evidence for shared intentionality in the way that stone tools are manufactured. By at least the Late Pleistocene, there should be evidence for cumulative cultural capacities.

2.3.8 Shared intentionality

Tomasello (1999; Tomasello et al., 2005) argues that human cognition is unique among humans because of shared intentionality - the ability to participate with others in collaborative activities with shared goals and coordinated action roles for pursuing those goals. Shared intentionality depends on theory of mind, which is the ability to attribute mental states to oneself and others, but theory of mind alone cannot account for human cognitive uniqueness, because other primates exhibit theory of mind. When sharing intentionality, interactants are engaged with one another in a particular way and the goals of each interactant includes the goals of the other. Shared intentionality requires cultural learning and the ability to read the intentions of others. When

26 individuals in complex social groups share intention with one another repeatedly in particular interactive contexts, the result is habitual social practices and beliefs. The way that humans understand the actions and perceptions of others creates species-unique forms of cultural learning, and humans are biologically adapted for participating in collaborative activities. These cognitive skills make possible the creation and use of artifacts and technologies that accumulate modifications over generations through cultural evolution. Shared intentionality and cooperation also provide the foundation of symbolism and language, which are inherently collaborative. Tomasello et al. (Tomasello et al., 2005) suggest that shared (or "we" intentionality) has antecedents in non-human primates; for example, apes understand that others have intentional actions, but the full expression of collaborative engagement and shared intentionality appeared at some point in the evolution of Homo. The repetitive form and reduction strategies represented by Acheulean assemblages by ~1.8 Ma suggests that both the goal and means of LCT production were transmitted between hominins, suggestive of shared intentions (Shipton, 2010). However, the nature and/or degree that instructive learning and collaboration played a role in LCT manufacture could be debated, and it remains possible that human-like capacities for shared intentionality that explains cumulative culture change is unique toHomo sapiens (Tomasello et al., 2005).

2.3.9 Level 4 intentionality

Dunbar (2003) uses a concept of enhanced intentionality, or an increased level of intentionality, to distinguish between the nonhuman and modern human theory of mind. He argues that because there is a lineal relationship between level of intentionality and frontal lobe volume in primates, we can use frontal lobe volume as a proxy for level of intentionality in extinct hominins. These are the following levels of intentionality he recognizes:  Level 1 – no theory of mind (the ability to attribute mental states to oneself and others), primates at this level of intentionality know what they know  Level 2 – minimal theory of mind, representing the upper limit for nonhuman primates, primates at this level know that others know something  Level 3 – presumably a level of intentionality intermediate between extant nonhuman primates and humans, hominins at this level would know that someone else knows that someone else knows something

27

 Level 4 – the level characteristic of normal human adults and the minimum level required for (e.g. I know that you know that there is a supernatural force that knows that you and I know)

Based on frontal lobe volume estimates, the Australopiths cluster around Level 2 intentionality, H. erectus around Level 3, and ‘archaic H. sapiens’ (presumably H. helmei and H. heidelbergensis s.l.) at 3.5. Anatomically modern humans are above Level 4 and Neanderthals straddle the Level 4 line. Dunbar (2003) suggests this indicates that religion was lacking in H. erectus and probably only came into being with archaic H. sapiens.

2.3.10 Gene-culture coevolution

The cultural evolutionary program of Boyd and Richerson (e.g. Boyd and Richerson, 1985; Boyd and Richerson, 1987; Boyd and Richerson, 2005; Richerson and Boyd, 2005) holds that a continuous range of variation exists in the cultural information that each individual harbors and that this variation is transmitted between individuals through observation and learning. In this view, culture is “any kind of mental state, conscious or not, that is acquired or modified by social learning and affects behavior” (Richerson and Boyd, 2005:5). Culture makes human evolution different from the evolution of other organisms, in part because social learning allows human populations to accumulate of adaptive behaviors and technology and this process is faster than genetic evolution. Because culture change is also faster than genetic evolution, it is adaptive to changing environments. In other words, social learning and culture permit adaptation to spatially and temporally heterogenous environments, such as those characterizing the last ~2 Ma of hominin evolution.

Boyd and Richerson (2005) argue that individual human intelligence is only a small part of being able to create complex adaptive behaviors. Complex cultural traditions are the product of a population of rather than individuals, and humans seem to depend on socially-learned strategies more so than their own cognitive prowess.

The capacity for this kind of complex adaptive behavior has a biological/genetic basis. Boyd and Richerson (2005:260-264) posit that the innate social instincts that characterize humans and permit cultural evolution first developed in the Pleistocene, a idea they describe as the ‘tribal instincts hypothesis’. They point out that genetic change probably occurred in response to

28 humans living in social groups and resulted in a culture-gene coevolutionary process that largely explains the large-scale cooperative institutions characteristic of humans today. For Boyd and Richerson, the concept of is an important premise of this argument; larger groups of cooperative nonkin outcompeted smaller groups with less cooperation. This had an impact on the genetic makeup of humans – genes favoring cooperative prosocial behavior would have been selected for, and also the culture of humans in the Pleistocene – cultural transmission biases would have favored the transmission of socially learned cooperative and prosocial behaviors. At some point in the Pleistocene, humans developed the psychological machinery required to participate in interfamilial cooperation and then the human capacity to live in larger and larger networks evolved through a gene-culture coevolutionary ratchet.

Boyd and Richerson (2005) do not explicitly pinpoint the timing for the appearance of these pro- social behaviors, though their discussion does hint at some important elements that may be observable. Because paleoanthropologists consider humans behaviorally modern by ~50 ka during the Upper Paleolithic, human pro-social behavior was certainly in place by then. Likely, the origin is much earlier than that based on the appearance of traits consistent with modern human behavior in the African MSA (McBrearty and Brooks, 2000). They suggest that Pleistocene hunter-gatherers probably lived in nonkin groups based on the role of big game hunting in Paleolithic society. Big game hunting generally requires cooperation on a larger scale than plant collecting or small game hunting (Steward, 1955). They also argue that the persistence of complex technologies requires large populations of people with high degrees of interaction in order to favor imitation of highly skilled performances and to compensate for loss of skills due to imperfect transmission (Henrich, 2004).

2.3.11 The ratchet effect and imitation

Some researches hold that human culture is unique because it accumulates modifications over time (the "ratchet effect", Tennie et al., 2009; Tomasello, 1999; Tomasello et al., 2005; Tomasello et al., 1993). One generation does things a certain way, and the next generation does it the same way, except they add modifications or improvements. The generation after that learns the modified or improved way. There is little loss or backward slippage and new strategies do not have to be relearned every generation. The process relies on both faithful transmission from one generation to the next and inventiveness to generate modifications and improvements.

29

Inventiveness, i.e., independent learning, could be a characteristic shared with primates, but only humans transmit cultural knowledge across generations (Tomasello et al., 1993).

Tennie et al. (2009) demonstrate that human social learning differs from learning in nonhuman primates. In humans, learning is more oriented toward process than product, with more emphasis on imitation than emulation. Here, imitation learning describes when subjects copy the strategies of their teacher in order to achieve their goal, while emulation learning describes when subjects learn about parts of their environment and use this to achieve their goal. The transmission of nut- cracking behaviors in chimpanzees might be an example of emulation. Juvenile chimpanzees learn how to crack nuts in the proximity of adults, facilitated by the co-presence of nuts, , and anvils. They try many methods involving these objects until they learn how to extract nuts themselves. In other words, chimpanzees re-construct the product rather than copy the process leading to it. In a sense, when a learns a new behavior they are actually re-inventing the strategy. They observe an effect and then use their own behavioral strategies to reproduce the effect.

Humans rely on emulation learning, but in many situations they also pay attention to the actual behavioral strategies of others, and copy the actions and techniques. Imitation seems to have been important for stone tool manufacture by the , as evidence by the repetitive form of LCTs and the shared reduction strategies for their manufacture (Shipton, 2010). Because humans imitate behaviors, they are able to adopt successful strategies and create complex technologies that they would have been unlikely to invent on their own in a single lifetime. Through imitation, human social learning also provides knowledge accumulated across many generations that an individual alone could never produce.

Tennie et al. (2009) use the phrase ‘zone of latent solutions’ (ZLS) to describe the kinds of behaviors that an individual could learn on their own under the appropriate ecological and social conditions. Nut-cracking with stone and/or wooden hammers and anvils among chimpanzees is within their ‘ZLS’. Without the kind of imitation learning that characterizes humans, chimpanzees are not able to learn behaviors that are outside their ZLS. Results of an experiment conducted by Tennie et al. (2009) illustrate this clearly. Apes and 4-year old human children were given the opportunity to gain a reward. This reward was on a platform behind a cage and could only be accessed by making a loop with a string, positioning that string over a handle on

30 the platform, and then using that loop to pull a reward-bearing platform toward them. First, the subjects were given the string, and no demonstration of what to do with it. No or child was able to gain the reward. Second, the subjects were given the string and the platform was pushed closer to the cage to demonstrate that the platform could be moved. No apes were able to gain the reward, but one child out of 12 made a loop with the string and pulled the platform towards them to gain the reward. Third, the subjects were given a string and demonstration of how to make a loop with the string and gain a reward. In this trial, no apes, but 9 out of 12 children successfully imitated the procedure to gain the reward (Tennie et al., 2009:2409-2411).

2.3.12 Synthesis

Lithic technology and the origins of Mode 3 technology play a pivotal role in several interpretations about hominin evolution in the Middle Pleistocene, which are summarized above. Some of these interpretations focused on what technological change in the Middle Pleistocene indicate about hominin behavioral and cognitive capacities. Attention is paid to the significance of hafted technology and what it demonstrates about working and constructive memory, which are tied to aspects of human problem solving and planning capacities (Ambrose, 2001; Ambrose, 2010; Wynn and Coolidge, 2011). A shift in technology toward flake and blade production is also important because it marks the beginning of human levels of variation and adaptability (Foley and Lahr, 2003; Foley and Lahr, 1997; Shea, 2011b). Mode 3 is also generally associated with the manufacture of lithic tipped armatures, which indicate new strategies for hunting and could reflect enhanced capacities for resource extraction and more reliable access to food resources compared to earlier hominins (Foley and Gamble, 2009). Hafting could also demonstrate human-like capacities for grammatical language (Ambrose, 2001; Ambrose, 2010). Big game hunting, presumably carried out by the use of hafted spears, could indicate increased capacities for cooperation with non-kin beyond the nuclear and extended family, since big game hunting in modern hunter-gatherer populations is generally a cooperative endeavor (Boyd and Richerson, 2005). This capacity for nonkin cooperation is what Boyd and Richerson (2005) characterize as the 'tribal-instinct'. To summarize, Mode 3 technologies demonstrate innate capacities for problem solving, planning, cooperative behavior, adaptation to variable selective pressures, and perhaps language.

31

Some interpretations of Middle Pleistocene hominin behavior address the question of why there was technological change during this period. A direct cause for technological change is offered by Potts (1998), who argues that increased climatic variability ~700 ka could explain changes in hominin morphology and behavior in the early Middle Pleistocene. That is to say, new and flexible technologies may have served to cope with rapid changes in local environments. Diverse technologies may have also served to adapt to an increased variety of environments on a spatial scale. The degree of behavioral variability that characterizes the shift to flake and blade assemblages in the Middle Pleistocene could have resulted from diverse strategies for adapting to ecological variability (Shea, in press; Shea, 2011b). Indirect causes of technological change are provided by Foley and Lahr (2003; 1997), who link the origins of Mode 3 technology with the speciation of H. helmei. Increased encephalization and changes in cognition, which may have been selected for a variety of purposes, could have also indirectly caused technological change. The large size of the human brain and our enhanced capacities for problem solving, planning, and language could be advantageous for any number of reasons, with the technological changes of the Middle Pleistocene reflecting an exaptation of these capacities, rather than an explanation for them. For example, the social brain hypothesis posits that our large brain was selected for in order to cope with the increased cognitive demands of living in large social groups (Aiello and Dunbar, 1993; Dunbar, 1998; Dunbar, 2003). Changes in brain structure and organization in response to these social pressures could have had implications for technological behaviors and indirectly resulted in changes to how tools were manufactured and used to extract resources. Cognitive changes related to brain expansion could have also influenced changes in what Tennie et al. (2009) identify as the “ZLS”, or the range of behaviors that an individual can learn on their own, while also affecting human capacities and strategies for social learning.

Technological changes in the Middle Pleistocene may have also caused social and morphological changes. If hafted hunting technology reflects an increased reliance on high-quality resources like meat, and resulted in reduced variability in foraging returns, then the origins of Mode 3 technology could have decreased adult and juvenile mortality, with important implications for the length of childhood and the onset of senescence (Kaplan et al., 2000). In other words, technological changes that improved foraging returns could have affected hominin life history. These life history changes could also be linked to changes in the relative importance of social learning and the kinds of strategies used to transfer knowledge between the generations. In

32 addition, increased foraging returns may have resulted in increasing populations, which would in turn favor increased and more complex social interactions across longer distances and consequently, brain expansion (Foley and Gamble, 2009), and eventually promote the kind of ultrasociality that is unique to our species (Hill et al., 2009).

Other research described above presents ideas about the definition and roots of human uniqueness. Multiple traits are put forth for what makes humans unique, including social learning (Boyd and Richerson, 2005; Hill et al., 2009; Richerson and Boyd, 2005; Tomasello, 1999; Tomasello et al., 1993), cooperation (Boyd and Richerson, 2005; Hill et al., 2009; Hill et al., 2011), enhanced and/or shared intentionality (Dunbar, 2003; Tomasello, 1999), enhanced working memory (Wynn and Coolidge, 2011), constructive memory (Ambrose, 2010), cumulative culture change (Boyd and Richerson, 2005; Tennie et al., 2009; Tomasello, 1999), imitation (Tennie et al., 2009), cultural conformity (Boyd and Richerson, 2005; Tennie et al., 2009), symbolism (Henshilwood and Marean, 2003), and behavioral variability (Shea, 2011b). In some cases the roots of these traits are traced to the Middle Pleistocene and in others, to the Late Pleistocene. However, with some exceptions, the archaeological record has so far played a minor role in establishing the origins and evolution of the above traits, probably in part because of issues of archaeological visibility, and because the middle range theory connecting these concepts to archaeological evidence is underdeveloped. From an archaeological perspective, the ‘modern human behavior debate’ has so far focused mainly on archaeologically visible traits like the use of symbolic resources (Henshilwood and Marean, 2003; McBrearty and Brooks, 2000), which is related to capacities for communication and social interaction, and ‘behavioral variability’ and adaptability, which is reflected in the diverse strategies used to manufacture stone tools (Shea, 2011b). Most other aspects of human uniqueness require consideration from an archaeological perspective in order to establish how these traits might be recognized archaeologically, in order to test hypotheses about what makes modern humans unique.

2.4 Summary

The shift to Mode 3 technology involved a movement away from manufacturing large core tools such as handaxes, cleavers, and picks, toward the production of diverse types of flake tools that were probably hafted onto handles. In Eurasia, this technological change is represented by the beginning of the Middle Paleolithic and generally associated with Homo neanderthalensis. In

33

Africa, Mode 3 is represented by the MSA, the earlier portion of which is generally associated with Homo helmei (or late ‘archaic Homo sapiens’) and by anatomically modern Homo sapiens after ~200 ka. The origins and evolution of Mode 3 technology is significant because of the role it plays in accounts of hominin speciation and dispersion, the origins of behavioral modernity and variability, the development of human sociality and encephalization, and advances in cognitive facilities related to long-term planning and language.

34

3 Technological Change in the Middle Pleistocene

This chapter focuses on Middle Pleistocene lithic technology, and begins with a review and analysis of chronological data currently available for the late Acheulean and early MSA. Because the KP1 stratum 4a assemblage is assigned to the Fauresmith Industry, I also provide a summary of the current state of knowledge regarding the Fauresmith Industry, some of the problems and debates associated with the Fauresmith concept, and the role of KP1 within this context. Regardless how the KP1 assemblage is labeled, the stratum 4a KP1 lithic assemblage is chronometrically dated to the early Middle Pleistocene and documents the technological behaviors of early Middle Pleistocene hominins. This thesis focuses on three aspects of KP1 technological behavior for which I provide the necessary background information in three sections of this chapter; Section 3.3 examines blade production, section 3.4 considers functional analyses of MSA points, and section 3.5 provides an overview of raw material foraging strategies in the ESA and MSA.

3.1 Chronology of the ESA-MSA ‘transition’

Handaxes, cleavers, and/or picks characterize the Acheulean. These tools are sometimes lumped together using the terms "Large Cutting Tools" (LCTs, McNabb et al., 2004) or "bifaces". Neither term is ideal. LCT imposes the unknown and restrictive function of cutting. "Biface" is technically inaccurate because it excludes unifacially worked handaxes, cleavers, and picks, and is the same term used to describe the small finely made points characteristic of later time periods (i.e. Still bay). Here, I will use the term LCT, except when referring to specific tool forms within this category.

The African Acheulean persisted for at least a million years. Acheulean sites occur throughout most of the continent, but most chronological information about the African Acheulean comes from East Africa and South Africa. The earliest Acheulean occurrence is dated to ~1.8 Ma in West Turkana, Kenya based on sedimentation rates (Lepre et al., 2011). Other early occurrences are 1.4 Ma (Asfaw et al., 1992; McDougall and Brown, 2006).The earliest dated Acheulean assemblages in southern Africa could be as old at 1.6 Ma (Chazan et al., 2008; Gibbon et al., 2009; Kuman and Clarke, 2000; Matmon et al., 2012).

35

LCTs persisted in at least some parts of Africa until ~160 ka. Handaxes are present in the Herto Member of the Bouri Formation in the Middle Awash, Ethiopia in sediments dated to 160-154 ka based on 40Ar/39Ar (Clark et al., 2003). In , at the site of Abdur, handaxe-bearing deposits date to 125±7 ka based on U-Th dating of coral (Bruggemann et al., 2004; Walter et al., 2000), but the handaxes could also be considered large bifacial points (Chazan, personal communication).

The appearance and elaboration of new technologies (Levallois reduction, blades, and points) in the early Middle Pleistocene makes refining and determining the chronology of the later Acheulean is challenging. The discovery of assemblages containing LCTs and these new technologies, led to the definition of 'transitional' industries such as the Sangoan and Fauresmith. There are various issues with these industrial designations, and consequently, some researchers avoid them, employing the terms “Late” or “terminal ESA” or “early MSA”. For some, a single occurrence of an LCT is enough to designate an assemblage as Acheulean, whereas others will designate an assemblage as MSA if Levallois reduction is common, even if rare LCTs are present. Inconsistent terminology affects an effective consideration of chronology. The approach to chronology that I take here focuses just on the reported artifact types for the relevant assemblages, rather than the industrial designations assigned by the excavators. I will use these terms to discuss chronology in the following section:

 Acheulean assemblages: contain LCTs, and no aspects of Levallois technology, laminar (i.e., blade) technology, and no points.

 ‘transitional’ assemblages: contain LCTs, and Levallois technology, laminar technology, and/or points2. The term ‘transitional’ is used by paleoanthropologist in many ways (Riel-Salvatore, 2009). I use the term here solely to describe a type of assemblage in the African archaeological record, and I do not hold that the assemblage type indicates that hominin behavior, cognition, or social characteristics were transitional.

2 Bifacial points are distinguished from LCTs somewhat arbitrarily and based mainly on size, especially thickness In most cases, the presence of points is not contentious, because they are unifacial and manufactured on small flake blanks.

36

 MSA assemblages: characterized by Levallois technology, laminar technology, and/or points, no LCTs.

Chronometric dating methods for the Middle Pleistocene have undergone significant improvements in the last couple of decades. The current consensus that the ESA-MSA transition dates to ~300-200 ka is based on the observation that there are both ESA and MSA-designated assemblages dated to this period. However, Herries (2011) highlights that most of our chronological data for the Middle Pleistocene actually provides only minimum age estimates for the lithic assemblages of interest. A reconsideration of published dates using the designations I define above reveals that most securely dated Acheulean assemblages are older than ~500 ka in Africa. African assemblages securely dated to between ~500 and 300 ka contain both Acheulean and MSA elements of technology.

3.1.1 Dates for Acheulean assemblages

Many later Acheulean assemblages that have been dated provide minimum age estimates (Herries, 2011). At Sai Sudan, an OSL sample from sands overlying the Acheulean assemblage gives a minimum age of 223±19 ka (Van Peer et al., 2003).

At Ismila, Tanzania, an Acheulean assemblage consisting of a minimally trimmed cleaver and no MSA components (Howell et al., 1962) gave an estimated age range of 330-220 ka based on U- series analyses of bone (Howell et al., 1972), but like other uranium series analyses on open systems that include bone (Herries, 2011), these dates might be best interpreted as minimum age estimates (Herries, personal communication).

Secure ages for Acheulean assemblages in East Africa come from , Kenya. The archaeological sequence of at Olorgesailie starts at Member 1, which dates to 992±39 ka based on 40Ar-39Ar analyses. The uppermost Member in the sequence, Member 14, dates to 493 ka ± 2 ka, (Deino and Potts, 1990, 1992). The lithic assemblages from these members contain abundant LCTs, with no diagnostic MSA elements (Isaac, 1977).

37

The Acheulean assemblages at , Zambia (defined as the Bwalya Industry), contain LCTs but no prepared or Levallois cores, blades, or points3, and provide only minimum age estimates. Two U-series ages on wood from the Acheulean layers were dated to 182±16 ka and 182±10 ka (Clark, 2001:appendix D). These should be considered minimum age estimates (Clark, 2001:27) because of the open nature of the system (Herries 2011), indicating that the assemblage is older than 166 ka.

The Acheulean is represented in Excavation 1 at Wonderwerk Cave, South Africa in strata 6-10 (Chazan and Horwitz, in press), possibly extending on either side of those strata to strata 5 and 11 (Chazan et al., in press), but the lithic artifact sample in 5 is heavily weathered (Chazan, personal communication), and the size of the lithic artifact sample in 11 is too small for confident designation (Chazan, in press; Chazan et al., in press; Chazan and Horwitz, in press). Strata 5-8 were originally designated to the Fauresmith (Beaumont and Vogel, 2006), but the assemblages do not contain Levallois, blades, or smaller bifaces than the underlying strata, and so are better assigned to the Acheulean (Chazan, in press; Chazan et al., in press; Chazan and Horwitz, in press; Chazan et al., 2008). An OSL sample from stratum 9 in Excavation 1 gives a minimum age estimate of 256±21 ka for the associated and underlying Acheulean assemblages

(Chazan et al., 2008). This is a minimum age estimate because the De values are beyond the limits of OSL methods (Chazan et al., 2008). The OSL age is consistent with two minimum U- series age estimates from speleothem of >349 and >350 from the overlying stratum 6 (Beaumont and Vogel, 2006). However, it is not clear whether the dated speleothems are broken fragments or in situ growths and so it remains difficult to thoroughly evaluate the meaning of the U-series results (Herries, 2011:12). This method dates the formation of the stalactites and provides a maximum age estimate for associated artifacts if the dated items are not in situ (Herries, 2011). If the dated speleothems in strata 6 of Excavation 1 are fragments, then they provide no chronological control because strata 6 could be younger than or older than ~350 ka. Paleomagnetic and cosmogenic nuclide analyses provide better chronological information for the Acheulean at Wonderwerk Cave. Stratum 9 has normal polarity and the lower part of stratum 9 is

3 There are a few points in this assemblage, but they are rare and most can alternatively be interpreted as convergent scrapers or small handaxes (Clark et al. 2001). If one preferred to classify these assemblages as ‘transitional’ the conclusions presented here would not be changed because the dates are minimum age estimates (Clark, 2001:27; Herries 2011).

38 best interpreted as falling within the Jaramillo subchron (0.99-1.07 Ma) based on cosmogenic nuclide analysis (Matmon et al., 2012). A weak reversal signal (Chazan et al., 2008) and a major unconformity (Matmon et al., 2012) within Stratum 9 could suggest that the upper part falls within the Bruhnes chron and post-dates 780 ka, but it is also possible that it represents a continuation of the Jaramillo (Chazan et al., 2008). At this point, it is possible to assert that some of the Acheulean at Wonderwerk Cave is younger than 0.99 Ma, and some is older than 0.99 Ma. It is likely that the Acheulean in strata 6-8 and the top of stratum 9 is younger than 780 ka.

At Duinefontein 2, South Africa, U-series analysis of calcrete blocks that overlie the archaeology-bearing horizons gives a minimum age estimate of ~150-160 ka (Cruz-Uribe et al., 2003; Klein et al., 1999). The ‘subtraction method’, which combines OSL, thermoluminescence (TL), and infrared-stimulated luminescence (IRSL) to calculate an age estimate, gives range of 217-355 for the archaeological horizons (Feathers, 2002). Based on the problems with variable dose rates and the limits of feldspar luminescence signal, these dates could be considered minimum age estimates (Herries, 2011:7). Klein et al. (1999) point out similarities between the Duinefontien and Elandsfontein fauna consistent with a Middle Pleistocene age. The Elandsfontein fauna appears similar to or could be marginally younger than the Cornelian Land Age, which dates to between ~1.1 and 0.8 Ma at Cornelia-Uitzoek (Brink et al., 2012; Herries, 2011).

The Cave of Hearths Acheulean in Beds 1-3 lacks Levallois technology, blades, and points (McNabb, 2009) and securely postdates the Bruhnes-Matuyama boundary at 780 ka (Herries and Latham, 2009). Together, the paleomagnetic and ESR data suggest an age range of ~700-400 ka, but towards the older end of this range (Herries, 2011:7-8). Cave of Hearths could be one of few known ‘non-transitional’ Acheulean sites in southern Africa that securely postdates the Bruhnes- Matuyama boundary. Strata 6-8 and the top of stratum 9 at Wonderwerk Cave probably also postdates the Bruhnes-Matuyama boundary, as discussed above.

Other dated Acheulean assemblages such as Swartkrans (Herries et al., 2009), Cornelia-Uitzoek (Brink et al., 2012), Gombore II in the Melke Kunture Formation, Ethiopia (Morgan et al., 2012), Members 1-12 at Olorgesailie, Kenya (Deino and Potts, 1990, 1992), and Bodo, Ethiopia (Clark et al., 1994) are more than at least 500 ka.

39

3.1.2 Dates for ‘transitional’ assemblages

Many sites with ‘transitional’ assemblages in Africa, defined here by the co-occurrence of LCTs and Levallois, blades, and/or points, give us only minimum age estimates, but based on current evidence, these types of assemblages are at least as old as 500 ka (Kapthurin Formation, Johnson and McBrearty, 2010) and perhaps as young as 132 ka (Abdur, Bruggemann et al., 2004). Both these age estimates come from East Africa. In southern Africa, ‘transitional’ assemblages are at least as old as 435 ka (KP1, Porat et al., 2010). There is less chronological information from other parts of the continent, but at Cap Chatelier (Casablanca, North Africa) an Acheulean assemblage that also contains Levallois technology has a minimum age of 200 ka (Rhodes, 1990, as cited in Raynal et al., 2001:71) and is consistent with dates from East and southern Africa. Below is a summary of chronological information for ‘transitional’ assemblages in East Africa and southern Africa.

At Sai Island, Sudan, three deposits are identified as Sangoan, based on the presence of core-axes (Van Peer et al., 2003). ‘Foliate’ points from these levels were subjected to a use-wear analysis and interpreted as ‘projectiles’ (Rots et al., 2011). The lowest Sangoan unit, BLG, overlies ES, which is dated by OSL to 223±19 ka , so BLG is less than 242 ka. The middle Sangoan unit, TLG, is dated by OSL to 182±20 ka (202-162 ka). The uppermost Sangoan in UB is in secondary context and dated to 152±10 ka. Thus, the artifacts from UB are more than 142 ka, but less than the age for TLG. To summarize, all of the Sangoan at Sai island is younger than 242 ka and older than 142 ka.

At Herto, Middle Awash, Ethiopia the Lower Herto Member has yielded a “Later Acheulean” assemblage with handaxes and possibly Levallois technology (DeHeinzelin et al., 2000:133), and this member contains a tuff dated by 40Ar/39Ar to 260 ±16 ka (Clark et al., 2003). The Upper Herto Member yields LCTs, Levallois, points, and rare blades, and is securely constrained by 40Ar/39Ar to 160±2 ka and 154±7 ka (Clark et al., 2003).

LCTs, Levallois cores and flakes, and points were recovered from archaeological strata at site A1 at Aduma, Middle Awash that underlie the Ardu B sediments (Yellen et al., 2005). The Ardu B sediments are dated to ~90 ka based on U-series and OSL and provide a minimum age for some of the lithics at site A1.

40

Unit 9, at Gademotta, Ethiopia, yields handaxes, Levallois technology and retouched points and is capped by a tuff (Unit 10) dated to 276±4 (Morgan and Renne, 2008).

At Lake Eyasi, Tanzania, a "Sangoan" assemblage with core-axes, and possibly Levallois technology and points, is argued to have a minimum age of 141 ka based on U-series analysis of bones from nearby Mumba Rockshelter (Mehlman, 1987).

Two archaeological occurrences in the Lower Omo Valley Kibish Formation, Ethiopia date to 195±5 ka based on correlation with the dated sediments in which the earliest Homo sapiens fossils were recovered (McDougall et al., 2005; Shea, 2008). Both these localities, Kamoya’s Hominid Site and Awoke’s Hominid Site, yield typical Levallois technology and rare LCTs (Shea, 2008).

At Olorgesailie, Kenya, two members of an as-of-yet unnamed formation overlie the ~490 ka Acheulean-bearing Member 14. These are the Olkesiteti and the Oltepesi beds, and together they span the period from ca. 490 to 64 ka (Brooks et al., 2007; Deino and Potts, 1990, 1992). The beds contain both LCTs and MSA technologies (Brooks et al., 2007), though formal descriptions of these assemblages are pending.

At Simbi in the Lake Victoria Basin, Kenya, K/AR estimates on tuffs associated with a Sangoan assemblage give an age between 200 and 50 ka (McBrearty, 1991, 1992).

At Abdur, Eritrea, an assemblage containing handaxes4 and blades dates to 125±7 ka based on U-Th dating of coral (Bruggemann et al., 2004; Walter et al., 2000). This site provides the youngest date for handaxes in Africa.

The Kapthurin Formation, Kenya provides some of the best chronological control in Africa for the earlier part of the Middle Pleistocene. The MSA-designated assemblages associated with the upper levels of the Bedded Tuff (K4) section of the formation, including Koimilot and Rorop Lingop, contain Levallois cores and flakes, and unifacial handaxes (McBrearty and Tryon, 2006; Tryon, 2006). The artifacts are dated based on tephrostratigraphic correlation to less than 284±12 ka (Deino and McBrearty, 2002; Tryon, 2006). Some of the Kapthurin assemblages that are

4 Or very large bifacial points (Chazan, personal communication).

41 designated as Acheulean also include Levallois cores and flakes, blades and/or points (Deino and McBrearty, 2002; Johnson and McBrearty, 2010; McBrearty and Tryon, 2006). Two Acheulean assemblages, the Leakey Handaxe Area (LHA) and Factory Site (FS), occur in the Middle Silts and Gravel Member (K3), which is bracketed by tephra dated to between 284±12 and 509±9 ka (Deino and McBrearty, 2002). Both these assemblages contain LCTs and Levallois technology, and the LHA assemblage also contains blades (Tryon, 2006). The sites of GnJh-42 and GnJh-50 from the Lacustrine Facies of the Middle Silts and Gravel Member are bracketed by 40Ar/39Ar dates of 545 ±3 ka on underlying Tuff and 509±9 ka on the overlying Grey Tuff. These assemblages yield the earliest and confidently dated non-Levallois blades, blade cores, and the earliest minimum age estimate for a ‘transitional’ assemblage, as defined above.

At Kalambo Falls, Zambia, the Sangoan (or Chipeta) Industry assemblages contain LCTs, blades (“long triangular flakes”), points, and Levallois technology (Clark, 2001; Sheppard and Kleindienst, 1996). The only available dates at this point are infinite radiocarbon age estimates ~36-46 ka and a U-series estimate on wood of 76±10 ka, which is also a minimum age estimate (Clark, 2001). The ages for the underlying Acheulean are also minimum estimates (see above), so they do not provide chronological control for the overlying levels.

“Early MSA” containing LCTs (core-axes), points, and backed pieces were recovered from the F and A blocks at Twin Rivers, Zambia, (Barham, 1998, 2002; Barham and Smart, 1996). For the F-block, U-series of speleothem indicate a minimum age of 195±19 ka based on mass spectrometric methods (Barham, 1998), or >230 +35/-28 ka based on conventional methods (Barham and Smart, 1996). The older A block dates to between 266 and >400 ka (Barham, 2002).

At Rooidam 1, South Africa, the assemblage was designated as Fauresmith based on the co- occurrence of LCTs, Levallois cores and flakes, ‘flake-blades’, and points. As Tryon (2003) points out, the points are only concentrated in one excavated spit. The other tool-types are distributed relatively evenly through the deposit, however. The densest artifact concentration occurs in unit B (Butzer, 1974), which is the equivalent of levels 12’ 6’’ – 18’ 6’’ reported by Fock (1968) and stratum 9 reported by Beaumont (1990f). This unit underlies unit C, which was dated using uranium-series methods. However, recrystallization gave only a minimum age for unit C that is too young (Th: 108±9 ka, Pa:108+40/-20), based on the U-series dates from unit G

42 higher in the sequence (Szabo and Butzer, 1979). Unit G (Butzer, 1974) is equivalent to approximately 5' feet below surface as reported by Fock (1968), and stratum 4 as reported by Beaumont (1990f). There is no evidence for recrystallization in Unit G according to Szabo and Butzer (1979), and U-series analysis gives a Th age of 174±20 and a Pa age of 151+∞/-35. These ages could date the associated artifacts, but the open nature of the system suggests that both samples may represent minimum age estimates (Herries, 2011:14). Even Fauresmith artifacts recovered from above Unit G in Stratum 1 could have an actual age greater than ~174 ka.

At Bundu Farm, South Africa, the assemblage from Groups 4-6 are designated as Fauresmith and the overlying Groups 2-3 are designated as MSA (Kiberd, 2006). Groups 4-6 are characterized by prepared cores and flake-blades distributed evenly throughout the section. One LCT and 3 worked points were found at the bottom of the sequence in Group 6. Group 4 was dated using coupled ESR/U-series on teeth to 145.7±16 ka. Unprovenienced tooth samples from earlier excavations (most likely from Groups 4 and 5) give an age range of 144-371 ka. This means that for Group 6, which is the only unit documenting the coexistence of LCTs, prepared cores, worked points, and flake-blades, there is a minimum age estimate of ~371 ka (Herries, 2011:12). The overlying units with no LCTs or points could be as young as ~129 ka.

At Wonderwerk Cave, South Africa, Fauresmith-designated assemblages were recovered from Excavation 1 near the front of the cave (Beaumont and Vogel, 2006; but c.f. Chazan, in press; Chazan et al., in press; Chazan and Horwitz, in press; Chazan et al., 2008), 2, 3, and 4 in the middle of the cave (Beaumont and Vogel, 2006), and 6 at the back of the cave (Beaumont and Vogel, 2006; Chazan and Horwitz, 2009; Chazan and Horwitz, in press). Strata 3-5 in Excavation 2 provided three U-series ages on speleothems recovered from these sediments (Beaumont and Vogel 2006). As was the case for the Acheulean in Excavation 1 discussed above, it is not clear whether these are broken fragments or in situ growths (Herries, 2011:12). Two stalactites, presumably fragments, recovered from the Fauresmith-bearing stratum 3 in Excavation 2 gave age estimates of 276±29 and 278±26 ka. Thus, the Fauresmith artifacts in stratum 3 of Excavation 2 are younger than 304 ka. A speleothem from stratum 4 gave an estimate of 286±29 ka. Assuming this is a fragment and not in situ, the Fauresmith –designated assemblage in this level is less than 315 ka. However, recent investigations at Wonderwerk Cave have not yet confirmed the Fauresmith-designation of these strata (Chazan, personal communication).

43

No Fauresmith-bearing strata have been identified by the recent investigations of Excavation 1 (Chazan, in press; Chazan et al., in press; Chazan and Horwitz, in press; Chazan et al., 2008), but Beaumont (Beaumont, 2004b; Beaumont and Vogel, 2006) attributed strata 5-8 in Excavation 1 to the Fauresmith. Even if one prefers to accept the Fauresmith-designation for strata 5-8, the only chronological information currently available is that it probably post-dates the Bruhnes- Matuyama boundary 780 ka, and certainly post-dates the Jarimillo 0.99 Ma (Chazan et al., 2008; Matmon et al., 2012, see above).

Chazan and Horwitz (2009) argue that all the Excavation 6 material at Wonderwerk Cave should be classified as Fauresmith and that there is no MSA in this area of the cave, despite claims by Beaumont and Vogel (2006) for MSA in the upper part of stratum 3. Stratum 3 in excavation 6 gave a minimum U-series date of >187±8 ka (Beaumont and Vogel, 2006). If this is an in situ stalagmite, then the deposit must be older than ~187 ka. If it is a fragment, then the fragment formed sometime beyond 179 ka, and was deposited in the sediment sometime after it formed (Herries, 2011). This gives us no chronological information. The best chronological information that we have for Excavation 6 is that the lithic-bearing sediments have normal polarity and are thus less than 780 ka (Chazan and Horwitz, 2009; Matmon et al., 2012).

3.1.3 Dates for MSA assemblages

The earliest age estimate for a MSA assemblage may come from Cartwright's and Wetherill's sites on the Kinangop Plateau, Kenya, which produced controversial K-Ar dates of >440 ka and <557 ka for 'pseudo-Stillbay' occurrences there (Evernden and Curtis, 1965), but there are issues with the relationship between the dated tuffs and the occupation horizons and this date is not generally accepted (McBrearty and Brooks, 2000:489). The dating question could be resolved by continuing research at Cartwright's site where numerous bifacially worked points5 were recovered from recent excavations and systematic surface collections (Waweru, 2002). A K-Ar date of 240 ka was also reported from Malewa Gorge, Kenya (Evernden and Curtis, 1965), but no excavations have been conducted in that area.

5 Dimensions for the points are not given separately, but the maximum dimension for all recovered artifacts from Cartwright’s site is 60 mm, most are less than 20 mm (Waweru, 2002), thus the points are smaller than handaxes.

44

The site of Florisbad, near Bloemfontein, South Africa gives an early age estimate for an MSA assemblage. OSL and ESR age estimates suggest an age of 279±47 ka (326-232 ka) for the MSA-bearing basal layers (Units N, 0 & P) of the site (Grün et al., 1996; Kuman et al., 1999).

In northern Africa, multiple dating methods (OSL, AAR, ESR, TL & U-series) indicate that the MSA at Bir Tarfawi and Bir Sahara East in southwestern Egypt, characterized by Levallois technology and bifacial points, begins ~230 ka (Wendorf et al., 1993).

MSA assemblages at Gademotta and Kulkuletti, Ethiopia associated with sediments between Unit 10 and Unit D are at least 183±10 ka based on 40Ar/39Ar analysis (Morgan and Renne, 2008). As mentioned above, the earliest levels at this site are capped by a tuff dated to 276±4 ka (Morgan and Renne, 2008). Except for the lowest level at this site, handaxes are absent, and there are small bifacial points.

Beaumont reports MSA deposits at Wonderwerk Cave in Excavation areas 2, 3, 5, 6 and 7 (Beaumont and Vogel, 2006). There are several U-series dates on speleothems from MSA- bearing strata in each of these excavation areas, and these dates range from ~73-220 ka (Beaumont and Vogel, 2006). If we assume the speleothems are fragments and not found as in situ growths (Herries, 2011, see above), then each provides a maximum age for the deposit they were recovered from. Excavation 2, stratum 2 gives the oldest age estimate of 220±14 ka, and the MSA-designation for this assemblage has been confirmed by recent investigations (Chazan and Horwitz, in press). Thus, all the MSA assemblages in Wonderwerk Cave are probably younger than ~220 ka. The youngest MSA deposit identified by Beaumont is less than 73±5 ka (Beaumont and Vogel, 2006). This age estimate comes from Excavation 5, stratum 2, which has not been yet been analyzed by the current research team (Chazan, personal communication).

Thermoluminescence (TL) dates from Mumbwa , Zambia, suggest that the basal MSA levels there are >172 ka (Barham 2000, as cited in McBrearty and Brooks, 2000). The lowermost excavated layers at , South Africa, Unit 5 WA, contains MSA artifacts, with ESR dates on tooth enamel of 174±9 ka and 227±11 ka (Grün and Beaumont, 2001).

MSA assemblages at Lincoln Cave have been dated using uranium series methods to between 253±36 and 115±8 ka years old (Reynolds et al., 2007).

45

Other dated MSA assemblages, such as those at 13B (Jacobs, 2010; Marean et al., 2007) and Klasies River (Feathers, 2002), date to less than 170 ka.

3.1.4 Synthesis of Acheulean, ‘transitional’, and MSA chronologies

Based on current evidence, assemblages that we confidently know date to between ~500 and 300 ka in African appear to contain LCTs and Levallois technology, blades, and/or points (Figure 1), and are ‘transitional’ based on the terminology employed here. By using the term ‘transitional’ I do not intend to imply that hominin behavior itself is necessarily ‘transitional’, but just that the lithic assemblages during this period contain diagnostic elements of both the ESA and MSA. There are many possible explanations for the co-occurrence of ESA and MSA fossil directeurs in lithic assemblages that I discuss further in Chapter 8. The assemblages I identify here as ‘transitional’ include the sites in the Kapthurin Formation, Kenya, where there is good chronological control and assemblages described as ‘Acheulean’ contain blades and/or Levallois technologies and assemblages described as ‘MSA’ contain LCTs. There may be no assemblages with only LCTs younger than ~500 ka. Dated Acheulean assemblages with apparent ages younger than this, such as Duinefontein 2 and Ismila, could be minimum age estimates. The earliest MSA assemblages without LCTs, such as the basal levels at Florisbad, are no older than ~300 ka. Some assemblages that post-date this time also appear to contain LCTs, and one of these sites, Abdur, is younger than ~130 ka. Research focusing on the so-called ESA-MSA “transition” have been focusing the time period between 300-200 ka, when LCTs first disappear from some assemblages. But, if we are interested in when these new technologies first appear, we need to be looking at the time period between roughly 500 and 400 ka (Herries, 2011).

3.2 Defining and Evaluating the Fauresmith Industry

3.2.1 Definition

The Fauresmith industry was originally defined in the first half of the twentieth century, based on surface collections and field observations in what was called the Fauresmith district and neighboring areas in the Free State, South Africa (Goodwin and Van Riet Lowe, 1929). The original definition of the Fauresmith focused on small finely made handaxes on flakes with an S- shaped twist or curve, slightly trimmed flake points, rare cleavers, and scrapers (Goodwin and Van Riet Lowe, 1929:72-85), while later discussions emphasized the co-occurrence of handaxes

46

Figure 1 Summary of chronology for three types of Middle Pleistocene assemblages. The three types of assemblages are those with LCTs, (Acheulean), those with LCTs, and blades, points and/or Levallois technology (‘transitional’ assemblages), and those with blades, points, and/or Levallois technology, but no LCTs (MSA). and Levallois points (Beaumont and Vogel, 2006), and prepared cores and blades (Beaumont, 1990b:79). Based on the co-occurrence of LCTs, and MSA elements, the Fauresmith assemblages were seen as representing a transitional industry between the core-tool dominated ESA and the flake-tool dominated MSA (Goodwin and Van Riet Lowe, 1929:83-84). However, through the years descriptions of Fauresmith assemblages have been inconsistent and contradictory (Underhill, 2011). Many researchers have expressed concerned over the Fauresmith concept (e.g. Humphreys, 1970; Sampson, 1972) , sometimes highlighting the fact that members of the Terminology Committee of the Pan-African Congress of in 1961 formally agreed that the term should be dropped (Sampson, 1972:52).

A large number of sites have been attributed to the Fauresmith Industry. In the Atlas of African Prehistory (Clark, 1967), an overlay marks the location of ~170 ‘transitional’ sites designated as

47

“Levallois-Mousterian”, “Acheulio-Levallois”, and Fauresmith. One hundred and sixteen of these sites are named and coordinates are supplied. Site locations and industrial attributions to create this map were based on regional consultations, including C.K. Cooke, J. G. Fock, B. D. Malan, H. J. Deacon, and compilation was done by Charles M. Keller, who visited museums at Kimberley, Bloemfontein, Pietermaritzburg, and Durban. However, there are no descriptive or quantitative data available for these assemblages and there are serious concerns about accepting their designation as Fauresmith (e.g. Sampson, 1972:52).

Fewer Fauresmith-designated sites have been formally reported. Table 3 provides a list of 27 Fauresmith occurrences reported in other publications, indicating the site type, the stratigraphic context of the assemblage, the types of artifacts present, dominant raw material types, and age estimates, if available. Figure 2 shows the location of the sites.

3.2.2 The Fauresmith in the Free State

The Fauresmith type site, Brakfontein, is located in the Free State and the original definitions of the Fauresmith industry are based on assemblages from this area (Goodwin and Van Riet Lowe, 1929). Only four sites from this region are listed in Table 1; Brakfontein, Fauresmith (Town Lands), Lockshoek, and Onder Dwars Rivier, because these are the only sites discussed in any detail by Goodwin and Van Riet Lowe (1929), even though 18 additional sites are listed. Clark (1967) and Humphreys (1970) also map several more sites in this region, but I was unable to locate published records for these additional sites. The four sites mapped here and discussed by Goodwin and Van Riet Lowe (1929) all contain LCTs. These LCTs are described as small and finely-made handaxes manufactured on flake blanks with an S-shaped twist. According to (Goodwin and Van Riet Lowe, 1929:72), the average length for a Fauresmith handaxe is 108 mm, compared to 170 mm for the Acheulean handaxe. Other tools types that are mentioned are slightly trimmed flake points, rare cleavers, and scrapers. A burinated blade on hornfels from the site of Fauresmith (Town Lands) is also illustrated by Van Riet Lowe (1933). When identified, hornfels was usually the main raw material type used at the Free State Fauresmith sites. All of the assemblages in the Free State come from surface collections at unsealed open air locations. The sites all yielded LCTs, and some seem to have also yielded retouched points, but it is not clear whether blades or prepared cores were recovered from any of these localities.

48

Table 3 Summary of published Fauresmith-designated assemblages.

Site Location SiteType Surface Fauresmith- Stratigraphic Raw Age References Notes Collection/Excavatio bearing Units Context of 6 Estimate

LCTs Material

Cores Points ns Fauresmith Blades

Assemblage Prepared Retouched Retouched

Bestwood 1 Northern open air Surface collections n/a Lying on top of yes yes no? yes banded n/a (Chazan et al., Cape and excavations, gravel under sands ironstone 2012b) Chazan 2011 Biesiesput 1 Northern pan Surface Collections stratum 2 and 3 underlying LSA in yes yes yes yes hornfels n/a (Beaumont, Cape Fock and Humphreys stratum 1 1990f; 1971, Excavation Beaumont and Beaumont 1983 Vogel, 2006) Brakfontein Free State open air Surface collections, n/a n/a yes ? yes? ? hornfels n/a (Goodwin and (No. 231) Lowe 1926 Van Riet Lowe, Fauresmith 1929) Type Site Bundu Farm Northern pan Excavation, Kiberd Groups 4-6 underly MSA yes yes yes yes mainly 144-371 (Kiberd, 2006) only one LCT Cape 1998-2003 Groups 2-3 ka recovered from lowest Fauresmith group (see discussion of this site in chronology section) ?Bushman's Limpopo rock- excavations 1965 Layers 28-43 stratified below yes? yes yes yes hornfels n/a (Louw, 1969) handaxes recovered Rockshelter shelter LSA from layer 34 and 41, interpreted as intrusive, Herries (2011) cites this site as potential Fauresmith site. Canteen Northern Open air, Excavations, base of "Hutton Hutton sands yes yes yes yes mainly n/a (Beaumont Koppie Cape koppie, Beaumont and Sands" and top contain MSA, andesite and McNabb, Vaal river McNabb, Kuman of Stratum 2a gravels (strata 2a 2000; Kuman, (Younger and 2b) contain in press; Gravels II, or Acheulean (2a is McNabb and Rietputs A) designated as Beaumont, Victoria West) 2011)

6 In some literature, the terms indurated shale, lydianite, and baked shale, are used to describe hornfels; Ventersdorp Lava is used to describe andesite; and jasper or jaspilite is used to describe what I refer to as banded ironstone. In most cases these identifications require further investigation.

49

?Cave of Limpopo cave Excavations, Gardner Beds 1-3 underlies MSA yes no no no mainly ~700-400 (McNabb, McNabb (2009) Hearths and Kitchen 1947-49, strata (Bed 4, see quartzite ka 2009) reports only 10 Mason 1953-1954 below) (Herries, blades and 2011) designates the Beds 1-3 assemblages as Acheulean without MSA technologies Bed 4 overlies Beds 1-3 ? yes yes yes various, n/a (Sinclair, 2009) Sinclair (2009) does (Acheulean) and mainly not distinguish this underlies MSA quartzite bed from the strata overlying MSA beds. ?Cofimvaba Eastern open air Surface collections, n/a sealed by yes yes yes ? shale/ n/a (Goodwin and "long slender Cape 1925 overlying gravels hornfels Van Riet Lowe, parallel-sided flakes and thick deposit 1929) (up to 5 inches in of sandy loam length" pg. 93 - interpreted here as blades Fauresmith Free State open air Surface collections, n/a water-borne yes yes yes? ? hornfels n/a (Goodwin and Van Riet Lowe gravels Van Riet Lowe, 1929; Van Riet Lowe, 1933) Inhoek 1 Northern Vaal river Surface collection, n/a thin spread of yes ? ? yes hornfels n/a (Sampson, Cape Sampson gravel 1972) ?Kalambo Zambia/ open air, Excavations 1956- Ocherous sands stratified below yes yes yes yes mainly n/a (Clark, 1954, Now designated as Falls Zimbabwe Kalambo 1966 at Sites A-D MSA and above quartzite, 2001; Sangoan/Chipeta River ESA at sites A and also Sheppard and (Sheppard and B, below MSA at Kleindienst, Kleindienst, 1996; sites C and D 1996) Clark, 2001) KP1 Northern pan/ Excavation, Stratum 4a underlies MSA yes yes yes yes mainly 682-435 (this study, Cape doline Beaumont 1978- stratum 3, overlies banded ka Beaumont, 1990, Dating Porat et Acheulean ironstone 1990b; al. 2004 stratum 4b Beaumont, 2004a; Porat et al., 2010; Wilkins and Chazan, 2012) Lockshoek Free State open air Surface collections, n/a water-borne yes ? yes? ? hornfels n/a (Goodwin and Lowe 1926 gravels Van Riet Lowe, 1929) ?Montagu Western cave Excavations, Isaac Layer 3 underlies MSA yes ? ? ? quartzite n/a (Keller, 1973) Cave Cape 1919, Keller 1964-65 (HP) stratum, overlies Acheulean Muirton Northern open air, Excavations, n/a at base of red yes ? ? ? quartzite (Helgren, Cape base of Sampson 1963, sands, on top of 1978; koppie, Humphreys 1968 calcified sands (ie. Sampson, Vaal river within Riverton 1972) Formation)

50

Nooitgedach Northern open-air, Surface collections n/a gravels layer at yes yes yes yes chert, n/a (Beaumont, t 2 Cape Vaal river base of "Hutton quartzite 1990c) Sands" Onder Dwars Free State open air Surface collections, n/a n/a yes ? yes? ? ? n/a (Goodwin and Rivier Lowe 1926 Van Riet Lowe, 1929) Pneil 1 Northern open-air, Surface collections, Stratum 3 - underlying MSA, yes yes yes yes andesite n/a (Beaumont, Fauresmith' (Power's Cape base of Power 1930-1955, ferruginized overlying 1990d) separated from Site) koppie, Beaumont 1984- sand and Acheulean Acheulean based on Vaal river 1985 cobbles weathering (surface of Rietputs) Pneil 6 Northern open-air, Excavation and Stratum 3 - Excavated from ? yes yes yes hornfels n/a (Beaumont, unclear how this site Cape base of surface collections, colluvium stratum 3, that 1990e) relates to "Pniel site" koppie, Beaumont 1984 based on survey discussed by Helgren Vaal river and surface (1978), stratum 3 also collections contains backed underlies MSA and pieces? (Beaumont overlies ESA and Vogel, 2006: supplementary information) Riverview Northern open-air, Surface collections, n/a Fauresmith yes yes yes yes ?hornfels, n/a (Malan, 1947; stratigraphic Estates Cape Vaal river Van Riet Lowe 1934 associated with andesite, Van Riet Lowe, observations based sands that overly diabase? 1935, 1937) on diamond Acheulean-bearing shafts (Victoria West) gravels, MSA artifacts at the top of the sands Rooidam 1 Northern pan Excavations Fock, Stratum 3-9 separated from yes yes yes yes mainly >174 ka (Beaumont, points only occur in Cape 1964-5 (Beaumont, overlying MSA hornfels 1990f; Butzer, one excavated spit 1990e), B-H stratum by 1 m of 1974; Fock, (Butzer, 1974), sterile sediments 1968; Szabo ~5'-18’ 6’’ and Butzer, (Fock, 1968) 1979) Rooidam 2 Northern pan Surface Collections, n/a Fauresmith yes yes ? yes mainly n/a (Beaumont, Cape Beaumont artifacts hornfels 1990f) associated with calcreted stratum Roseberry Northern open-air Surface collections n/a on bedrock yes yes yes yes hornfels (or n/a (Beaumont, Plain 1 Cape underlying felsite?) 1990c) "Hutton Sands" Sheppard Northern open air, Surface collections C and D1 gravel mixed with MSA yes ? yes ? hornfels n/a (Van Riet Island Cape Vaal river beds material?, Lowe, 1929) overlying ESA in D gravels

51

van der Elst Guateng open air Surface collections, on surface of yes no? no? yes diabase (van der Elst, Donga van der Elst sandy , ? 1950) (Badfontein) "Fauresmith"- bearing sediments overly ESA-bearing gravels, and underly calcified sandy-clay with MSA tools Wonderwerk Northern cave Excavations, Malan Excavation 2 In Excavation area yes yes yes yes banded <315 ka?, (Beaumont, current research calls Cave Cape and 1943- strata 3-5?, 2, underlies MSA?, ironstone certainly 1990g, 2004b; into question some of 1944, Butzer 1974- Excavation 3 In Excavation area and less than Beaumont and the original industrial 1977, Beaumont and strata 4-6?, 3, underlies MSA?, chert 780 ka Vogel, 2006; designations of Thackeray 1978- Excavation 4 Chazan, in Beaumont 1979, Beaumont stratum 4?, press; Chazan (Beaumont, 2004b; 1978-2004, Chazan Excavation 6 et al., in press; Beaumont and Vogel, 2004-pres strata 3UP-5 Chazan and 2006), see text for Horwitz, 2009; discussion Chazan and Horwitz, in press; Chazan et al., 2008; Matmon et al., 2012)

52

Figure 2 Map of Fauresmith-designated sites mentioned in text. A. Close up of sites in Vaal River Valley. B. Close up of sites in Fauresmith district in the Free State.

53

3.2.3 The Fauresmith in the Northern Cape

3.2.3.1 The Vaal River Valley

Many Fauresmith-designated sites have been identified in the Vaal River Valley in the area northwest of Kimberly in the Northern Cape. These Vaal River Valley sites include Canteen Koppie, Pniel 1 (Power’s Site), Pniel 6, Nooitgedacht 2, Muirton, and Riverview estates (Table 3, Figure 2). Sheppard Island was also located on the Vaal River approximately 150 km upstream from the majority of the other Vaal River sites, but has since been flooded due to damming. Van der Elst Donga is located >250 km upstream from Sheppard Island, but still within the Vaal River basin. Roseberry Plain is located about 25 km east of the modern Vaal River, and Rooidam and Biesiesput 1 are about 30 km south. Inhoek 1 is also located on the Vaal River, but its exact location could not be mapped.

The Vaal River Fauresmith sites are all open air sites. Some are associated with pan margins and some are located at the base of koppies. In most cases, the Fauresmith assemblages in the Vaal River Valley were manufactured predominately on hornfels but there are exceptions, with some assemblages predominately manufactured on andesite, chert and/or quartzite. In general, the Vaal River Fauresmith assemblages contain LCTs, blades, points, and prepared cores, though in a few cases it was not possible to ascertain from the reports whether certain tool types were present or not. Those cases are marked with a question mark in the particular tool type column of Table 3.

3.2.3.1.1 The Vaal River sequence

Vaal River Valley surveys in the early led Van Riet Lowe to recognize that Fauresmith deposits were stratified between Earlier Stone Age and Middle Stone Age deposits at several localities (e.g. Van Riet Lowe 1937, 1952) and he argued that there were stratified variants of the Fauresmith Industry evidenced in these deposits. Using a combination of geological and archaeological evidence, Van Riet Lowe defined a fairly long stranding chronological sequence for the Vaal River deposits and their associated industries. According to Van Riet Lowe, the Basal Older Gravels consist of Ventersdorp Lava cobbles, as well as some , hornfels, and rare other types. No artifacts are associated with the Basal Older Gravels. The Older Gravels contain higher concentrations of quartzites and hornfels, and were also called "Red Gravels" or "Potato Gravels". These Older Gravels were covered by red Kalahari sands, and seem to be associated with Oldowan and/or early Acheulean artifacts. The

54

Younger Gravels consist mainly of Ventersdop Lava cobbles, but also some quartzites, hornfels, and rare other types of raw material largely from Dwyka tillites. The Younger Gravels are associated with Acheulean artifacts and were also covered by sands and silts in many areas. What Van Riet Lowe identified as Lower Fauresmith assemblages occur mainly at the peneplane between some of the overlying sands and calcified sands, and the Upper Fauresmith supposedly occurrs within the upper levels of the sands. The MSA occurrs in the uppermost levels of the sands. The Youngest Gravels occur mainly in tributaries, and as hill-washed , and the kinds of cobbles depends on the immediate catchment area. The Lower Fauresmith was also recovered from the Youngest Gravels.

Van Riet Lowe’s Vaal River Valley sequence was constructed by correlating unconnected deposits across the river valley. The correlation of deposits between sites that are not physically linked by continuous terraces is extremely problematic, and the sequence of events that resulted in the laying down of gravel terraces, their , their subsequent by sand and other gravel deposits, is not necessarily represented at each site the same way. A revised alluvial history for the Vaal was proposed by Helgren (1978), who designated the Acheulean-bearing gravel deposits as Rietputs A-C and the overlying deposits as Riverton I-V. Some of the deposits that Van Riet Lowe (1937) designated as Younger Gravels are equivalent to the Rietputs, and some Youngest Gravels are equivalent to Riverton III and IV. More details about correlations between the alluvial of Helgren (1978) and Van Riet Lowe (1937) are presented by Wadley and McNabb (2009). A more conservative, simplified, and fairly well-established sequence for the Vaal River is (1) the Older Gravels (Windsorton Formation), (2) the Acheulean- bearing Younger Gravels (Rietputs Formation), (3) the sands of the Riverton Formation and the Youngest Gravels in modern tributary beds (de Wit, 2008; de Wit et al., 2000). Near Windsorton, the Rietputs Formation is dated by cosmogenic nuclides to 1.57 ± 0.22 Ma (Gibbons et al. 2009).

At Canteen Koppie, Acheulean artifacts are associated with gravels that are interpreted as the Younger Gravels (Rietsput Formation) or colluvial sediments (de Wit, 2008:55) and Fauresmith artifacts are associated with upper most levels of those gravels and the junction between the “Hutton Sands" (possibly the Riverton Formation) and the Younger Gravels (Beaumont, 1990a; Beaumont and McNabb, 2000; McNabb, 2001). MSA artifacts are recovered from within the sands above the Fauresmith artifacts (Beaumont and McNabb 2000, de Wit 2008). At other Vaal River sites, Fauresmith artifacts are associated with either the sands, the base of the sands, or the

55 top of gravels (e.g. Riverview Estates, Sheppard Island, Pneil 1 and 6, Muirton, Nooitgedacht 2, Table 3), while Acheulean artifacts are associated with the gravels lower in the interpreted stratigraphy at sites like Riverview Estates, Sheppard Island, and Pniel 1 and 6. MSA artifacts are in the upper levels of sands at Riverview Estates and Pniel 1 and 6.

3.2.3.2 Bundu Farm

The Bundu Farm site is located on a pan in the Upper Karoo region of the Northern Cape about 19 km southwest of the modern Orange River (Figure 2). Excavations were carried out by Kiberd between 1998-2003. Twenty-six trenches were opened near and within the pan. Seven main sediment horizons, i.e., ‘Groups’ are recognized across the trenches. Groups 4-6 contain the “Late Acheulean or transitional ESA/MSA” assemblage (Kiberd, 2006:195) that underlie the MSA in Groups 2-3. The Group 4-6 assemblage is manufactured mainly on locally-available quartzite, contains prepared cores and ‘flake-blades’. One large LCT and three points were recovered from Group 6. Group 6, which shows the most superficial affinities with other Fauresmith-designated assemblages, and as discussed in section 3.1.2, has a minimum age of ~371 ka (Herries, 2011; Kiberd, 2006).

3.2.3.3 The Kuruman Hills and surrounds

The Kuruman Hills are located about 150 km northwest of Kimberley in the Northern Cape and this region is the setting for KP1 (Figure 2). KP1 is discussed in detail in Chapter 4. Other Fauresmith-designated sites nearby are Wonderwerk Cave and the newly reported site of Bestwood 1. Kathu Townlands has not been designated to the Fauresmith, but because of its proximity and potential significance as a raw material quarry site, it will be discussed here as well. Furthermore, Beaumont and Vogel (2006, supplementary information) present handaxes, blades, and prepared cores from this site, and the combination of these tool types is consistent with a Fauresmith designation.

The well-known site of Wonderwerk Cave is located on the eastern flank of the hills and may be the only cave located in this area. The earliest occupation in the cave dates to ~2 Ma (Chazan et al., 2008; Matmon et al., 2012). The site has a long history of research, including excavations by Malan and Wells (1943-1944), Butzer (1974-1977), Beaumont and Thackeray (1978-1979), and Beaumont (1978-2004). Current investigations since 2004 are led by Michael Chazan and Liora

56

Horwitz. There are seven excavation areas, and the Fauresmith was originally identified in five of them (Beaumont and Vogel, 2006). Recent investigations call the Fauresmith-designation in some parts of the cave (i.e., Excavation 1) into question because of the lack of Levallois, small handaxes and blades (Chazan, in press; Chazan et al., in press; Chazan and Horwitz, in press; Chazan et al., 2008). However, the entire sequence in Excavation 6 is now attributed to the Fauresmith (Chazan and Horwitz, 2009), even though Beaumont attributed the top of stratum 3 to the MSA (Beaumont and Vogel, 2006). The Fauresmith at Wonderwerk cave is manufactured on banded ironstone and black chert from the immediate area outside the cave.

The Fauresmith-designated deposits at Wonderwerk Cave are of further significance because of potential evidence for symbolically-mediated behaviors. Banded ironstone slabs with linear markings were recovered from the back of the cave in Excavation 6 (Chazan and Horwitz, 2009), but nearly all these linear markings are now attributed to geogenic processes (Jacobson et al., 2011; Jacobson et al., in press). Nonetheless, quartz crystals, small rounded pebbles, and specularite have also been recovered from Excavation 6, and because these kinds of items are generally considered non-utilitarian, their presence supports the idea that the back of the cave could have been occupied because of its unique sensory properties (Beaumont and Vogel, 2006; Chazan and Horwitz, 2009; Chazan and Horwitz, in press).

Bestwood 1 is a newly discovered site located near the town of Kathu between two hills at the northernmost edge of the western flank of the Kuruman Hills (Chazan et al., 2012b). The site has only been preliminarily investigated, but yields LCTs (including finely-made handaxes), Levallois cores, and rare blades on mainly banded ironstone. The artifacts appear to be in primary or near-primary context and hold the potential for broad horizontal exposures, and the overlying sands are suitable for OSL analysis to provide a minimum age estimate for the site. Lithic artifacts are abundant throughout the immediate area around Bestwood 1, especially on the immediately adjacent hills both to the east and west.

Kathu Townlands (or Uitkoms) is rich surface scatter on the banded ironstone bedrock located just outside the town of Kathu, between KP1 and Bestwood 1. Two excavations were carried out there by Beaumont in 1982 and 1990 (Beaumont, 1990b; Beaumont, 2004a). The richness of the site and reported high frequency of handaxe roughouts suggest that the site could have served as quarry/workshop. Artifact types include “smallish” handaxes, prepared cores, blades, but no

57 points (Beaumont, 2004a:52). The site has a large area, potentially extending nearly continuously to the hills immediately west of Bestwood 1.

3.2.4 Temporal Fauresmith variants

Van Riet Lowe (1937, 1952) and Beaumont and Vogel (2006) identify variants within the Fauresmith, supposedly representing chronological trends. Van Riet Lowe's variants were tied largely to the Vaal River sequence that he defined, the details of which are now mostly discredited (e.g. Helgren, 1978; Sampson, 1972:52). As discuss above, the Lower Fauresmith was recovered from three contexts at various localities; the Youngest Gravels, the top of the Younger Gravels, and the base of red sands that sit on top of calcified sands on higher terraces (Van Riet Lowe, 1937, 1952). Van Riet Lowe's (1937) Upper Fauresmith was recovered from within the red sands on a lower at Riverview Estates. The differences between these variants appears to have been the quality of the handaxes and the Levallois cores, with more finely made handaxes and more typical cores characterizing the latter period. However, because of the problems with linking the stratigraphy across all the Vaal River sites, and because assemblages representing these variants were never actually fully reported, little can be done to evaluate the existence or significance of these variants.

Beaumont and Vogel (2006) also identify three chronological variants of the Fauresmith as represented at Wonderwerk Cave. It is not clear if, or how, these variants related to Van Riet Lowe's scheme. Beaumont’s Major Units 5-9 (Excavation 1, strata 8-11) were tentatively designated as Early Fauresmith or Acheulean based on the presence of a refined handaxe, cleaver, and a prepared core, but there is no mention of blades or points (Beaumont and Vogel, 2006). However, in the supplementary information, Biesiesput 1 is cited as an example of an Early Fauresmith site and illustrations include blades and points. The Major Unit 4 deposits at Wonderwerk Cave are designated as Middle Fauresmith and the assemblage includes small refined handaxes, blades, points, and prepared cores (Beaumont and Vogel, 2006). Other Middle Fauresmith sites supposedly include Roseberry Plain/Samaria Rd., KP1, Pniel 6, and Brakfontein (Beaumont and Vogel, 2006:222, supplementary fig. 12). The Major Unit 3 deposits at Wonderwerk Cave are identified as Late Fauresmith and include less refined and larger handaxes compared to the small refined handaxes of the Middle Fauresmith. Again, assemblages representing these variants have never been fully reported and Chazan et al. (2008) identify the

58 assemblage from strata 6-10 in Excavation 1 at Wonderwerk as Acheulean, and call into question the Fauresmith designation for the front part of the cave where these supposed variants are stratified. There may be chronological trends within the Fauresmith, but the terms “Early”, “Middle”, and “Late” Fauresmith should be avoided until empirical data demonstrates it.

3.2.5 Fauresmith outliers

The majority of Fauresmith-designated assembles are situated in central South Africa, mainly in the Northern Cape and Free State, but the term Fauresmith was applied to a few sites that occur outside this area. In most cases, researchers have abandoned the Fauresmith designation for these sites. For others, there has been little research done to establish true affinities with other Fauresmith-designated assemblages.

The lithic assemblage recovered from Beds 1-3 at Cave of Hearths, Limpopo Province, South Africa dates to ~700-400 ka (see Herries, 2011), and was originally identified as Fauresmith by Mason (1959), who later called it Acheulean (Mason, 1962). Beaumont and Vogel (2006) highlight the presence of blades in this assemblage, but McNabb (2009) reports only 10 blades and designates the Beds 1-3 assemblages as Acheulean, with no evidence of Levallois reduction. Beaumont and Vogel also (2006) suggest that Bed 4 is Fauresmith because of the presence of handaxes (citing Sampson, 1974a:157-158), but Sinclair (2009) does not distinguish this bed from the overlying MSA beds.

Montagu Cave is a multicomponent ESA-MSA site located in the Western Cape. At this site, Layer 3 underlies an MSA deposit with Howiesons’ Poort affinities, and overlies a sterile unit and then an Acheulean assemblage in Layer 5 (Keller, 1973). The Layer 3 assemblage is characterized by LCTs, ‘discoid’ cores, and small scrapers mainly on quartzite. Goodwin (1929) reports differences in shape between the Layer 3 and Layer 5 handaxes, and suggest that shape in Layer 3 shows a trend toward the Fauresmith, but Keller (1973) does not support this assertion. There is no mention of Levallois or blade technology (Keller, 1973).

Acheulean artifacts were recovered at Elandsfontein, South Africa (also called Hopefield, or Saldanha), where an adult calvaria designated as Homo rhodesiensis or Homo heidelbergensis was recovered (Chapter 2). The handaxes are generally described as well-made and regular in form and were originally attributed to the Fauresmith or “Final Acheulean” (Klein, 1978; Klein,

59

1988; Singer and Wymer, 1968). Blades, points, and Levallois products were also recovered from surface localities where deflation probably resulted in a palimpsest of ESA and MSA artifacts (Singer and Wymer, 1968). At the Cutting 10 site, handaxes were excavated from beneath dune sand that sealed the ESA occurrence and separated it from the overlying MSA (Singer and Wymer, 1968). No blades, points, or Levallois products, were recovered from this sealed occurrence, which makes the assemblage heredifferent from the many Fauresmith- designated sites in central South Africa that contain LCTs, blades, points, and Levallois technology (Table 3).

There are four excavation areas at Kalambo Falls, Zambia, and in all four areas, Sangoan assemblages (characterized by core-axes and trihedral picks) are associated with ochreous sands underlying Lupemban (MSA) horizons (Clark, 2001; Sheppard and Kleindienst, 1996). These Sangoan assemblages overly Acheulean assemblages at two of the sites. The Kalambo Falls Sangoan assemblage is being discussed here because its Fauresmith affinities were originally highlighted (Clark, 1954), before it was designated as Sangoan. These Sangoan assemblages contain blades, points, and Levallois technology. There are no counts of blades, which are subsumed within the flake category. Clark (2001:246) states “while the Chipeta Industry continues Acheulean flake technology, an increasing number of flakes, though still small, are blade-like and long triangular and are indications of a trend toward more specialized flake forms”. Points are usually unifacial and rare compared to the frequency of LCTs that are larger and ‘crude’ compared to the underlying Acheulean handaxes. The morphology and frequency of these LCTs seems to distinguish the Kalambo Falls Sangoan from many Fauresmith-designated assemblages.

Covimvaba is an open-air locality in the Eastern Cape. Surface collections in 1925 recovered LCTs (“not well-made” but typical), blades, and retouched points on shale and hornfels (Goodwin and Van Riet Lowe, 1929:92-94). The assemblage housed at the Albany and South Africa Museums, and the site requires further investigation to be properly evaluated.

Another site requiring further investigation is Bushman’s , which is located in the Limpopo Province of South Africa. The rockshelter was excavated in 1965; the lowermost layers (28-43) yielded a MSA-designated assemblage with points, blades, as well as some pigments and beads (Louw, 1969). Most of the artifacts were manufacted on hornfels. Two handaxes were

60 recovered from layers 33 and 39, and interpreted as intrusive by Louw (1969). Herries (2011) cites this site as potential Fauresmith site, but without further study that statement seems unwarranted.

3.2.6 Fauresmith chronology

A relative position of Fauremith-designated assemblages between more typical ESA and MSA assemblages is well-established. Besides KP1, Fauresmith-designated assemblages appear to be stratified above Acheulean and beneath MSA deposits at two other excavated sites in the Northern Cape– Pniel 6 (Beaumont, 1990e) and Canteen Koppie (Beaumont and McNabb, 2000). At the excavated sites of Bundu Farm (Kiberd, 2006) and Rooidam 1 (Beaumont, 1990f; Butzer, 1974; Fock, 1968; Szabo and Butzer, 1979) Fauresmith assemblages are stratified beneath MSA deposits.

The multicomponent cave site of Wonderwerk Cave has the most potential for establishing a robust temporal record for the ESA-MSA transition in central South Africa. The stratigraphy across the seven excavation areas was linked by the concept of “Major Units” (Beaumont and Vogel, 2006). As identified by Beaumont and Vogel (2006), Major Units 3-4 (and maybe 5) contain the Fauresmith-designated assemblages, and underlay Major Unit 2, which contains MSA assemblages, and overlay Major Units 6-8, which contain Acheulean assemblages. However, there are serious issues with correlating the stratigraphy across the site in this way (Chazan and Horwitz, in press) and recent research at Wonderwerk Cave implies that the situation may be a bit more complex there than originally presented. As discussed above, reanalysis of the Wonderwerk assemblage has led Chazan et al. (Chazan, in press; Chazan et al., in press; Chazan and Horwitz, in press; Chazan et al., 2008) to question the identification of the Fauresmith in Excavation 1, where the Fauresmith assemblages are supposedly stratified between MSA and Acheulean assemblages. The Fauresmith assemblages from the middle of the cave (Excavation 2, 3, 4, and 7) have not yet been analyzed by the new research team (Chazan, personal communication). The assemblage from Excavation 6 identified as Fauresmith does show clear evidence of systematic blade production in association with bifaces (Chazan and Horwitz, 2009), and according to Beaumont and Vogel, this underlies an MSA assemblage, but Chazan and Horwitz (2009:527) suggest that the upper levels are also Fauresmith. It is far from

61 clear how the Fauresmith-designated deposits in Wonderwerk Cave relate stratigraphically to other archaeological units across the multiple excavation areas in the cave.

An absolute chronology for Fauresmith-designated assemblages is less secure than its relative position between the ESA and MSA. As discussed above, the Fauresmith-designated assemblages at Bundu Farm (Kiberd, 2006) and Rooidam 1 (Szabo and Butzer, 1979) provide OSL age estimates of 145.7 ± 16.0 ka and 174 ± 20 ka, respectively. Some of the deposits at Wonderwerk Cave identified as Fauresmith by Beaumont are less than 286±29 ka (Beaumont and Vogel, 2006). However, the dates at Bundu Farm and Rooidam 1 may be considered minimum age estimates for the levels that contain both LCTs and Levallois technology and blades (Herries, 2011:12,14), and Chazan et al. (Chazan, in press; Chazan et al., in press; Chazan and Horwitz, in press; Chazan et al., 2008) question the Fauresmith designation of some deposits at Wonderwerk Cave. The lithic assemblage from Excavation 2, which provides the U-series age estimate of <286±29, has not yet been reassessed. Therefore, it is impossible to establish a robust chronology for the so-called Fauresmith Industry based on current evidence. Nonetheless, relative and absolute dating techniques situate many of these Fauresmith-designated assemblages in the Middle Pleistocene.

3.2.7 Problems with the Fauresmith concept

The Fauresmith as an industry has never been thoroughly defined, and the term has been used inconsistently. Underhill (2011) provides a comprehensive overview of the history of the Fauresmith concept, highlighting the ad hoc manner in which tool types and technological strategies were added and or removed from the definition of the Fauresmith through time. For example, though the original definition focused on finely-made, small, almond-shaped handaxes made on flake blanks, later descriptions of Fauresmith-designated assemblages described cordiform, lanceolate, and double pointed handaxes (Table 1 in Underhill, 2011). Blades and points were not part of the original definition, though they were mentioned in early descriptions. With Beaumont's work, blades and points became part of the definition of the Fauresmith. Problems with Fauresmith concept were recognized early on and there were formal decisions to abandon the term, as established by the 1967 Pan-African Congress on Prehistory (Humphreys, 1970; Sampson, 1972). Nonetheless, the term ‘Fauresmith’ has persisted.

62

Part of the reason the term 'Fauresmith' has persisted despite its problems could be linked to the influence of J. Desmond Clark (Underhill, 2011), who used the term Fauresmith to highlight technological changes at the end of Acheulean in his general synthesis The Prehistory of Africa (Clark, 1970). Furthermore, before the development of radiometric dating techniques for the Middle Pleistocene, industrial designations were a necessary part of identifying a relative chronology for an archaeological site and solving chronological problems (Clark and Riel- Salvatore, 2005). By identifying an assemblage as Fauresmith, an archaeologist is communicating its chronologically relative position between the ESA and MSA based on the mixed character of its technological components. However, today there are robust methods for establishing absolute chronologies, and industrial designations are much less important for establishing age then they once were.

Based on the current state of knowledge, the term “Fauresmith” is likely to continue being used to describe ‘transitional’ assemblages, especially in the interior of southern Africa, that contain LCTs, blades, points, and prepared cores, because there are assemblages that do not fit well within the definition of the ESA or MSA. I acknowledge that the KP1 stratum 4a assemblage is labeled as Fauresmith, but my description and analysis of the assemblage is not necessarily representative of all Fauresmith assemblages, nor does it confirm the existence of the “Fauresmith Industry”. If usage continues, however, the term ‘Fauresmith’ must be divorced from the idea of ‘culture’ or ‘tradition’, until further research establishes a thorough definition based on artifact types, frequencies, and core reduction strategies at numerous localities, demonstrates that these localities are contemporaneous using radiometric dating techniques (as was done for the , Jacobs et al., 2008), and verifies that Fauresmith assemblages as a group are distinct from other “transitional” assemblages in Africa.

3.3 Blade Production in the Middle Pleistocene

3.3.1 Blades: definition

Here, blades are defined as detached lithic pieces that are at least as twice as long as they are wide along the axis of percussion. Some technical definitions also emphasize parallel or slightly convergent edges, with one or more dorsal ridges running parallel to them (Inizan et al., 1999:71). A distinction is sometimes made between ‘flake blades’ and ‘true blades’ in MSA assemblages (e.g. Beaumont and Vogel, 2006; Volman, 1984). Flake blades were distinguished

63 from true blades because they were not manufactured using the type of production that characterizes the Upper Paleolithic. For reasons discussed below, that terminology is not adopted here.

From a European perspective, blades have traditionally been associated with modern humans. The association of blades with Homo sapiens originally developed out of the contrasting nature of ubiquitous laminar technologies of the Upper Paleolithic compared to the mainly flake-based technologies of the Middle Paleolithic (de Mortillet, 1883:339, as cited in Conard, 1989:243). Blades were argued to provide an adaptive advantage in raw material efficiency and blade- dominated assemblages using prismatic reduction strategies became part of the support for the Upper Paleolithic Revolution model for modern human origins ~50 ka (Bar-Yosef, 2002; Klein, 1989:356). Based on well dated contexts, it is now established that even in Europe and the , early laminar technologies are not uniquely associated with the Upper Paleolithic or Homo sapiens (e.g. Delagnes and Meignen, 2006; Hershkovitz et al., 2011; Shimelmitz et al., 2011).

Furthermore, researchers now recognize that there are multiple reduction strategies that can result in laminar end products. Bar-Yosef and Kuhn (1999) emphasized three types of blade production: prismatic, Levallois, and Hummalian. Prismatic blade production generally starts with the creation of a crested blade to set up the initial guiding ridge for blade removals and the blade removals wrap around a core’s full or partial perimeter. Levallois blade production occurs on the upper surface of a bifacial core that is prepared to exhibit appropriate latitudinal and longitudinal convexities. The lower surface of the core serves as the platform. Hummalian blade production occurs on two or three adjacent surfaces of a minimally prepared core (Boëda, 1995). This tripartite scheme does not model all inter and intra-site diversity in blade production strategies, nor is there evidence for a simple linear trajectory from one strategy to another through time. For example, non-Levallois prismatic blade production characterizes the Upper Paleolithic, but is not exclusive to it (Bar-Yosef and Kuhn, 1999; Delagnes and Meignen, 2006; Marks and Monigal, 1995; Revillion and Tuffreau, 1994).

3.3.2 Blade production in the MSA

Blades have always been an important component of the definition of the African MSA. The relative abundance of blades was used to distinguish different subphases in the MSA of southern

64

Africa (Volman, 1984); MSA 1 was mainly flake-based, MSA 2 mainly blade-based, HP blade and bladelet-based and characterized by backed blades, and MSA 3 was mainly blade-based. Despite the importance of blades in the MSA record, there have been a limited number of technological analyses of MSA blade assemblages focused on identifying core reduction strategies. Exceptions include analyses of MSA 2, HP, and MSA 3 deposits at Klasies River Mouth (Villa et al., 2010; Wurz, 2000; Wurz, 2002), the Nubian Complex in North Africa (Van Peer, 1992), and the HP assemblage at Rose Cottage Cave (Soriano et al., 2007). For ‘transitional’ assemblages, a description of ~500 ka old blades and blades cores at Kapthurin Formation (Johnson and McBrearty, 2010) has been presented. Some observations and trends in MSA laminar technologies are highlighted below.

There are differences in the ways that blade cores can be organized, prepared, and exploited for blade production. Organization summarizes the manner in which the volume of the core was perceived and utilized for detaching pieces. Preparation refers to whether and how removals were taken off the exploitation surface in order to control convexities for the exploitation surface. Exploitation refers to the how the ‘predetermined’ removals were detached from the exploitation surface. Following Boëda (1995), exploitation can be preferential (core preparation occurs between each removal), or recurrent (one prepared surface yields multiple endproducts). Blade removals can also be unidirectional (off a single platform), or bidirectional (off two opposed platforms). Convergent exploitation uses a central guiding ridge for the detachment of a triangular endproduct.

In the later portion of the MSA at least, there was a diversity of blade production strategies (Table 4). At some sites, like Klasies River Mouth, there is evidence for temporal change in blade production strategies, but the dominant pattern for the MSA blade production variability is not lineal (e.g. there is no trajectory from Levallois to non-Levallois, or preferential to recurrent).

The MIS 6 and MIS 5 assemblages at Pinnacle Point 13B, Mossel Bay, also document some temporal changes in blade production, but the trends do not follow roughly contemporary trends at KRM, and there is as much spatial variability between the two excavated areas of the cave as there is between the temporal units (Thompson et al., 2010) . The methods of analysis differ for these assemblages compared to the other MSA assemblages discussed here, because they focus more on quantifiable characteristics of endproducts with less information about core reduction

65

Table 4 Summary of MSA blade core reduction strategies during MIS 5 and 4 (Late Pleistocene). MIS 5 MIS 4 (~130-74 ka) (~74-60 ka) KRM MSA 2a KRM MSA 2b Nubian 1 Nubian 2 RCC HP KRM HP KRM MSA 3 Taramsa 112 (Wurz, 2000; (Wurz, 2000; Method Method (Soriano et (Villa et al., (Villa et al., Refits 6 and 3 Wurz, 2002) Wurz, 2002) (elongated (elongated al., 2007) 2010) 2010) (Van Peer, points) points) 1992) (Van Peer, (Van Peer, 1992) 1992) Core pyramidal or two two two pyramidal two pyramidal two organization flat (non- hierarchical hierarchical hierarchical (non- hierarchical (non- hierarchical Levallois) surfaces surfaces surfaces Levallois) surfaces Levallois) surfaces (Levallois) (Levallois) (Levallois) (‘Levallois- like”) Preparation minimal indeterminate distal bilateral minimal centripetal crested blade centripetal/ preparation divergent removals preparation – removals bidirectional removals some crested blades Exploitation mainly recurrent preferential preferential recurrent recurrent recurrent bidirectional recurrent unidirectional convergent convergent unidirectional unidirectional unidirectional or unidirectional convergent or unidirectional bidirectional Technique direct soft direct hard direct hard direct hard direct soft direct soft direct hard direct hard ? hammer hammer hammer stone stone hammer hammer? hammer? hammer?

strategies. For that reason, PP13B blade production cannot be summarized in Table 4. Related traits such as platform types and the relative frequency of blades to flakes imply that the recorded variability reflects core reduction strategy, however, and the study highlights the reality thattemporal change is not the only explanation for lithic variability in the MSA (Thompson et al., 2010).

The earliest evidence for blade production comes from >500 ka deposits in the Kapthurin Formation, Kenya (Johnson and McBrearty, 2010). The Kapthurin blade cores are generally pyramidal or flat. Blades generally follow the long edge of one preparatory flake scar with little to no core preparation. Blades were extracted recurrently from unidirectional or centripetal cores. (Johnson and McBrearty, 2010:195) describe the reduction strategy as similar to the Hummal volumetric concept (c.f. Boëda, 1995).

Blades were manufactured using various reduction techniques in the MSA. Lithic reduction technique is defined as the manner in which flakes are detached from stone cores, employing direct or indirect percussion or pressure, and using hard or soft percussors (Inizan et al., 1999:30- 32). The earliest knapped stone tools ~2.6 million years ago were manufactured by striking stone with stone, a technique described as direct hard hammer percussion. Stone tools manufactured by modern human populations include a range of other techniques. Soft hammers made of bone,

66 , wood, and ivory have been argued to remove larger, thinner flakes (Bordaz, 1970). Indirect percussion is accomplished by using a bone, antler, or wood punch between the percussor and the stone core so that the force is directed more precisely (Whittaker, 1994). At Klasies River Mouth, MIS 5 assemblages attributed to the MSA 2a subphase (<115 ka) exhibit small, lipped platforms, diffuse bulbs of percussion, and extensive preparation, which Wurz (2000; 2002) suggests could indicate direct soft hammer percussion. In the subsequent MSA 2b subphase, platforms are thick and the bulbs are often splintered, consistent with direct hard hammer percussion (Wurz, 2000; Wurz, 2002).

The time period that has received the most attention with respect to lithic reduction techniques is the HP, and analyses provide conflicting results. Different researchers have concluded that HP blades were manufactured using direct hard hammer percussion (Winter, 2000), direct soft hammer percussion (Parkington et al., 2005; Wurz, 2002), indirect percussion (Deacon and Deacon, 1999; Singer and Wymer, 1982), and soft stone hammer percussion (Soriano et al., 2007; Villa et al., 2010). Many of these contradicting conclusions are based on independent analyses of the same site, Klasies River Mouth, reflecting the limits of current literature on identifying reduction techniques at MSA sites. However, evidence from the sites such as Diepkloof (Parkington et al., 2005) could potentially reflect variability related to geographic, raw material, or functional difference. More research is required to explain these various interpretations. Post-HP assemblages at Klasies River Mouth suggest that direct hard hammer percussion was used for blade detachment (Villa et al., 2010). The oldest securely dated blades from the Kapthurin Formation, Kenya (500-548 ka) were probably manufactured using direct hard hammer percussion based on prominent bulbs and bulbar scars, and the fragmentary nature of the recovered blades (Johnson and McBrearty, 2010:196). To summarize, the earliest blades in Africa were manufactured using direct hard hammer percussion, but by MIS 5 hominins used various techniques, and there is no lineal temporal trend.

3.3.3 The significance of blade production

Many considerations of the significance of blade production have focused on explaining its presence in Upper Paleolithic assemblages. Bar Yosef and Kuhan (1999) proposed that Upper Paleolithic blade production contrasts with earlier laminar technologies in its ubiquity across Western Eurasia after 45 ka (Bar-Yosef and Kuhn, 1999:330). In the Upper Paleolithic, the use

67 of soft-hammer and indirect percussion may result in finer, more standardized products suited for composite technology, and the widespread production of small blades (bladelets) during this time may be reflective of multi-component technology, which offers increased effectiveness, reliability, and maintainability. The costs associated with ‘front-loading’ efforts might have been offset with coordinated and cooperative social structures that were not in place prior to the Upper Paleolithic. One can call this the economic hypothesis for blade production.

An alternative to the economic hypothesis for Upper Paleolithic blade production is that blade production reflects technological choices related to stylistic or symbolic concerns (Eren et al., 2008). Despite the general consensus that blade production more efficiently exploits raw material than flake production, an experimental study reveals that blades from prismatic cores do not actually provide more cutting edge by volume of raw material than Levallois flakes do (Eren et al., 2008). Flakes have a longer use-life when resharpened because blades are rapidly exhausted due to their narrow form. Furthermore, investigations of archaeological collections do not support the hypothesis that Upper Paleolithic blade-based tools are more standardized in form than Middle Paleolithic flake-based tools (Chazan, 1995). If raw material efficiency and end product are ruled out as explanations for practicing blade production, despite the costs associated with it (the risk of production failures and thus, the wastage of raw material), then perhaps it is worth considering the social benefits of blade production within the context of hunter-gatherer group dynamics. For my purposes, I will identify this argument as the stylistic hypothesis. Eren et al. (2008) draw from literature on signaling theory, but researchers could also draw from the theory of style (Sackett, 1982, 1985; Wiessner, 1983; Wobst, 1977) or the anthropology of technology (Dobres, 2000) to understand the active role that lithic artifacts and technology play in constructing and reaffirming social relations, and cultural transmission (Boyd and Richerson, 1985; Boyd and Richerson, 1987; Boyd and Richerson, 2005; Richerson and Boyd, 2005) to understand how ideas about technology pass from generation to generation. Though Bar Yosef (1999) acknowledge a social component to their economic hypothesis, the stylistic hypothesis differs because it suggests that social factors explain the production of blades, rather than just offsetting the costs associated with laminar technologies and ‘front- loading’.

Systematic blade production is lacking in Oldowan and early Acheulean assemblages, so it still worth asking what the appearance of blade production technologies before the Upper Paleolithic

68 signifies in the evolution of the hominin lineage. Bar Yosef and Kuhn (1999) argued that prior to the Upper Paleolithic, blades came and went, that the advantages of blade production were situationally relevant, and that the choice to practice blade production was related to local economic reasons. However, there remains little research into the explaining the appearance of blade production in some early Upper Pleistocene and Middle Pleistocene assemblages. Furthermore, when considering the economic and stylistic hypotheses for blade production, there is no reason to assume that every early blade assemblage can be explained the same way or that the two hypotheses – economic vs stylistic – are exhaustive or mutually exclusive.

In a recent critique of the ‘modern human behavior’ concept, (Shea, 2011b) argues that behavioral variability, evidenced by the co-occurrence of the first four technological Modes including blade technology (Clark, 1977), is essentially identical to Upper Paleolithic humans by at least 200 ka in East Africa. He argues that hominins practiced a variety of lithic reduction strategies in response to human behavioral strategies related to energetic cost/benefit balances and situational variables. The early appearance of 500 ka blade production in the Kapthurin Formation has also been used to argue for diversification in hominin behavior (Johnson and McBrearty, 2010). New technologies may reflect innovative behaviors of some hominin groups and the incremental addition of novel behaviors to MSA technological repertoires (Tryon, 2006; Tryon et al., 2005). Foley and Lahr (Foley and Lahr, 2003; Foley and Lahr, 1997) suggested that the shift to MSA technology marked an important move toward technological and behavioral diversity. Levallois endproducts served as blanks, from which any number of tool forms could be fashioned. In the MSA, these blanks were often fashioned into projectile or spear points that exhibited spatially and temporally limited distributions. There is general consensus that the spatial and temporal pattern of MSA point styles implies that artifact variability in the MSA is associated with cultural patterns and “regional traditions” (Clark, 1982; Foley and Lahr, 2003; Foley and Lahr, 1997; McBrearty and Brooks, 2000), though there has been little effort to identify how environmental variables may have influenced extractive behaviors, point function and form. The concept of “diversity”, as it stands as an explanatory framework for technological behaviors in the late ESA and MSA, is linked to both the ‘economic’ and the ‘stylistic’ hypotheses for the origins of blade technology, because spatial and temporal diversity can reflect adaptation to different physical and/or social conditions.

69

3.4 MSA Points and Point Function

3.4.1 Points: definition

Points are triangular lithic products that may be either retouched or not, and are one of the defining characteristics of MSA assemblages (Goodwin and Van Riet Lowe, 1929). During the MSA, flakes and blades were often shaped into a pointed form using unifacial or bifacial retouch. For example, in the post-Howiesons Poort levels at Sibudu Cave, retouched points are mainly unifacially worked; both lateral edges are retouched on the dorsal side so that the distal end converges into a point (Villa et al., 2005). More invasive bifacial retouch is characteristic of the Lupemban lanceolate points known from Central Africa (see Taylor, 2011) and the Still Bay points from southern Africa (see Mourre et al., 2010; Villa et al., 2009b).

Levallois methods of reduction were also used during the MSA to manufacture triangular flakes of a ‘predetermined’ form that were not modified by retouch. The Nubian Industry is an example of this type of technology (Van Peer, 1992). In Southern Africa, the Mossel Bay Industry, or MSA II substage, is also characterized by the production of convergent triangular flakes using Levallois reduction methods (Wurz, 2002).

3.4.2 Functional studies of MSA points

It is frequently suggested that MSA points were hafted and used as hunting weapons, and there are functional studies that support this view. The most direct evidence for a lithic tipped armatures used for hunting may come from MIS 5 deposits (~130-74 ka) at Klasies River Mouth, South Africa, where a point tip embedded in a giant buffalo ( antiquus) vertebra was recovered (Milo, 1998).

Surface wear and residue deposits on MSA points provide indirect evidence for MSA hunting technologies (Table 5). Retouched points from the Still Bay levels (72.7±3.1, Jacobs et al., 2006) and non-retouched convergent flakes from the underlying levels (98.9±4.5 ka, Jacobs et al., 2006) at have impact damage consistent with use as hunting implements (Lombard, 2007). At White Paintings Rockshelter, Botswana, five retouched points from sediments dated to between 66.4±6.5 and 94.3±9.4 ka exhibit impact damage and are interpreted as spear tips (Donahue et al., 2004). At Sodmein Cave, Egypt use wear on tool surfaces is

70

Table 5 Types of evidence used to determine whether MSA points were used as hunting implements Type of evidence Characteristic of hunting implement Microscopic use wear Longitudinal striations parallel to point margins (Donahue et al., 2004; Lombard, 2005a; Rots et al., 2011) Linear polish parallel to point margins (Rots et al., 2011) Hafting polish/bright spots at proximal end (Lombard, 2005a; Rots et al., 2011; Van Peer et al., 2008) Binding scars on lateral edges near proximal end (Van Peer et al., 2008) Macroscopic use wear Impact fractures (Brooks et al., 2006; Donahue et al., 2004; Lombard, 2005a, 2007; Lombard et al., 2004; Rots et al., 2011; Villa and Lenoir, 2006) Edge Damage Distribution Increased frequency of damage at tip and in hafting zone, equal frequency of damage on left and right laterals (Bird et al., 2007; Schoville, 2010; Schoville and Brown, 2010) Residue Animal tissue resides at distal end (Lombard, 2004a, 2005a) Plant and ochre residues at proximal end (Lombard, 2004a, 2005a) Retouch and shaping Basal modification for hafting (Brooks et al., 2006; Villa et al., 2005; Villa and Lenoir, 2006) Geometric morphometrics Symmetry through all stages of reduction/resharpening (Iovita, 2011)

consistent with hafting and impact (Rots et al., 2011). The site yields Nubian Complex type tools, which date to the Last (127-110 ka) at other North African sites. Six out of ten selected pieces were interpreted as hunting weapons based on macrowear and microwear characteristics (Rots et al., 2011:645). The Lupemban levels at Sai Island, Sudan (recovered from the Black Silts unit, similar to MIS 5 assemblages in North Africa) have two lithic pieces with impact fractures that are interpreted as 'projectiles' (Rots et al., 2011:642). A lanceolate point presumably from the early Upper Pleistocene at Taramasa-8 in North Egypt shows evidence for hafting, but broke and was discarded before use (Van Peer et al., 2008). Retouched points from ≠Gi, Botswana (~77 ka) and Aduma, Middle Awash, Ethiopia (<70 ka) exhibit basal modification and impact fractures consistent with use as hunting implements (Brooks et al., 2006).

Not all MIS 5 MSA points were used as hunting weapons. Functional analyses of convergent flakes dated to between 162 and 90 ka (MIS 6 and 5c-e) from the site of Pinnacle Point 13B suggest that points there were used as cutting tools rather than hunting weapons (Bird et al., 2007; Schoville, 2010). This conclusion is based on the distribution of damage along point edges. Rather than interpreting tool function for individual points, this approach looks at the distribution of damage at the assemblage level (n=31 for Bird et al., 2007, and n=238 for

71

Schoville, 2010). Damage on PP13B points is not concentrated at point tips as expected for hunting implements based on experimental research (Schoville and Brown, 2010). Instead, edge damage is distributed along point edges with decreasing frequencies at point tips and increased frequency on the left side compared to the right. The PP13B study expands the variation in point function and implies that points discarded at occupation/cave sites may serve different or a greater variety of functions than points discarded at kill/open-air sites; it does not rule out the likely possibility that points were also used as hunting implements in other contexts (Schoville, 2010).

Kuman (1989:241, 286) suggests that the dominant function of MSA points at ≠Gi may have been cutting based on point shape (rounded tips), retouch characteristics (steeper retouch on one lateral edges), and distribution of use wear (extending from tip to base).

Tanged tools in Northern Africa, often interpreted as the hafted tip of hunting implements, are retouched in a manner more consistent with cutting or scraping tools than armatures (Iovita, 2011). The smaller tools, which have undergone multiple episodes of retouching/resharpening, are more assymetrical than the large tools. If these tools were hafted as armatures, we would expect a symmetrical point to be maintained through the life history of the tool. Points from the Final Paleolithic of North Germany with a known function as weapon tips exhibit less asymmetry than the Aterian tools (Iovita, 2011).

Sites from more recent MIS 4 (~74-60 ka) indicate that at least some points were used as hunting implements. At Sai Island, Sudan, five bifacial foliates from deposits <60 ka were apparently used as hafted projectiles or hafted /axes, and one Levallois flake shows evidence that it was used as a hafted projectile (Rots et al., 2011:652). At Sibudu Cave, retouched points from deposits dated by OSL to 50-64 ka exhibit diagnostic impact fractures, microwear, and plant and animal residue distributions consistent with use as hafted hunting weapons (Lombard, 2004b, 2005a). A quarter of the points from the post-Howieson’s Poort levels have modified bases, which is an additional observation consistent with hafting (Villa et al., 2005).

There has been only one functional analysis of points confidently predating MIS 6 (>130 ka). The oldest securely dated MSA points that have undergone a functional analysis come from the Sangoan levels at Sai Island, Sudan. One point fragment from the Middle Sangoan unit (202-162 ka) exhibits evidence for hafting and impact damage and is argued to be a projectile fragment

72

(Rots et al., 2011:642). One foliate from the underlying Lower Sangoan exhibits no use wear (Rots et al., 2011:642). As it stands, the evidence for hafted hunting weapons prior to MIS 5 is scant.

3.4.3 Spears vs. projectiles

The term ‘projectile’ is often restricted to technologies that significantly increase the distance between the hunter and the prey, i.e., arrows and darts launched with a spear thrower, whereas thrusting and thrown spears are not considered projectiles (Shea 2006). Alternatively, the term “complex projectile” refers to bow/ and spearthrower/ composite technologies that propel at high velocity and “simple projectile” are those relying solely on human mechanical energy like hand-cast spears, javelins, and throwing sticks (Sisk and Shea, 2011). However, not all researchers recognize the same terminological distinctions or explicitly state what they mean when they used the term ‘projectile’. The distinction has evolutionary importance, however, because complex projectiles and spears represent different prey acquisition strategies with implications for adaptive fitness. Ethnographic evidence suggests that the effectiveness of spears and projectiles is dependent on the environment and targeted prey (Churchill, 1993), and adaptive technologies can reduce the risk of injury, while increasing the probability of a successful hunt.

Methods for distinguishing projectile tips from spear tips are still under development. Experimental research demonstrates that similar types of damage related to impact and hafting occur on both projectile tips and spear tips (eg. Fischer et al., 1984; Lombard et al., 2004). Much of the recent MSA research related to hunting technology has focused on the backed pieces characteristics of the HP subphase. Use wear and residue studies provide evidence that backed pieces were hafted and used as hunting weapons (Lombard, 2005a, b, 2006, 2008, 2011; Lombard and Pargeter, 2008; Lombard and Phillipson, 2010). The small size, light weight, and standardized form of quartz backed pieces from Sibudu Cave are used to argue that they were used as hafted arrowheads (Wadley and Mohapi, 2008). A small bone point may provide additional support for technology during this time (Backwell et al., 2008). However, Villa et al. (2010) are skeptical of this interpretation, arguing that backed pieces could alternatively represent hafted spear or spear thrower dart elements. More recently, Lombard (2011) argues that because small backed pieces were hafted transversely, at least some backed

73 pieces were used as arrowheads. If this hypothesis is accepted, then the HP marks the first appearance of bow and arrow technologies ~60 ka.

Concerning MSA points, there are few studies that support their use as projectiles. For the functional analysis of a few selected lithics from Sodmein Cave, projectile elements were distinguished from thrusting spear elements based on ‘twisting’ evidence in the latter (Rots et al., 2011:646). However, this method for distinguishing projectiles from spear tips has yet to be demonstrated experimentally. Rots et al. (2011:642) argue that projectile technology dates to at least the Sangoan (>200 ka) occupation of North Africa based on their analysis of a single lithic fragment from Sai Island, Sudan, but their method for distinguishing projectiles from spears so far has little basis.

MSA researchers more often use size to distinguish points used as spear tips from points used as projectile tips. Ethnographic and experimental evidence demonstrates that for projectile tips to function as such, they need to be small and tend to be smaller than spear tips (Shea, 2006, and references therein). At Sibudu, post-Howiesons’ Poort (~53 ka) point size, based on maximum width measurements, is more consistent with use as spears than arrowheads (Villa et al. 2005). Tip cross-sectional area (TCSA) is a variable related to size that can be an effective discriminator (Hughes, 1998; Shea, 2006). The Sibudu Cave post-Howiesons’ Poort points have a mean TCSA of greater than 100 mm2, which is inconsistent with arrow or dart technologies (Villa et al. 2005). Shea (2006) presents point metrics from Klasies River Mouth, Blombos Cave, Porc Epic, North Africa (tanged Aterian points), and Still Bay points from various sites. TCSA values for these assemblages are significantly higher than those of arrowheads and dart tips, but significantly lower than experimental thrusting spears, which Shea (2006) interprets as evidence that if the points were used as hafted hunting weapons, they were used as tips for throwing or thrusting spears instead of projectiles. Tip cross-sectional perimeter (TCSP, Sisk and Shea, 2009) is perhaps a superior size-related measure that can be used to discriminate projectile from spear technologies. A TCSP analysis of the same assemblages examined by Shea (2006) suggests that points from Porc Epic, Ethiopia and some Aterian assemblages are consistent with TCSP values for dart tips (Sisk and Shea, 2011). However, morphometric data only provides information on functional potential, not actual use, and using the mean values for an entire assemblage of points to determine functional potential will obscure differences; there could be more than one weapon type present in an archaeological assemblage (i.e., projectiles and non-projectiles) Furthermore,

74 new research on the tanged Aterian tools indicates that cutting is the more probable function for the majority (but not necessarily all) of these pieces (Iovita, 2011).

Brooks et al. (2006) argue that because some points at ≠Gi, Botswana (~77 ka) and Aduma, Middle Awash, Ethiopia (<70 ka) are below the size range for spears based on length, width, thickness, and mass measurements, that at least some MSA points may have functioned as projectile tips. Brooks et al. (2006:251) suggest some may have been used to tip spear thrower darts, even though there is no ethnographic evidence for spear thrower darts in Africa. At Rose Cottage Cave, South Africa, it is argued that large points from post-Howiesons Poort layers (~50-29 ka) were more likely used as spearheads, but small points from one of the final MSA layers (~31-29 ka) may have been used as arrowheads based on the small mean TCSA value (Mohapi, 2007; Villa and Lenoir, 2006).

Velocity-dependent features of macrofractures such as Wallner lines and fracture wings (Hutchings, 2011) provide the most promising avenue for identifying projectile technologies on very fine-grained materials, but this method has also not yet been applied to the MSA record.

To summarize, metric data of MSA points are generally more consistent with throwing or thrusting spear technology, than projectile technology. HP backed pieces and bone points may provide some evidence for projectile technologies by 60-65 ka. In later time periods, based on the final MSA at Rose Cottage Cave, some small points could have been used as arrowheads. If the small size of some points at ≠Gi indicates that they were used as projectile elements (Brooks et al. 2006), then one might push back the origins of projectile technology to ~77 ka. Even if the origins of projectiles can be traced this far back, spears either continued to be used or were re- adopted in some contexts, based on evidence from the post-Howiesons’ Poort levels of Sibudu.

3.4.4 Problems with current evidence

The evolution of hunting technologies in Africa is far from resolved. The question of backed pieces and their role in bow and arrow technology aside, there have been few adequate functional studies of MSA lithics. Potential spear armatures have received much less attention than potential projectile armatures, even though the origins of technology-aided hunting of large game is evolutionarily significant. The function of points in the Middle Stone Age is generally assumed. However, few studies adequately identify whether MSA points were (1) hafted, (2)

75 used to process animal remains, and (3) subjected to impact. All these criteria must hold if they were used as hunting weapons, but alone do not indicate what kind of hunting technology was utilized. Some researchers argue MSA points were spear tips based on mean TCSA values (eg. Shea 2006), others argue that the smallest MSA points are indicative of projectile technology (Brooks et al., 2006; Mohapi, 2007; Sisk and Shea, 2011).

Equifinality is serious concern in identifying MSA point function. The three criteria above would also hold true for hafted daggers and carving tools used on bone (Shea 1991). Even the most direct evidence for MSA hunting weapons, the lithic fragment embedded in the giant buffalo vertebra at Klasies River Mouth (Milo, 1998), could potentially represent a butchery tool because the penetration angle towards the front and below of the animal is unlikely for hunting (Marean and Assefa, 1999). There has been little work regarding wear patterns on butchery tools. Based on evidence from PP13B, there is good reason to believe that some points were used as cutting/butchery tools rather than hunting weapons (Bird et al., 2007; Schoville, 2010; Schoville and Brown, 2010). The same may also be true at ≠Gi (Kuman, 1989). Studies indicating a cutting function for points show that it is invalid to assume function based on morphology alone. Recent research is also identifying the role post-depositional processes play in creating edge damage, especially scars at the distal end interpreted as impact damage (Pargeter, 2011b). Trampling and other non-behavioral processes can result in so-called “diagnostic impact fractures”. For these reasons, functional studies of points need to consider multiple lines of evidence and be able to rule out alternative hypotheses for observed wear characteristics and patterns.

3.5 Raw Material Foraging in the ESA and MSA

Raw material form and mechanical properties impose technological constraints on artifact production and use (e.g. Andrefsky, 1994; Clark, 1980; Luedtke, 1992), and lithic technology involves making selective decisions about raw material choice. Hominins at some Oldowan sites ~2 Ma in West Turkana selected specific raw materials for the implementation of certain techniques because of the internal structure of these rocks (Braun et al., 2008; Braun et al., 2009; Goldman-Neuman and Hovers, 2012; Stout et al., 2005). There is also clear evidence for raw material selection and differential use at African Acheulean sites. Some Acheulean assemblages are characterized by the use of durable coarse-grained rocks for the production of LCTs (Jones, 1979, 1994; Sharon, 2008). Raw material use for smaller 'light-duty' tools seems to be more

76 variable during the Acheulean, but with a preferential use of finer-grained rocks or quartz visible at some sites like Ismilia, Tanzania (Howell et al., 1962) and Kalambo Falls, Zambia, (Clark, 2001; Sheppard and Kleindienst, 1996). Raw material foraging behaviors in the MSA are generally considered more selective than earlier periods, with more time expended on search, extraction, and processing. An illustrative example of this is provided by early Late Pleistocene quarry sites in Egypt where MSA hominins were digging open pits nearly 2 m deep and across huge areas of ~1000 m2 to extract chert (Vermeersch et al., 1990).

Raw materials play a significant role in understanding and explaining lithic assemblage variability, forager mobility, landscape use, and interaction in the MSA. The HP subphase of the MSA, for example, has resulted in multiple hypotheses regarding the increased reliance on fine- grained raw materials that characterizes this period. One of these hypotheses is that the increased reliance on fine-grained raw materials in the HP reflects larger foraging territories and increased mobility (Ambrose and Lorenz, 1990). However, surveys on the southern coast of South Africa near PP5-6 and Sibudu suggest that fine-grained raw materials are in fact, local (Brown, 2011; Minichillo, 2006; Wadley and Mohapi, 2008). Here, local raw materials are defined as those that can be recovered from distances within a typical one day roundtrip, which averages around 15- 30 km (Binford, 2001:238; Kelly, 1995:133 ). Half of this average roundtrip, 8-15 km, represents the daily foraging radius of a hunter-gatherer community. Instead of representing larger territories or increased mobility, (Minichillo, 2006) argues that increased frequencies of fine- grained raw material at HP sites represent increased search time at cobble beaches adjacent to HP cave sites. Deacon (1989) suggests that the high frequencies of fine-grained raw material in the HP resulted from the exchange of tools, as in modern San ethnography (e.g. Wiessner, 1983), and the desire to add both functional and symbolic value to weapons in the context of exchange. Exchange, in the context of this model, served to stimulate the expansion of social networks and interaction that would minimize the risk associated with deteriorating climatic conditions. Brown and Marean (2010) recently proposed a connection between heat-treatment of silcrete, the HP, and the availability of firewood for heat treatment. Silcrete was heat-treated at Pinnacle Point to improve its flaking quality (Brown, 2011; Brown et al., 2009). This procedure requires a long processing time and large amounts of wood fuel, the availability of which would have peaked during periods when summer rains were dominant in the South African Cape. The appearance

77 and disappearance of the HP seems to correlate with climatic events that would have influenced the availability of wood fuel.

3.5.1 Distance of raw material transfers

The movement of raw materials across the landscape has important implications for the ranging patterns and resource networks of stone-tool using populations. Paleolithic data on raw material transfers indicate that 100 km is an important threshold for raw material movement. Lower and Middle Paleolithic raw material transfers rarely exceed 100 km, whereas Upper Paleolithic transfers are sometimes in excess of 200 km (Gamble, 1999). Raw material transfers in the Early and Middle Paleolithic are thought to be associated with regular seasonal trips during which raw material resources were collected in addition to subsistence items (Feblot-Augustins, 1993), whereas raw material transfers in the Upper Paleolithic are indicative of long distance trade (Gamble, 1999).

There is evidence from some parts of Africa for raw material transfers that exceed 100 km, and even extending beyond 300 km, during the MSA. In the central Kenyan Rift Valley, a dozen distinct petrological groups among the flows were identified using electron microprobe and x-ray florescence (Merrick et al., 1994; Merrick and Brown, 1984). These sources have been traced to a number of sites in Kenya and Tanzania, over 300 km distant. Some lithics from Muguruk and Songhor, Kenya are from sources 145-190 km away. Lithics from Nasera Rock Shelter, Tanzania have been sourced to an obsidian flow 240 km away and lithics at Mumba Rock Shelter are derived from a source 305 km away (Merrick et al., 1994). Raw material provenience studies from other parts of Africa also demonstrate long-distance movement of obsidian. The site of Porc-Epic Cave, Ethiopia yields lithic material exhibiting a distance of 250 km from site to source (Negash and Shackley, 2006). Transfers in excess of 100 km during the MSA probably indicate interaction and trade between groups, rather than the movement of individuals. The dates for the above sites are not well-established, but there is no evidence to suggest that they are more than150 ka.

In contrast to the MSA, Acheulean sites do not show evidence for raw material transportation beyond 100 km; most raw material is transferred from less than 10 km away (Merrick et al., 1994), consistent with the daily foraging radius of hunter-gatherers. At Gadeb, Ethiopia, obsidian handaxes may have been transported ~100 km (Clark and Kurashina, 1979). Four obsidian

78

LCTs, were recovered from this locality and the nearest primary source of obsidian is 100 km away. The handaxes seem to have been curated and abandoned as finished tools at the site. There is no mention of potential secondary sources (i.e., streams and river beds that may have transported the obsidian closer to Gadeb), however, so it remains possible that the source of the obsidian was actually closer. In Zimbabwe, at the site of Lochard, silcretes were used to manufacture Acheulean tools and the nearest outcrop is 65 km away (Bond, 1948). Again, there is no mention however, if there may have been closer secondary sources.

Little is known about distance of transfer for the early MSA or ‘transitional’ assemblages. Beaumont and McNabb (2000) mention that some blades in the Fauresmith assemblage at Canteen Koppie were manufactured on banded ironstone, which is only available 200 km away. This scale of long-distant transport is possible but needs to be verified.

3.5.2 Increased fine-grained raw material in the early MSA

For the early MSA and ‘transitional’ assemblages, there is evidence for an increased reliance on fine-grained raw material compared to Acheulean assemblages at some sites. In the Kapthurin Formation, a single raw material, phonolitic lava, is used for stone tool manufacture at the Acheulean sites of the Leakey Handaxe Area and the Factory Site, In contrast, a range of fine- grained lavas are used at the early MSA (i.e., ‘transitional’) site of Koimilot (Tryon et al., 2005). Tryon et al. (2005) suggest that the shift to fine-grained raw materials in the MSA could be related to suitability of these materials for Levallois reduction, or a functional emphasis on tools with sharper but potentially less durable edges in the MSA. At Kalambo Falls, there are increases in the frequencies of fine-grained raw materials in the Sangoan horizons compared to the ESA horizons (Sheppard and Kleindienst, 1996). At Kudu Koppie, there is also increased use of locally available fine-grained raw materials like chert, chalcedony, and banded ironstone when comparing the MSA deposit to the Late ESA (i.e. 'transitional’) assemblage with LCTs and Levallois technology (Wilkins et al., 2010).

In contrast to the above studies, at Casablanca (North Africa) the same raw material is used throughout the sequence that spans the Acheulean and a 'transitional' assemblage with LCTs and Levallois technology (Raynal et al., 2001). All levels exhibit similar frequencies of the locally- available abundant quartzites and rare (Raynal et al., 2001:73).

79

3.5.3 The Fauresmith and raw materials

Humphreys (1970) argued that the Fauresmith is not an entity distinct from the Acheulean, but appears to be distinct because of raw material factors. The potential influence of raw material on Fauresmith handaxe form and the production of larger flakes was also recognized by other workers and early on (Van Riet Lowe, 1927). Humphreys (1970) argued that most Fauresmith sites near Kimberley were located away from the Vaal River Valley and associated with Ecca Beaufort geology, where hornfels outcrops and is abundant on the landscape. The Acheulean sites are located mainly in the Vaal River Valley, where Ventersdorp Lava (andesite) and dolerite is available. Humphreys (1970) argued that raw material either determined the nature of or produced an 'advanced' appearance for the Fauresmith assemblages. Hornfels is considered easier to knap and it was argued that it may lend itself to improvements in knapping technique. Humphreys also cites evidence from the stratified Acheulean/Fauresmith/MSA site at Sheppard Island (Goodwin and Van Riet Lowe, 1929:235-243) where both Ventersdorp and Ecca deposits are found. The Acheulean at this site was manufactured on gravel materials (quartzite, ‘amygdaloidal lavas’, and dolerite), whereas the Fauresmith was manufactured almost exclusively on hornfels.

It is important to identify the raw materials utilized at Kathu Pan to address the question of the raw material and the Fauresmith. If the Fauresmith assemblage was manufactured on different raw materials than the underlying Acheulean, then one could hypothesize that raw material explains the difference between the Fauresmith and Acheulean technology (sensu Humphreys 1970). If there is little difference between the raw materials, than raw material does not explain the difference. Furthermore, identifying raw material foraging strategies at Kathu Pan 1 will provide answers to the following questions: How far are raw materials transferred during the Fauresmith occupation of KP1? How does this compare to the underlying Acheulean? Is there an increase in higher-quality raw materials compared to the underlying Acheulean? In other words, is the KP1 stratum 4a lithic assemblage consistent with a typical ESA raw material foraging strategy focused mainly on abundant, local raw materials, or with a characteristic MSA strategy focused on finer-grained raw materials that are transported longer distances and/or require increased search time?

80

3.6 Summary

In this chapter I provided the necessary background required to evaluate the significance of the KP1 stratum 4a lithic assemblage. Thisassemblage is one of very few dated to the early Middle Pleistocene. Other sites dated to this period document the co-occurrence of LCT technology typical of preceding periods, as well as the addition of blade technology, points, and Levallois core reduction strategies. Even though it has become commonplace to situate the transition from the ESA to MSA at ~300 ka, a review of the current chronological evidence indicates that securely dated assemblages between ~500 and 300 ka contain tool types diagnostic of the MSA. LCTs persist for a long time in Africa after the appearance of these technologies, until at least ~130 ka, but there are actually no known assemblages undoubtedly younger than ~500 ka that contain only LCTs in the absence of what are considered typical MSA technologies. KP1 is among just a handful of sites, including the Kapthurin sites, and Olorgesailie, Kenya that are dated to between ~500 and 300 ka, and as such, documents the important technological changes reflected in the early Middle Pleistocene archaeological record.

The stratum 4a assemblage at KP1 is designated to the Fauresmith Industry. The concept of the Fauresmith is problematic and I highlighted some of these issues. I also presented a list of assemblages that have been attributed to the Fauresmith, and identified some similarities between them. However, thorough analyses have not been conducted on the majority of these assemblages. Defining the Fauresmith and/or establishing its existence is not one of the goals of this analysis. To do so would require a cross-site comparison of lithic assemblages with good chronological control. My position is that the term 'Fauresmith' has become a convenient short hand for what I identify as 'transitional' assemblages in southern Africa, those that contain LCTs, and blade, points, and/or Levallois, but there is no current evidence to support the idea that it reflects a 'cultural' phenomenon. That the stratum 4a assemblage at KP1 is designated as Fauresmith is a moot point when considering most aspects of hominin behavioral evolution, because we know that the assemblage dates to the early Middle Pleistocene regardless of its ‘cultural’ or industrial designation.

It is now established that blade production has great antiquity, with early occurrences dated to the early Middle Pleistocene. There are diverse strategies for producing blades, and the MSA documents several permutations of the various methods used for core organization, preparation,

81 exploitation, and percussor type. Blade production is often considered as an efficient means of using raw material and/or producing standardized endproducts, but experimental and archaeological studies do not always support these assertions. Alternative interpretations focus on the role that core reduction plays in social contexts, because core reduction is probably learned and performed in a social environment. Diverse core reductions strategies that include blade production are indicative of new degrees of behavioral variability, which is arguable one of the most important characteristics of modern human adaptation. KP1 provides a rare opportunity to consider the degree and nature of behavioral variability with respect to core reduction strategies in the early Middle Pleistocene.

In general, MSA points as a tool category are considered tips on the ends of hafted hunting weapons. There are, however, some studies that highlight the role that points play in cutting or scraping tasks as well. Morphology alone should not be used to assume function. Furthermore, there have been no functional studies of points dated to more than ~250 ka, so it has not been established whether hafted hunting technologies are associated with the onset of the MSA.

During the ESA, lithic raw material was foraged from nearby sources, with less effort expended on search and extraction time than the MSA. During the MSA, particularly in the Later Pleistocene, raw material was sometimes transferred long distances, more than 100 km, and hominins often intensified their search and extraction time to focus on rarer, fine-grained raw material types. Longer transport distances could in part be associated with increased mobility, but distances of greater than 100 km probably represented trade and interaction between groups. A focus on fine-grained raw material could in part reflect functional requirements, but some researchers also highlight the role that high-quality raw material types can play in symbolic and social interactions. While the contrast in raw material foraging strategies is clear when comparing later MSA assemblages (mainly from MIS 5 and 4) to ESA assemblages, it is not clear how earlier MSA or ‘transitional’ assemblages compare to the ESA in this regard.

82

4 Kathu Pan 1 (KP1) 4.1 Geological and ecological setting

KP1is located about 4.5 km northwest of the town of Kathu, in the Northern Cape Province, South Africa (27º39’59 S, 23º00’26 E, Figure 3). This part of the Northern Cape is located on South Africa’s inland plateau, which consists of expansive , and geologically and ecologically contrasts with the Kalahari Basin to the north and the Great Escarpment to west, south, and east. There are two prominent hill ranges on this part of the inland plateau that run in a north south direction and about 400 m high; the Langeberg and the Kuruman Hills. The inland plateau falls within the summer rainfall region, with mean annual rainfall between 200-400 mm, which is intermediate compared to the drier regions to the west and the wetter regions to the east and along the southern coast (Barnard et al., 1972, as cited by Humphreys and Thackeray, 1983). With some exceptions like the Vaal and Orange, rivers are generally episodic. Changes in temperature are great; in the hottest month, January, temperatures can rise to mid-30ºC and in the coldest month, July, temperatures can drop to below 0ºC (Humphreys and Thackeray, 1983).

Pans and springs are major sources of water in the Northern Cape. Pans are depressions in the landscape that are often dry, but temporarily fill with water, and are dotted across the Northern Cape landscape. During the rainy season they fill up and may carry water for long periods of time. Springs are geological features where water flows to the surface from underground, and are often related to karstic systems. These features are abundant around the Kuruman Hills and important determinants of human land use and settlement patterns.

Intensive farming operations have changed the local vegetation and faunal distribution patterns. Reconstructions of vegetation patterns for A.D. 1400 suggest that the inland plateau once supported Bushveld vegetation composed mainly of a mixed bushveld with a dense variety of shrubs and trees (Humphreys and Thackeray, 1983:22). River valleys and other types of water sources probably encouraged the growth of Acacia karoo trees. The Northern Cape once supported a rich faunal biomass that included multiple species of , hippo, rhino, guagga (extinct ) and (Humphreys and Thackeray, 1983:23). Humphreys and

83

Figure 3 Map showing location of Kathu Pan and KP1. a. Relative to towns of Kathu and Kuruman, other archaeological sites in the region, and major geological features.Green stippling marks the location of volcanic raw material outcrops. b. aerial view of Kathu Pan. c. Outline sketch of pan and relative location of sites after Beaumont (1990b)

84

Thackeray (1983: 26-29) emphasize the ‘uniform’ nature of the Northern Cape environment on the inland plateau, suggesting that there is no marked resource zonation. The one exception is probably water, which is more abundant and accessible to the east of the Langeberg, than west of the Langeberg, based on Khoisan and Afrikaans linguistic patterns and historical land use patterns (Humphreys and Thackeray, 1983).

The site of KP1 is located in Kathu Pan, which is situated between the only two prominent hill ranges on the inland plateau. The Langeberg rise ~30 km to the west and the Kuruman Hills ~ 7km to the east. The Gamagara River is located about 11 km west of Kathu Pan, though minor tributaries run within 4 km of the pan today and the pan itself is part of the Gamagara drainage system. The Gamagara drainage systems runs northward until it meets the Kuruman River, which is a tributary of the Molopo River that forms part of the border between South Africa and Botswana. These rivers are dry except after flash floods. About 7 km south is Kolomela (previously Sishen) Mine that is exploiting the extensive iron and manganese deposits located there. The pan is ~0.3 km2 in area and is perennially flooded by high water table levels and artesian seepage (i.e., springs). The pan represents one of few permanent sources of water for the three farms whose boundaries run together there. Sinkholes into the underlying karstic system have formed and are still forming in and around the pan, and in some of these sinkholes, including KP1, stratified Stone Age deposits have accumulated. Today, the pan is covered with dwarf grass species and bordered by a thicket of Acacia karoo trees.

4.2 Previous Research at KP1

4.2.1 Excavations by Peter Beaumont

KP1 was discovered in 1974 when handaxes and faunal remains were observed eroding out of the wall of a newly-formed sinkhole about 3 m deep and 6 m wide. KP1 is one of eleven locations within Kathu Pan that were excavated between 1978 and 1990 by Peter Beaumont (Beaumont, 1990b). KP1 was excavated first, and then investigations spread to other parts of the pan. Table 6 summarizes the finds at each of the 11 Kathu Pan localities and Figure 3c shows their relative locations. Only 9 of the Kathu Pan excavations yielded archaeological remains. Several of the Kathu Pan localities, including KP1, are in-filled sinkholes or dolines that formed within calcretes of the -aged Kalahari Group. KP1 preserves the longest lithostratigraphic and archaeological sequence of the sites with Acheulean, Fauresmith, MSA,

85

Table 6. Summary of Kathu Pan sites based on Beaumont (1990b). Site Description KP1 Multicomponent doline site, Acheulean, Fauresmith, MSA, LSA, focus of this study KP2 Scant LSA artifacts, a few faunal remains, pollen data reveals Holocene paleoenvironmental shifts (Butzer, 1984) KP3 No archaeological remains KP4 No archaeological remains KP5 Multicomponent doline site, MSA, LSA KP6 Multicomponent doline site, MSA 1, Howiesons Poort, LSA KP7 Doline site, mixed LSA and Fauresmith (?), ‘items that compare best with the Acheulean assemblage from stratum 4a at KP1 KP8 Doline site, stratified LSA KP9 Doline site, Fauresmith (?), Howiesons Poort, LSA KP10 “Identical to KP8”, doline site, stratified LSA KP11 Small paleochannel with Ceramic LSA scatter

and LSA deposits. Acheulean assemblages were not recovered from other sites, but KP7 and KP9 yielded artifacts that Beaumont suggests are similar to the stratum 4a Fauresmith-designated assemblage at KP1.

KP1 was excavated in a total of about two and half months over two field seasons between 1979 and 1982 with teams of 8-14 people. To access the artifact-rich levels, Beaumont removed near- sterile sediments over an area of ~50 m2 and a depth of ~3 m2 before starting systematic excavations. Based on what is housed at the McGregor Museum (see below), the material removed must equate to what Beaumont identifies as strata 1 and 2, with systematic excavations guided by a 1m by 1m grid beginning at the top of stratum 3. Excavation procedures are not described in further detail by Beaumont (1990b; 2004a), but he refers to a 3mm mesh sieve (Beaumont, 1990b:82) and based on the well-labeled collection contents (see below), he excavated by arbitrary spits of 10 cm for most units, though occasionally 20 cm spits were used.

Beaumont identified 5 strata at KP1 (Beaumont, 1990b; Beaumont, 2004a) that are summarized in Table 7. Stratum 1is characterized by 1.5 -2 m of interdigitating calcified sand and organic peats. Stratum 2 is characterized by well-sorted aeolian sand that becomes increasingly calcified toward the top. Artifacts in strata 1 and 2 are very sparse, but Beaumont tentatively suggests a ceramic LSA or designation for stratum 1 and potential Robberg affiliation for stratum 2. Stratum 3 is characterized by a MSA-designated assemblage recovered from gravel with sub-

86

Table 7 Summary of KP1 stratigraphy. After Beaumont (Beaumont, 1990b; Beaumont, 2004a), Klein (Klein, 1988), and Porat et al. (2010) from the top of the sequence (stratum 1) to the bottom (stratum 5).

Stratum Lithostratigraphy Faunal Assemblage Lithic Assemblage Age (Porat et al. 2010)

1 1.5 -2 m of interdigitating calcified sand - Ceramic Later Stone - and organic peats indicative of Age or Iron Age? marshland or standing water, lies immediately below modern topsoil

2 1.8 m of well-sorted aeolian sand that Ostrich eggshell beads, Later Stone Age with OSL exhibits increasing calcification towards consistent with Later Stone possible Robberg 10.0±0.6 ka the top, implies arid phase Age occupation affiliation? 16.5±1.0 ka

3 0.8 m of unstratified gravel in grey-white Narrow range of taxa, Middle Stone Age OSL 291±45 sand matrix, possibly deposited by predominately grazers ka paleostream or intermittent flash floods that swept across the Kathu area (Butzer, 1974)

4 1- 4 m of uncalcified yellowish- 4a Mainly grazers indicating a Fauresmith OSL 464±47 white sand that was deposited grassy, savannah environment, when seasonal pond levels were consistent with U-series-ESR higher than any later time, small lithostratigraphic interpretation 542+140-/- vertical zones of coarser debris that of high water levels 107 ka represent old spring vents, divided into two substrata based on faunal and lithic components 4b Mainly grazers indicating a Acheulean ~2.85-400 ka, grassy, savannah environment, based on time recki resemble Todd’s range for Group III, which spans the Elephas recki broad time range of 2.85 – 400 ka >435-682 ka based on dates for stratum 4a

5 Up to 3.5 m of orange aeolian sand down - sterile - to bedrock

angular to sub-rounded pebbles in a greyish sand matrix. Stratum 4 consists of two levels, 4a and 4b, which are distinguished from each other based mainly on the lithic and faunal assemblages, though Beaumont (1990b:76) also reported a thin pebble lens that divided the subunits. The artifacts were recovered from a yellowish sand matrix. Stratum 4 includes spring vent features

87 that cut vertically through the stratum. These features were identified and described by Beaumont (1990b), but excavations were conducted by arbitrary levels and did not follow these features. The stratum 4a assemblage was designated to the Fauresmith, and the underlying stratum 4b assemblage was designated to the Acheulean. Stratum 5 is calcified pale orange sand and sterile.

4.2.2 Fauna and palaeoenvironmental interpretations

Richard Klein (1988) studied the entire KP1 faunal assemblage from the1979-1980 excavation. The faunal component at KP1 is heavily fragmented. Of the identification that could be made, the most abundant species in stratum 4 are Reck's (Elephus recki), equids, white , and -size alcelphines. In stratum 3, Reck's elephant is absent, Burchell's zebra and giant alcelaphines appear, and bovines are relatively abundant (Klein, 1988). Based on the presence of mainly grazers, the palaeoenvironment is interpreted as grassy savannah. Klein also draws potential chronological ties between the KP1 Acheulean assemblage and Elandsfontein.

Klein (1988) did not distinguish stratum 4a and 4b, but lumped spits 80-420 cm as the Acheulean levels, and 0-80 as the MSA levels. However, there are some differences in the upper levels of stratum 4, 80-180 cm, which correlates roughly with stratum 4a, compared to the lower levels, which correlate to stratum 4b. The upper levels (80-180 cm) contain Burchell's zebra, hippopotamus, warthog, large tragelaphines, reedbuck, small alcelaphines, and the giant long- horned buffalo, while the lower levels do not (see also Beaumont, 1990b; Porat et al., 2010). Only stratum 4b contains Elephas recki. If hominins were at least in part responsible for the accumulation of fauna at KP1, than the appearance of new species in stratum 4a compared to stratum 4b could indicate dietary changes and/or new strategies for accessing large game. However, Klein (1988) warns against making behavioral assumptions about the KP1 faunal assemblage specifically, and most Acheulean faunal assemblages, generally. Hunting or scavenging by hominins is only one of several possibilities for the accumulation of faunal remains at a pan/spring like Kathu Pan. Furthermore, the boundary between stratum 4a and 4b is not at the same level across the site (see below). Assessing the significance of the KP1 faunal assemblage requires further investigation of the curated material and new excavations.

88

The sedimentary sequence of KP1 provides some further palaeoenvironmental information and suggests that, in addition to moister periods experienced at Kathu Pan, there were intervals during the Pleistocene that were substantially drier than any part of the Holocene (Butzer, 1984). The desiccated beds might indicate that during periods when the water table dropped below those beds, protracted peat fires burned out the organic matter. Butzer (1984) suggested the following sequence of events from the bottom of the sequence to the top:

 Subsurface karstic solution and formation of doline, followed by draining out of residual sediments during period of low water table  KV-4 (Stratum 5 in Beaumont, 1990b) wet, with peat marsh and perennial water, slow accumulation of fresh sediment at base of doline  Water table drops, protracted peat fires  KV-3B (Stratum 4 in Beaumont, 1990b) pan with seasonal water, then periods with high water table and spring eruptions (Acheulean and Fauresmith)  Water table drops, protracted peat fires  KV-3A (Stratum 3 in Beaumont, 1990b) flash floods washed in poorly sorted detritus with MSA artifacts, lower water levels (MSA)  KV-2 (Stratum 2 in Beaumont, 1990b) decrease in sedimentation rates (primary or derivative aeolian component) and establishment of current pan environment  KV-1 (Stratum 1 in Beaumont, 1990b)

Together the sedimentary sequence and faunal assemblage indicate that there were climatic fluctuations during the Pleistocene that affected water table levels and the mammalian populations at Kathu Pan. Faunal remains of hippopotamus might indicate that some periods were moister than the current conditions, presumably periods associated with human occupations, and sedimentary observations suggest that there were periods drier than any part of the Holocene. For now, these observations are indefinite and unspecific. Future work at KP1 focused on site formation processes and paleoenvironmental reconstruction could lead to stronger hypotheses about the influence of the local environment on hominin behaviors at Kathu Pan during the Pleistocene.

89

4.2.3 Dating the KP1 sequence

Recent investigations (Porat et al., 2010) were able to confirm many of the stratigraphic observations reported by Beaumont and reported absolute dates for strata 2-4 (Figure 4). In 2004, KP1 was reopened and Beaumont’s original ~4.5m section on the north wall was exposed and cleaned back up to 10 cm. The newly exposed section correlates well with the section published by Beaumont (1990b).

In the process of cleaning the section, two well-defined vertically-oriented spring vents in stratum 4a were exposed and described as the ‘Upper Vent’ and ‘Lower Vent’. The Upper Vent is densely packed with lithic artifacts and fauna, and the area outside the vent contains few, in any, artifacts. The Upper Vent is in the uppermost levels of stratum 4a and cut-off at the top by Stratum 3.

In 2004, after exposing the KP1 section, four sediment samples were collected from the section for OSL analysis, and one tooth sample for combined U-series/ESR. The OSL samples were collected by horizontally drilling 30–40 cm deep holes into the section with a handheld auger. Under an opaque sheet, the sediment samples were collected from the holes and placed in light- proof black bags. Samples were processed at the Geological Survey of Israel. Samples #1 and #2 were collected from stratum 2, sample #3 from stratum 3, and sample #4 from inside the Upper Vent in stratum 4a. Further details about the methods used for OSL age determination are published by Porat et al. (2010).

An upper cheek tooth of capensis was processed for combined U-series/ESR age determination at the ANU laboratory (ANU sample # 2261). This tooth was recovered from inside the Upper Vent directly associated with the concentrated lithic artifacts and adjacent to the location of OSL sample #4. Further details about the U-series/ESR methods are published by Porat et al. (2010).

The OSL ages for the KP1 section are in stratigraphic order. Sample #1 gave an age estimate of 10.0±0.6 ka and sample #2 gave an age estimate of 16.5±1.0 ka. Sample #3 provided an age of 291±45 ka.

90

Figure 4 Stratigraphic profile of KP1 from 2004 investigations (Porat et al., 2010) showing chronometric age estimates and methods (looking north).

OSL sample #4, collected from the Stratum 4a Upper Vent with Fauresmith artifacts, gave an age of 464±47 ka.

It is important to note here that at KP1, it is possible to get older OSL age estimates than is possible at many other South African sites, which tend to max out at ~ 200 ka (e.g. Jacobs, 2010). The upper limit of age by OSL is determined by the annual dose on the sediment, which is related to its content of U, Th, and K. Low levels of these in the sediment lead to very slow saturation of quartz. At KP1 the dose rate is low and quartz grains did not saturate despite the antiquity of the sediments. For OSL sample #3 at KP1, the equivalent does (De) value is only 138 Gy, well within the acceptable range of OSL dating of quartz (Porat et al., 2010). For OSL sample #4, the De value is 303 Gy, which is close to the current limit of OSL dating, implying that this age could even be a minimum estimate. Large error ranges given for samples #3 and #4 cover uncertainties associated with water content and burial history (Porat et al., 2010).

91

From the the E. capensis tooth, U-series/ESR ages of 497 +182/-138 ka (sample #2261A) and 608 +216/-169 ka (sample # 2261B) were obtained. This results in a weighted mean age of 542 +140/-107 ka.

With the significant error ranges taken into account, the two dating methods indicate that stratum 4a dates to more than 435 ka and less than 682 ka. OSL indicates that sand grains associated with the lithic and faunal assemblage were buried at least ~417 ka. U-series/ESR directly dates the burial of a large E. capensis tooth, which was at least 435 ka, but could have been as much as 682 ka. The two dating methods are independent, yet are consistent with each other. The most reasonable interpretation is that the stratum 4a sediments, lithic, and faunal finds that are concentrated with each other inside the Upper Vent were buried ~500 ka, but at least 435 ka. This is thus far the oldest age estimate for a Fauresmith assemblage (Porat et al., 2010). The overlying stratum 3 also provides one of the earliest OSL ages for an MSA assemblage at 291±45 ka, and provides additional chronological control for the underlying stratum 4a because it cross-cuts and overlies the Upper Vent.

4.3 Lithic Collection Catalogue and Documentation

All remains excavated from the 1978-1990 excavations of Kathu Pan 1 are housed at the McGregor Museum, Kimberley, Northern Cape Province, South Africa. Prior to the study described here, the lithic collection of Kathu Pan had not been described beyond basic industrial assignment (Beaumont, 1990b; Beaumont, 2004a) and a preliminary analysis based on a sample of 217 flakes and blades as part of the dating study described above (Porat et al., 2010). Because no database of the lithic collection detailing its contents has even been created, the first step in analyzing the collection was to document it. The collection existed within 253 cardboard boxes and each box was marked with the site ID number (6538) and provenience information (square and spit). Lithic, faunal, and other remains were boxed separately from each other, but otherwise, artifacts were entirely unsorted prior to this analysis. Every individual lithic artifact was hand- labeled with provenience data, thus, based on the data published in the guidebooks (Beaumont, 1990b; Beaumont, 2004a), it was possible to link every specimen to a 10 or 20 cm spit within one square meter of the excavation.

Initial documentation of the collection required identifying and taking the mass of lithic remains from each individual unit present in the collection. A summary of lithic contents by square is

92 presented in Table 8. The documentation process revealed that 20 square units were opened and excavated by arbitrary levels of 10 or 20 cm spits (Figure 5, Table 8). An additional three square units were dug, but for Y23, only LCTs between levels 260-310 are present in the collection, and for Z24 and Z25, only faunal remains from levels 280-310 are present in the collection. Documenting the collection in this way also revealed that there are four additional excavated squares not previously recorded in any available reports. These data also permitted a 3D reconstruction of the excavations and a density map to examine the original horizontal and vertical distribution of lithic artifacts at the site (Figure 5).

4.3.1 Assigning levels to strata

Spits were recorded as the depth (i.e., 60-70) in centimeters below the top of stratum 3, but were not identified with respect to what stratum Beaumont identified them with. According to Beaumont (1990b:81) levels 0-60 are generally stratum 3, 80-180 are stratum 4a, and 180+ are stratum 4b, but the levels for the strata are not the same in each square unit. Fortunately, I was able to use sedimentological and technological evidence to refine the assignments the squares I analyzed and confidently attribute most levels to either stratum 3, 4a, or 4b.

Sedimentological evidence consisted of correlated curated sediment samples from the collection with images of the site profile (Porat et al., 2010), and descriptions by Beaumont (1990b). Sediment samples were available for many levels in squares F23, C23, and C21. For squares F23 and C23 in the center/eastern area of the excavation, there is a clear shift in sediment color at 60 cm (Figure 4, Figure 6). Sediments above are a light-brownish grey (10YR6/2) gravelly sand and sediments below are a pinkish grey (7.5YR7/2) gravelly sand. The brownish grey gravelly sand sediments are consistent with those described for stratum 3 and the pinkish grey gravelly sand are consistent with those described for stratum 4. For square C21 in the more eastern area of the excavations, the change in sediment coloration occurred at a slightly higher depth. Sediments above 40 cm are light-brownish grey; sediments below 40 cm are pinkish grey.

The assignment of levels above 40 cm to stratum 3 is corroborated by technological evidence; the assemblage from these levels lack the blades and points that characterize the underlying levels attributed to stratum 4a. Furthermore, the artifacts from the levels that I assign to stratum 3 based on sedimentological and technological criteria are also heavily rounded and waterworn

93

Table 8 Summary of lithic contents and excavated levels by square in the KP1 lithic collection that is currently housed at the McGregor Museum, Kimberley Square # Boxes Total Mass Upper Depth Lower Depth Excavated Levels Excavated Levels Sterile Volume Estimated Density (kg/m3) of Excavation Spits (kg) Present (cm) Present (cm) Upper Lower Levels Excavated (m3) lithic artifacts (Beaumont, (Beaumont, 1990b) 1990b) A23 34 395 140 310 140 340 310-340 2 197.50 140-200: 20 cm, 200-310: 10 cm

A24 11 133 210 310 210 310 - 1 133.00 10 cm

A25 9 94 210 310 210 310 - 1 94.00 10 cm

B20 7 59 0 200 0 200 - 2 29.50 20 cm

B21 2 17 10 60 10 60 - 0.5 34.00 10 cm

B23 10 106 20 310 20 310 - 2.9 36.55 20-60: 20 cm, 60-310: 10 cm

C20 19 233 0 380 0 380 - 3.8 61.32 20 cm

C21 6 67 0 120 0 120 - 1.2 55.83 20 cm

C23 9 93 20 310 20 310 - 2.9 32.07 20-40: 20 cm, 40-310: 10 cm

D20 3 24 0 110 0 110 - 1.1 21.82 20-40: 20 cm, 40-110: 10 cm

D21 7 68 0 150 0 150 - 1.5 45.33 0-120: 20 cm, 120-150: 10 cm

E20 4 32 0 110 0 110 - 1.1 29.09 0-30: 30 cm, 30-90, 10 cm, 90-110, 20 cm E21 9 80 0 180 0 180 - 1.8 44.44 0-40: 20 cm, 40-180: 10 cm

F20 2 12 40 110 40 110 - 0.7 17.14 40-60: 20 cm, 60-110: 10 cm

F21 7 68 20 180 20 180 - 1.6 42.50 20-60: 20 cm, 60-180, 10 cm

F22 8 73 130 260 130 260 - 1.3 56.15 130-240: 10 cm, 240-260, 20 cm

F23 18 158 0 250 0 250 - 2.5 63.20 0-20: 20 cm, 20-250: 10 cm

F25 8 62 Level 1 Level 30 not reported n/a - n/a n/a 30 Levels

F26 10 90 Level 1 Level 20 not reported n/a - n/a n/a 20 Levels

Y23 1 3 260 310 260 310 - 0.5 6.00 Only handaxes kept

Z23 19 214 220 420 220 770 420-770 5.5 38.91 220-260: 20 cm, 260-420, 10 cm

94

Figure 5 Three-dimensional reconstruction of KP1 excavations showing lithic artifact density for 20 cm spits (kg/m3).

compared to the stratum 4a lithic artifacts that have edges that are relatively fresh in condition

There is no clear change in sediment reflective of the 4a-4b transition, so the sediment samples could not be used to assign levels to stratum 4a or 4b. However, the technological evidence alone was sufficient to confidently assign most levels to either stratum 4a or 4b. Stratum 4b levels are distinct from stratum 4a levels for lacking blades and points and having several LCTs. The stratum 4a artifacts are also fresher with sharper, unworn edges compared to the stratum 4b artifacts.

Figure 7 demonstrates the clear technological differences between levels designated to strata 3, 4a, and 4b. These data were compiled using three samples from the KP1 assemblage:

(1) The stratum 3 sample consists of all lithic artifacts recovered from square C23, level 20-40 (n=892).

(2) The stratum 4a sample consists of all lithic artifacts recovered from squares F23, F21, C23, C21 (n=6052). This is the sample that formed the basis for the rest

95

Figure 6 Correlation of sediment samples, level, and stratum.

96

Figure 7 Lithic artifact type frequencies for strata 3, 4a, and 4b. Stratum 3 sample from square C23 (levels 20-40, n=892). Stratum 4a sample from squares F23, F21, C23, C21 (n=6052). Stratum 4b sample from square F21 (levels 160-180, n=241)

of the stratum 4a analysis as presented in Chapters 5, 6, and 7, which is why the sample size is so much larger than that for strata 3 and 4b. Further details about this sample are presented in Chapter 5.

(3) The stratum 4b sample consists of all lithic artifacts for square F21, levels 160- 180 (n=241). This is the sample that forms the basis of my strata 4b raw material analysis that I present in Chapter 7.

97

Ideally, I would have data from the entire KP1 lithic assemblage, but the enormity of the collection made that an unreasonable goal and beyond the scope of this dissertation. The stratum 4a sample is larger for this technological comparison, because it is the one dated to the early Middle Pleistocene and relevant to the issues concerning the shift from ESA to MSA technologies that are the focus of my dissertation. The strata 3 and 4b samples were selected solely to address specific questions about comparability to stratum 4a.

From these samples it is apparent that the strata 3, 4a, and 4b assemblages differ radically from each other. The stratum 3 assemblage represents a mainly flake-based industry with Levallois flakes and relatively rare retouched points (Figure 7). These points require further examination, but both unifacial and bifacial points were observed. Blades occur in relatively low frequencies in stratum 3; blades make up less than 4% of all the lithic artifacts (Figure 7). In contrast, the underlying stratum 4a assemblage contains abundant blades (16%) and retouched points. LCTs are present, but in very low frequencies. The underlying stratum 4b assemblage contains abundant LCTs (nearly 6% of lithic artifacts), no blades, points, or Levallois flakes. Thus, the technological distinctions between the three excavated strata originally identified by Beaumont (1990b) are clear. His macroscale correlations between KP1 stratigraphy and technology are reliable and can be reconstructed based on the carefully curated collection that is housed at the McGregor Museum.

4.3.2 The relationship between the dated samples and the curated assemblage

The artifacts from the Upper Vent are the same assemblage that Beaumont designated as Fauresmith. This is secure based on two lines of evidence; the stratigraphy and the technological characteristics of the lithic artifacts themselves. Based on Figures 3 and 4 in Porat et al. (2010), OSL sample #4 was taken ~40 cm below the base on stratum 3. The tooth sample for ESR was recovered ~10 cm below the OSL sample #4. Thus, the two samples were taken within ~80-110 cm below the top of stratum 3. The location of Upper Vent in the exposed section would have abutted square F21, and the stratum 4a designated units square F21 are from 40-140 cm below the top of stratum 3. In other words, the depths associated with the stratum 4a lithic artifacts that I examine in this study span the depths of the dating sediment and tooth samples. In addition, sediment samples from 40-140 cm below the top of stratum 3 collected during Beaumont’s

98 excavations (see above) are consistent with sediments at that depth in the newly exposed profile and within the Upper Vent (Figure 6).

In 2004, the Upper Vent was partially excavated to collect a small sample of lithic artifacts (n=34) for comparison with the curated assemblage at the McGregor Museum. The lithic artifacts in the Upper Vent have relatively fresh edges, and there are blades and convergent points with facetted platforms. A statistical comparison of elongation between the Upper Vent flakes and blades, and a sample from Beaumont’s excavation in square F21 showed that there was no significant difference between the Upper Vent sample and Beaumont’s excavated material in this respect (Porat et al., 2010:272). Furthermore, as discussed above, the stratum 4a assemblage is very distinct from the overlying stratum 3 MSA assemblage, which is heavily rolled and has very few blades. It is also distinct from the underlying 4b, which lacks blades and prepared cores and is rich in LCTs. Technologically, the Upper Vent material is consistent with the stratum 4a Fauresmith-designated assemblage that is housed at the McGregor Museum.

4.4 A consideration of site formation processes

As a doline site, KP1 probably has a complex taphonomic history. Like many archaeological assemblages, especially Paleolithic ones, the artifacts at KP1 are not recovered from primary context. This is especially true for the MSA assemblage, which is heavily-rolled and water-worn and probably washed into the site with its gravelly matrix. The concentration of artifacts in the spring vents in stratum 4a demonstrates that geological processes have affected the location and position of stratum 4a artifacts, as well. The relatively fresh condition of the stratum 4a lithic artifacts do suggest however, that if the material was transported to the site by water, it was probably not over a very long distance. In other words, hominin activities during the stratum 4a occupation were most likely occurring in or near the pan.

There are two general types of site formation processes that influence lithic assemblages; subtractive distortion and additive distortion (Schick, 1984:131-132). Both these processes are probably responsible for the current state of the stratum 4a lithic assemblage. Subtractive distortion is the winnowing of the small fraction of lithic debitage, and additive distortion is the hydraulic reconcentration of different debitage events, the effect of these processes must be kept in mind when interpreting lithic assemblage characteristics. To examine the potential effects of

99 these processes on the stratum 4a assemblage at KP1 I looked at size distribution and weathering data on two subsamples of the stratum 4a assemblage.

4.4.1 Artifact size distribution

The winnowing of small artifacts from archaeological occurrences is common and can be identified using size distribution data. Lithic artifact assemblages representing in situ knapping events have high frequencies (>65 %) of small flaking debris (<20 mm in maximum length)(Schick, 1984). Maximum length of each lithic artifact was recorded for the stratum 4a sample from square F23 (levels 60-130, n=1938). A comparison of the KP1 lithic size distribution to 107 experimental lithic knapping studies (Schick, 1984) indicates that small size categories are underrepresented in the KP1 assemblage (Figure 8). The small lithic component may have been winnowed away due to alluvial processes before and/or after secondary within the doline deposit. It is also not entirely clear how recovery techniques may have affected the size distribution at KP1. Sediments were sieved, presumably with a screen at least 15 mm in size based on the abundance of artifacts greater than this in maximum dimension, and the rarity of artifacts smaller than this. It is likely that 3 mm sieve was used during excavation (see above), but if not, then the size fraction less than 15 mm would not be present in the curated assemblage anyway. Future excavations are the only way to confidently establish whether the small fraction is present in the KP1 stratum 4a sediments or not.

4.4.2 Weathering

Characteristics of surface weathering were also recorded to explore how taphonomic factors may have influenced the lithic assemblage. Weathering is the result of irreversible chemical and mechanical changes induced by the local environment, before and after burial. Differences in weathering have sometimes proven useful for separating artifacts from two different components at mixed sites (e.g. Purdy and Clark, 1987). Different weathering states could potentially indicate that additive distortion has affected assemblage composition in stratum 4a, resulting in a palimpsest of different lithic discard events with different depositional histories. The KP1 assemblage exhibits many types of weathering.

100

Figure 8 Size distribution of lithic artifacts. Experimental data from 107 knapping episodes (Schick, 1984). 4a sample from square F23 (levels 60-130, n=1938).

The following weathering states were assigned to each artifact in the analyzed sample, which included complete flakes, proximal flake fragments, blades, cores, retouched pieces, and LCTs from squares F23, F21, C23, C21 (n=1693): 0. No significant weathering 1. Rolling/rounding -- smoothed arises and rounded edges 2. White (sometimes pinkish) chalky patination on otherwise fresh artifacts 3. Severe disintegration that leaves artifacts with a yellowish white surface and lightweight 4. Rusty iron deposits (orange-red) on patinated, but otherwise fresh pieces 5. Glossy patina resulting in shiny and smooth silica deposit

101

Figure 9 presents examples of each type of weathering. Artifacts sometimes exhibited more than one weathering state and in those cases the artifact was included in frequency counts for each weathering state. Weathering type 2 – patinated but otherwise fresh condition - is the most common state for all artifact types in stratum 4a, with the exception of LCTs (Figure 10). The development of whitish patina is common on dark cherts and may be due to the dissolution of silica from the lithic surfaces in an alkaline environment (Luedtke, 1992:99) or the leaching of iron-oxides from the lithic surface in an acidic environment (Clark and Purdy, 1979).

Weathering type 5 is rare in the stratum 4a assemblage (Figure 10), but common in the stratum 4b levels of some squares. Some cherts develop gloss like this as a result of silica dissolution and re-precipitation on artifact surfaces, often resulting from exposure to water solutions for extended periods of time (Howard, 2002). High water table levels at KP1 could have created the soil conditions necessary to develop this kind of glossy patina. The increased frequencies of artifacts with glossy patina in stratum 4b compared to stratum 4a, and in the lowest levels of stratum 4a (Figure 11), are consistent with standing water due to high water table levels.

The majority of LCTs are rolled (weathering state 1, Figure 10), suggesting higher-energy water transport of LCTs, or longer surface exposure time prior to burial, compared to the majority of other artifact types. Thus, non-behavioral processes may explain the association of the LCTs with the rest of the stratum 4a assemblage. The implications of this observation are discussed further in Chapters 3 and 8.

All other artifact types exhibit similar frequencies of each the five different weathering states (Figure 10). There are multiple ways to interpret this observation. It is possible that the different weathering states reflect different discard events and the stratum 4a assemblage is a palimpsest of multiple periods. Even if this is true, there are technological similarities between the periods, since blades and points exhibit the whole range of weathering states. Also, weathering is dependent on lithic characteristics (structure and mineral composition) and the sedimentary micro- environment. Raw material type (e.g., banded ironstone vs. volcanic vs. quartzite) does account for some of the variety in weathering states, but not all of it. Even within a single raw material category such as banded ironstone, there are different weathering states present.

102

Figure 9 Example weathering states of KP1 lithic artifacts. 0. No significant weathering, 1. Rolling/rounded edge wear, 2. White (sometimes pinkish) chalky patination on otherwise fresh artifacts, 3. Severe disintegration that leaves artifacts with a yellowish white surface and lightweight, 4. Rusty iron deposits (orange-red) on patinated pieces, 5. Glossy patina resulting in shiny and smooth silica deposit.

This is explained at least in part by differences in mineral composition in the heterogenous banded ironstone, as evidenced on individual artifacts that exhibit different states of weathering for each type of band. Different types of chert, for example, may weather differently in the same environment, and the same type of chert may weather differently in the same environment (Luedtke, 1992:98). For this reason, it is not possible to determine which of two interpretations, palimpsest or differential in situ weathering, best explains the multiple types of weathering exhibited by the stratum 4a lithic artifacts, although a combination of these two explanations is likely.

4.5 Summary

The stratum 4a archaeological assemblage housed at the McGregor Museum imposes some analytical limitations. KP1 experienced a complex taphonomic history that remains largely

103

Figure 10 Frequency of weathering states for different artifact types in stratum 4a. Stratum 4a sample complete flakes, proximal flake fragments, blades, cores, retouched pieces, and LCTs from squares F23, F21, C23, C21 (n=1693).

unknown. Only further excavations focused on addressing questions of site formation can improve this situation. The lithic assemblage exhibits some characteristics consistent with distortion due to water transport. The smallest fraction of the lithic assemblage is missing. The artifacts are weathered and there are a variety of weathering states represented. This variety in weathering states could indicate additive distortion and a potential palimpsest of different discard events over a significant period of time.

Despite known issues due to unknown site formation processes, the stratum 4a lithic assemblage from KP 1 represents discarded material from a period for which little is known, and is suitable for identifying and describing early Middle Pleistocene technological behaviors. As a stratified Stone Age deposit with dated sediments and associated faunal remains, KP1 is an extremely rare

104

Figure 11 Frequency of weathering states for different levels within stratum 4a. Stratum 4a sample complete flakes, proximal flake fragments, blades, cores, retouched pieces, and LCTs from squares F23 and F21, and retouched points and LCTs from C21 and C23 (n=1597).

occurrence. The burial age of the lithic material is at least 435 ka based on the OSL and U- series/ESR age estimates, indicating that even if the material collected on the surface for some time before being deposited in the doline, the lithics were discarded before 435 ka. Most stratum 4a lithic artifacts have fresh edges, implying that even though they are not in primary context, most were not transported long distances. It appears that during the original excavation, all material was collected from sieves at least15 mm (but probably 3 mm), so all debitage > 15 mm is available for study. Provenience to square and depth below the top of stratum 3 was hand- written on each artifact and consequently, the museum collection is in a good state for further analysis.

105

Furthermore, all stages of lithic manufacture are present, as indicated by the presence of cortical pieces, cores, blanks, retouched pieces, and core-trimming elements, and this assertion is demonstrated in Chapter 5. Also in Chapter 5, it will be demonstrated that the flakes, blades, cores and retouched pieces are linked technologically as part of the same chaînes opératoires.

106

5 Technological Analysis of KP1 Stratum 4a Lithic Assemblage In this chapter, I present the results of a technological analysis of the KP1 stratum 4a lithic assemblage. Some of this research has been published in Journal of Archaeological Science (Wilkins and Chazan, 2012), and is reprinted here with permission from Elsevier. Characterization of flake and blade production methods is accomplished primarily through technical classification of blade, flake and core types, as well as attribute and metric analyses of subsamples within each category. The procedures and attributes used here were selected for the purpose of addressing these questions about the knapping activities represented in the assemblage: 1. How were raw materials transported to the site, as complete nodules, partially worked, or as preformed blanks)? 2. How were raw material nodules prepared for blank production? 3. For Levallois cores, was exploitation recurrent or preferential? If recurrent, how was the recurrent surface exploited (bidirectionally, unidirectionally, or centripetally) (Boëda, 1986, 1991, 1995; Boëda et al., 1990; Boëda and Pelegrin, 1980)? 4. How was the core maintained? 5. What were the characteristics of the blanks removed from cores? What were the objectives of core reduction? 6. Which pieces were chosen for retouch? How were they retouched? 7. What techniques were used for flake removals? What was the knapping tool? 8. Is the assemblage characterized by one overall “model reduction sequence”? Or are there multiple sequences? 9. How do the core reduction sequences at KP1 compare to other Pleistocene sites and what does this tell us about hominin technological behavior in the early Middle Pleistocene?

5.1 Methods

5.1.1 Sample selection

All lithic remains from stratum 4a in four square units, F21, F23, C21, C23, were analyzed. F21 was chosen because it was the focus of an initial analysis (Porat et al. 2010) of Fauresmith lithic reduction at KP1.The additional square units were chosen to ensure that a wide area of the site

107

Table 9 Analyzed 4a assemblage. Square Unit F23 F21 C23 C21 Analyzed 4a Levels (cm below top of stratum 3) 60-70 40-60 60-70 40-60 70-80 60-70 70-80 60-80 80-90 70-80 80-100 90-100 80-90 100-120 100-110 90-100 110-120 100-110 120-130 110-120 120-130 130-140 Total Number of lithic artifacts 1938 2264 302 1548

was examined, and because sediment samples were available for squares C23 and C21 to aid in establishing the upper boundary of stratum 4a. Table 9 summarizes the selected 4a sample. A total of 6052 lithics, which represent the complete stratum 4a assemblage from four out of twenty excavated square units, were analyzed.

5.1.2 Analysis

Analytical procedures were adapted from those outlined by Villa (2005) and Soriano et al. (2007) for South African Middle Stone Age assemblages. The analysis began with sorting based on basic lithic categories. The relative proportions of these basic lithic categories provide a general overview of assemblage composition and can be used to evaluate whether all stages of the knapping sequence occurred on site. Artifacts that preserve sufficient technological information - incomplete and complete flakes, blades and blade fragments, cores and core fragments and retouched pieces - were then selected from the assemblage for technological analysis. Villa (2005:405) advocates for a practice of laying out all pieces so that the process of grouping lithic products into discrete categories can be consistent and this procedure was followed here. This method has proved successful for describing lithic reduction strategies represented by other MSA blade assemblages, and shared procedures with these other analyses permits direct comparison of technical counts.

Following sorting, metric and technological attributes were recorded for a subsample of each technical category. Definitions for terms used can be found in the relevant results tables.

108

5.1.2.1 Technical categories

Selected lithic artifacts (n=3946), which include all proximal flake fragments, complete flakes, blades, including fragments, cores, and retouched pieces, including retouched points and LCTs, were grouped into technical categories. Counts of technical categories provide an overview of lithic products for consideration of reduction sequences (i.e., presence/absence of specific categories, relative frequencies of different categories), permit comparisons of blade frequencies between and within square units at KP1 and with other sites that have used the same classificatory schema, and when coupled with a detailed attribute analysis, provide the basis for interpreting and describing lithic reduction strategies.

Technical categories for blades (n=972) followed those in Soriano et al. (2007). Blades are defined as detached lithic pieces that are at least as twice as long as they are wide along the axis of percussion. Blade fragments are recognized based on their parallel edges and arises. Flake (n=1544) technical categories were created based on patterns observed within the KP1 flake assemblage and modeled after the numerical system used for blades. Cores (n=679) were categorized based on flake scar orientation and volumetric form. Any piece that was shaped or modified with retouch (n=731) was assigned to a typological category based on the distribution and location of retouch.

5.1.2.2 Attribute analysis

Metric and technological attributes were recorded for a subsample of the KP1 4a assemblage (Table 10). Subsamples were used to deal with time constraints, but large enough samples were obtained so that statistical tests could be used to test for significance. The attributes were recorded for blades (n=513) and flakes (n=227) from squares F23 and F21 and cores (n=123) and retouched pieces from square F23 (n=127). Attributes were recorded for all retouched points (n=149) and LCTs (n=20) from squares F21, F23, C21, and C23.

5.1.2.3 Bulb and platform analysis

A bulb and platform analysis was conducted on all flakes, blades, and retouched pieces (with bulbs and platforms) from square F=23 (n=340) to determine the percussion technique following methods used by Soriano et al. (2007), Villa et al. (2010), and developed by Pelegrin (2000). Table 11 summarizes the recorded attributes. For comparative purposes, an experimental sample

109

Table 10 Attributes recorded for the subsample from each technical category. Blades Flakes (with bulb) Cores Retouched pieces Retouched Points LCTs (exc. retouched (inc. fragments) pointed forms) n=513 n=227 n=123 n=127 n=149 n=20 From squares F23 From squares F23 From square F23, From square F23, From squares F23, From squares F23, and F21, excludes and F21, excludes excludes rolled cores excludes rolled F21, C23, C21 F21, C23, C21 category E and category 5 and and cores that could pieces and pieces blades without flakes without not be categorized that could not be measurable measurable categorized attributes attributes completeness completeness maximum length completeness of completeness of technological length (proximal, distal, (proximal, distal, (mm) blank (proximal, blank (proximal, (mm) mesial, complete) mesial, complete) distal, mesial, distal, mesial, complete) complete) technological length technological length maximum width blank classification blank classification technological width (mm) (mm) (mm) (mm) technological width technological width maximum thickness length (mm) length (mm) thickness (mm) (mm) (mm) (mm) (technological for (technological for complete blades and complete blades and flakes, maximum for flakes, maximum for other) other) thickness (mm) thickness (mm) number of scars width (mm) width (mm) blank (outside the bulb) (outside the bulb) dorsal directionality dorsal directionality technological length thickness (mm) thickness (mm) (radial, subradial, (radial, subradial, of final removal bidirectional, bidirectional, (mm) unidirectional, unidirectional, cortical) cortical) technological width dorsal directionality dorsal directionality of final removal (radial, subradial, (radial, subradial, (mm) bidirectional, bidirectional, unidirectional, unidirectional, cortical) cortical) termination of final retouch distribution retouch distribution removal (hinge, step, (any combination of (any combination of feather) left, right, distal, left, right, distal, proximal, dorsal, and proximal, dorsal, and ventral) ventral) remnant platform additional traits for type (facetted (2), functional analysis multifaceted (more (see Chapter 6) than 2), plain, cortical) Levallois exploitation type (for Levallois cores) (preferential, centripetal recurrent, bidirectional recurrent, unidirectional recurrent)

110

Table 11 Bulb and platform analysis. Platform width (mm) - Platform thickness (mm) - Exterior platform angle A-acute angle, R-right or near right angle, O-obtuse angle Platform preparation C-cortex, P-plain, F-faceted (with bulbs), R-residual faceting, O-other, Ind- indeterminate Dorsal preparation N-none, H-hinged removals, R-reduction of overhang by longer removals, L-lateral notching, O-other, Ind-indeterminate Platform abrasion N-none, S-slight, H-high, V-very high Fracture initiation point N-non-expressed, C-centered, L-lateralized, Ind-indeterminate Platform delineation R-regular curve, C-curved and overhanging without a break, B-curved and overhanging with a break, D-double curved, L-rectilinear, I-irregular, Ind-indeterminate Fissuring on the N-none, T-total circular fissuring at point of impact, A-arcuate partial fissuring, C- Platform contoured Hertzian cone, O-other, Ind-inderterminate Marks N-none, R-ripples on ventral surface, S-split or shattered bulb, H- hertzian cone on ventral surface, C-crushing, O-other, I-indeterminate Lipping L-lipped, S-small lip, N-no lip Bulb L-large bulb, S-small bulb, N-non bulb Platform morphology P-punctiform, N-narrow, O-oval or triangular (restricted), Q-quadrangular or trapezoidal (not restricted), C-curved, Ch-chapeau de gendarme, O-other, Ind- indeterminate

(n=147) of debitage from hard hammer percussion was also subjected to the same bulb and platform analysis. Flakes were removed from banded ironstone cobbles local to the Kathu region using volcanic and quartzite hard hammers.

5.2 Results

5.2.1 Blade production

Blade production at KP1 was a regular and purposeful practice, i.e., systematic, as evidenced in part by the following characteristics of the assemblage: (1) blades are a numerous component of the assemblage, (2) they are relatively large in size and have a high length to width ratio, (3) discarded blade cores have a distinct morphology, (4) platforms and dorsal surfaces were often intensively prepared immediately prior to blade detachment, (5) and retouched pieces in the KP1 assemblage were selectively manufactured on blades. If blade production was an unintentional goal of core reduction at KP1, one would expect fewer numbers, less exaggerated elongation, and a lack of highly prepared cores designed to produce elongated endproducts.

111

Blades and blade fragments make up 16.1 % of the entire lithic assemblage (Table 12, Figure 12, Figure 13, Figure 14). Blades and blade fragments represent more than a quarter (27.0%) of all discarded detached pieces, including flakes and flake fragments.

A summary of the KP1 blade metrics is provided in Table 13. The average length of complete blades is approximately 70 mm, but there is great variability in blade size. The upper limits of blade length in this sample approaches 150 mm and the minimum length is less than 15 mm. A frequency histogram of blade width (Figure 15) can be used to get a sense of size distribution (see Soriano et al., 2007, Villa et al., 2010), providing a larger sample than length because blades often break transversely. KP1 blade size distribution includes blades that would technically be classified as bladelets (<12 mm in width, Soriano et al., 2007), but these small blades are just at the lower end of a unimodal blade size continuum. The mean length to width ratio of the KP1 blades is 2.5 to 1 (n=92, sd=0.4).

5.2.1.1 Core preparation

Based on the distribution of blade technical categories (Table 14), it is possible to argue that KP1 cores for blade production were usually prepared using centripetal flaking. Blades with at least 50% cortex are present (Figure 12a), but in low frequencies (3.8%) and there is a low frequency of blades with any cortex at all (4.7%). In contrast, 5.3% of complete flakes and proximal flake fragments exhibit >50% cortex and 12.8% exhibit some cortex (Figure 16). The low frequency of blades with cortex compared to flakes suggests that blade cores were prepared first by flake removals. The rarity of crested blades (0.3%, Table 14) and the presence of blades with centripetal flake dorsal scars on one side only (4.2%, Figure 14e,h) are most consistent with a reduction strategy that prepared the blade exploitation surface with centripetal flake removals.

5.2.1.2 Core maintenance

Cores were not maintained with second generation crested blades (Soriano et al., 2007) or neo- crested blades (Villa et al., 2010), as these blade types are absent from the KP1 assemblage (Table 14). Centripetal flaking of dorsal and lateral convexities rejuvenated surfaces for blade production as evidenced by the presence of blades with centripetal flake removals (4.2%, Figure 14).

112

Table 12 Assemblage composition of analyzed sample. Includes all lithic contents from stratum 4a levels of four square units, F23, F21, C23, and C21. F23 F21 C23 C21 Total

0.7 1.0 0.2 0.8 2.7

n % n % n % n % n % Small flaking debris any piece less than 15 mm in 139 7.2% 106 4.7% 7 2.3% 17 1.1% 269 4.4% <15 mm maximum length products of the knapping process Chunks that are angular in form with no 93 4.8% 66 2.9% 16 5.3% 50 3.2% 225 3.7% apparent bulb or platform Flake fragments no bulb preserved 347 17.9% 458 20.2% 56 18.5% 226 14.6% 1088 18.0% Proximal flake bulb preserved but broken 219 11.3% 196 8.7% 21 7.0% 153 9.9% 589 9.7% fragments complete bulb of percussion and Complete flakes 232 12.0% 414 18.3% 41 13.6% 268 17.3% 955 15.8% distal end preserved including incomplete blades and blade fragments, with length/width Blades 213 11.0% 401 17.7% 43 14.2% 315 20.3% 972 16.1% ratio of 2 or greater, estimated for fragments Cores pieces with negative flake scars 274 14.1% 227 10.0% 41 13.6% 137 8.9% 679 11.2% detached pieces that were modified Retouched pieces after production, excluding 174 9.0% 178 7.9% 47 15.6% 184 11.9% 582 9.6% retouched pointed forms and LCTs converging edges shaped into a Retouched pointed point (retouch is usually unifacial, 42 2.2% 51 2.3% 1 0.3% 55 3.6% 149 2.5% forms (inc. frags) can be along one or both lateral edges) bifacially worked nodules with LCTs 4 0.2% 2 0.1% 3 1.0% 11 0.7% 20 0.3% lenticular cross-section nodules with high iron-oxide content (haematite, specularite) that leave a 6 0.3% 7 0.3% 0 0.0% 20 1.3% 33 0.5% red, black, yellow, or purple streak fragmentary, reddened pieces with Burned pieces 29 1.5% 40 1.8% 0 0.0% 35 2.3% 104 1.7% pock-marks rounded cobbles with pitting and other forms of wear that might be 2 0.1% 2 0.1% 0 0.0% 0 0.0% 4 0.1% indicative of use with too much rounding and/or Too weathered/non- dissolution to confidently assign to 164 8.5% 116 5.1% 26 8.6% 77 5.0% 383 6.3% cultural one of the above categories or non- cultural pieces Excavated volume (m3) 0.7 1.0 0.2 0.8 2.7 Total number of lithic artifacts 1938 2264 302 1548 6052 Lithic artifact density (1/m3) 2769 2264 1510 1935 2241

113

Figure 12 Photographs of selected stratum 4a KP1 lithic artifacts. a. entirely cortical blade (A2), b, c, g. blades from optimal phase of debitage (B1), d. blade preserving a portion of the opposite and bidirectional scars (B3), e. blade preserving a portion of the opposite platform (B3), f. refitting fragments of proximal blade fragment from optimal phase of debitage (B1), h, j. unifacially retouched pointed forms on blade blanks, i. blade core, k. unifacially retouched pointed form on blade fragment blank. All on banded ironstone.

114

Figure 13 KP1 stratum 4a blades and proximal blade fragments from the optimal phase of debitage (BI). All on banded ironstone. Arrows on sketches represented scar directionality. Circles indicated the presence of a negative bulb of percussion.

115

Figure 14 Other blade types. a. debordant flake, b. blade with more than 50% cortex (A3), c. blade with distal cortex (B2), d. blade with lateral cortex (B6), e. blade with centripetal dorsal scars on one side (B9), f. blade preserving opposite striking platform (B3), g. blade with cortical steep back and distal cortex (B10), h. blade with centripetal dorsal scars on one side (B9). All on banded ironstone. Arrows on sketches represented scar directionality. Circles indicated the presence of a negative bulb of percussion.

116

Table 13 Summary of stratum 4a blade metrics for all blades and B1 blades (produced during optimal phase of debitage) from squares F23 and F21 (n=513). Excluding pieces without measureable attributes or with weathering or breakage that prevented assignment to a blade technical category

All blades BI blades Thickness Length L/W Length (mm) Width (mm) (mm) L/W (mm) Width (mm) Thickness (mm) n=113 n=509 n=512 n=92 n=61 n=327 n=329 n=48 Mean 69.7 28.2 8.1 2.5 66.0 26.7 6.8 2.5 Min 14.9 7.9 2.0 2.0 14.9 7.9 2.0 2.0 Max 147.4 59.4 40.4 4.5 117.2 54.8 19.2 3.5 St. dev 19.3 9.2 4.2 0.4 17.7 9.1 2.8 0.4

Figure 15 Histogram of blade width (mm) from squares F23 and F21 of KP1 (n=511).

.

117

Table 14 Summary of blade technical category frequencies for KP1 stratum 4a (n=972). Blade technical categories based on Soriano et al., 2007. Blades that are rolled or extensively weathered are excluded from these counts. F23 F21 C21 C23 Total n % n % n % n % n % A: Initial 11 5.2% 17 4.5% 9 3.0% 1 2.9% 38 4.1% A1 Crested blades with one or two prepared versants 1 0.5% 2 0.5% 0 0.0% 0 0.0% 3 0.3% A2 Entirely cortical blades 4 1.9% 2 0.5% 1 0.3% 0 0.0% 7 0.8% A3 Blades with more than 50% of cortex (or natural surface) 6 2.8% 13 3.4% 8 2.6% 1 2.9% 28 3.0%

B: Main production phase 153 71.8% 275 72.6% 244 80.3% 29 82.9% 701 75.3% Blades from the central part of the debitage surface 126 59.2% 228 60.2% 201 66.1% 27 77.1% 582 62.5% B1 Blades produced during the optimal phase of the debitage, without cortex, with 114 53.5% 216 57.0% 194 63.8% 24 68.6% 548 58.9% unidirectional or bidirectional scars B1a B1 blades with centered triangular cross-section 20 9.4% 42 11.1% 41 13.5% 6 17.1% 109 11.7% B1b B1 blades with off-centered triangular cross-section 17 8.0% 40 10.6% 22 7.2% 5 14.3% 84 9.0% B1c B1 blades with trapezoidal cross-section 66 31.0% 119 31.4% 120 39.5% 13 37.1% 318 34.2% B1d B1 blades with converging edges (points) 11 5.2% 15 4.0% 11 3.6% 0 0.0% 37 4.0% B2 Blades with distal cortical edge 4 1.9% 6 1.6% 5 1.6% 0 0.0% 15 1.6% B3 Plunging blades preserving a portion of the opposite striking platform, and 6 2.8% 4 1.1% 0 0.0% 1 2.9% 11 1.2% unidirectional or bidirectional scars B4 Plunging blades preserving a portion of the opposite cortical end, and unidirectional 2 0.9% 2 0.5% 2 0.7% 2 5.7% 8 0.9% or bidirectional scars Blades from the sides of the debitage surface 27 12.7% 47 12.4% 43 14.1% 2 5.7% 119 12.8% B5 Blades directly underlying a crested blade with symmetrical or asymmetrical section 1 0.5% 0 0.0% 0 0.0% 0 0.0% 1 0.1% and unidirectional or bidirectional scars B6 Blades with a lateral cortical edge (less than 50% of cortex) and unidirectional or 6 2.8% 11 2.9% 18 5.9% 0 0.0% 35 3.8% bidirectional scars B7 Blades with a cortical or natural steep back and unidirectional or bidirectional scars 9 4.2% 10 2.6% 19 6.3% 0 0.0% 38 4.1% B8 Blades with a lateral and distal cortical edge (less than 50% of cortex) 0 0.0% 0 0.0% 0 0.0% 0 0.0% 0 0.0% B9 Blades with centripetal dorsal scars on one side only 10 4.7% 22 5.8% 5 1.6% 2 5.7% 39 4.2% 0 B10 Blades with a cortical or natural steep back and distal cortical edge 0 0.0% 2 1 0.3% 0 0.0% 3 0.3% .5% B11 Plunging blades of type B4 a B6 (blades with a lateral cortical edge and plunging on 0 0.0% 0 0.0% 0 0.0% 0 0.0% 0 0.0% a cortical end) B12 Plunging blades of type B4 and B7 (blades with a cortical steep back and plunging 0 0.0% 1 0.3% 0 0.0% 0 0.0% 1 0.1% on a cortical end)

118

B13 Plunging blade of type B3 and B6 (blades with a lateral cortical edge and plunging 1 0.5% 1 0.3% 0 0.0% 0 0.0% 2 0.2% on a portion of the opposite striking platform) B14 Plunging blade of type B3 and B7 (blades with a cortical steep back and plunging on 0 0.0% 0 0.0% 0 0.0% 0 0.0% 0 0.0% a portion of the opposite striking platform) C: Core maintenance blades 0 0.0% 0 0.0% 0 0.0% 0 0.0% 0 0.0% C1 Crested blade of second generation (crest along the midline of the blade) 0 0.0% 0 0.0% 0 0.0% 0 0.0% 0 0.0% C2 Crested blade of second generation (crest in lateral position) 0 0.0% 0 0.0% 0 0.0% 0 0.0% 0 0.0%

D: Other 49 23.0% 87 23.0% 51 16.8% 5 14.3% 192 20.6% D1 Generic crested blades (that cannot be classed as first or second generation) 0 0.0% 0 0.0% 0 0.0% 0 0.0% 0 0.0% D2 Blades that fall outside any of the listed category 7 3.3% 3 0.8% 0 0.0% 0 0.0% 10 1.1%

E: Indeterminate blades 42 19.7% 84 22.2% 51 16.8% 5 14.3% 182 19.5% 213 379 304 35 931

119

Figure 16 Percent frequency of blades (n=972) with cortex vs. complete flakes and proximal flake fragments (n=1544) with cortex.

Débordant flakes/blades with unidirectional or bidirectional dorsal scars and a preserved lateral platform surface are also present in the assemblage (n=13, 1.3%, Figure 14a) and may have sometimes been used to rejuvenate core edges.

5.2.1.3 Core exploitation: directionality and recurrence

Blades were generally exploited from two opposed platforms. Among blades for which dorsal scar directionality could be confidently determined, there are more bidirectional blades (n= 146, 66.4%) than unidirectional blades (n=33.6%, 74). For blades from the optimal phase of debitage alone (B1, Figure 13), 80.3% are bidirectional. Blades from the optimal phase of debitage are defined as those without cortex, thin regular lateral edges, and with dorsal scars parallel to the long axis of percussion. Blades preserving a distal striking platform (Figure 12d) further attest to the bidirectional production of blades.

Blades from the optimal phase of debitage are also consistent with recurrent, i.e., serial blade removals. The exploitation surface is not re-prepared between every removal.

120

5.2.1.4 Platform and dorsal surface preparation

Platforms and dorsal surfaces of blades were highly prepared. Many platforms were faceted immediately prior to blade detachment (Figure 17a, 36.4% faceted platforms with bulbs). Dorsal surfaces often exhibit extensive preparation with small hinged removals or small more elongated dorsal scars (48.8%, Figure 17d).

5.2.1.5 Discarded blade core morphology

Two types of blade cores were identified (Table 15); rare “unorganized” blade cores with few blade removals from unprepared surfaces, and Levallois-like bifacial hierarchical blade cores with domed upper surface, steep lateral edges, and bidirectional flake scars (Figure 18).

Together, the two types of blade cores make up 14.6% of all identifiable cores types in the KP1 assemblage. The remainder and majority of discarded cores in the KP1 assemblage are flake cores. Flake production was likely an additional objective of core reduction at KP1 and the low frequency of blade cores might be a consequence of changing reduction strategies through the life history of the cores. Some cores that were initially used for blade extraction might have been exploited in a different way for flake production just prior to discard. This pattern is apparent on some of the discarded cores where the final removals are slightly below a length to width ratio of 2:1, or flake scars overlie and truncate the final blade removals (Figure 18).

The KP1 organized blade core reduction is related to Levallois as defined by Böeda (1995). The cores consist of two asymmetrical convex surfaces with roles that were irreversible and the flaking surface was maintained in a way that created lateral and distal convexities to guide the shock wave of each predetermined blank (Figure 19). Blade core reduction at KP1 diverges slightly from Boëda’s (1995) definition in that the upper exploitation surface of the core carries more volume than the lower platform surface. In other words, the upper surface is more convex than the lower surface, and the lateral convexities are steep and sit lower than the platform (Figure 18, Figure 19).

5.2.2 Flake production

Many flakes in the KP1 assemblage are by-products of blade production, since flake removals were used to prepare and maintain core surfaces and platforms. However, there is evidence that

121

Figure 17 Bulb and platform attributes of KP1 blades. a. Facetted platform (with negative bulbs of percussion), quadrangular morphology, dorsal side-up, b. Facetted platform (with negative bulbs of percussion), oval (restricted) morphology, dorsal side-up, c. Platform with a small lip and residual facets (no bulbs of percussion), ventral view, d. Dorsal surface preparation, dorsal view, e. Broken, curved delineation between bulb and platform, ventral side-up, f. Shattered bulb of percussion, ventral view.

flake production was also an intentional and distinct goal of core reduction based on the presence of (1) large Levallois flakes with radial dorsal scars, (2) Levallois cores with preferential and recurrent flake removals, (3) multiple types of non-Levallois cores with flake removals, and (4) retouched pieces made on flake blanks. Flakes and flake fragments make of 43.5 % of the entire lithic assemblage (Table 6), representing 73.0% of all discarded detached pieces (including blades and blade fragments).

Based on the diversity of core types, there were multiple strategies for flake production at KP1. Here, these strategies will be grouped into two general categories: Levallois reduction and non- Levallois reduction.

122

Table 15 Summary of core types, counts, and frequencies from KP1 stratum 4a, squares F23, F21, C23, and C21. Cores that are rolled or extensively weathered are excluded from these counts. F23 F21 C23 C21 Total Core Type Definition n % n % n % n % n % Multiplatform Cores cores with removals from 3 or more platforms, or two distant platforms 17 13.8% 25 15.3% 2 16.7% 10 9.8% 54 13.5% Unifacial Partial type 1 and 3 in Stout et al. (2010) - includes Exploitation Cores unifacial choppers, and minimally exploited pieces 10 8.1% 9 5.5% 1 8.3% 4 3.9% 24 6.0% Bifacial Partial type 4 in Stout et al. (2010) – includes Exploitation Cores bifacial choppers, and minimally exploited pieces 20 16.3% 13 8.0% 2 16.7% 4 3.9% 39 9.8% Large Surface hierarchical core with minimal preparation Exploitation and a single preferential removal 20 16.3% 34 20.9% 3 25.0% 24 23.5% 81 20.3% Unifacial Centripetal radial core with removals from one surface Cores only, removals not parallel to the plane of intersection 3 2.4% 6 3.7% 0 0.0% 4 3.9% 13 3.3% Bifacial Centripetal radial core with removals from both Cores surfaces, removals not parallel to the plane of intersection, non-hierarchical organization of the surfaces 2 1.6% 15 9.2% 0 0.0% 6 5.9% 23 5.8% Preferential Levallois bifacial hierarchical cores with straight Cores plane of intersection between two surfaces, scar of single "predetermined" large removal taken parallel to plane of intersection 8 6.5% 20 12.3% 0 0.0% 11 10.8% 39 9.8% Recurrent Levallois bifacial hierarchical cores with straight Cores plane of intersection between two surfaces, multiple "predetermined" scars taken parallel to plane of intersection 13 10.6% 19 11.7% 2 16.7% 14 13.7% 48 12.0% Blade Cores bifacial hierarchical cores with domed upper surface, intersection of surfaces is not a plane, elongated removals on exploitation surface, bidirectional exploitation 17 13.8% 15 9.2% 0 0.0% 18 17.6% 50 12.5% “Unorganized” Blade minimally prepared cores with a few Cores opportunistic blade or bladelet removals 2 1.6% 3 1.8% 0 0.0% 3 2.9% 8 2.0% Miscellaneous Cores cores that did not fit one of the above and categories Core Fragments 11 8.9% 4 2.5% 2 16.7% 4 3.9% 21 5.3% Total 123 163 12 102 400

123

Figure 18 Blade cores at KP1. All on banded ironstone. Arrows on sketches represented scar directionality. Circles indicated the presence of a negative bulb of percussion.

124

Figure 19 KP1 blade core organization. a. proposed geometry of bidirectional organized blade cores at KP1. b. platform view of idealized blade core, c. lateral view of idealized blade core.

5.2.2.1 Levallois reduction

Levallois cores make up 21.8% of the cores (Table 15, Figure 20) and probable Levallois flakes with prepared platforms and regular form make up 20.7% of the flakes (Table 16). A summary of Levallois flake metrics is provided in Table 17. The average length of complete Levallois flakes is 55.9 mm. The upper limits of flake length in this sample approaches 118 mm and the minimum length is 24.7 mm.

125

Figure 20 KP1 stratum 4a Levallois cores. a. unidirectional convergent exploitation, b. recurrent unidirectional exploitation, c. preferential exploitation with centripetal preparation, d. recurrent bidirectional exploitation, broken distal end, e. preferential exploitation, centripetal preparation, hinged final removal, f. recurrent centripetal exploitation, g. preferential exploitation.

126

Table 16 Summary of flake technical category frequencies for KP1 stratum 4a (n=972). Flakes that are rolled or extensively weathered or were indeterminable with respect to technical category are excluded from these counts. F23 F21 C21 C23 Total n % n % n % n % n % 1A Flakes with entirely cortical surface 39 13.1% 61 15.8% 35 13.3% 3 11.5% 138 14.2% 1B Flakes with more than 50% cortex on dorsal surface 4 1.3% 7 1.8% 6 2.3% 0 0.0% 17 1.7% 1C Flakes with lateral cortex 18 6.1% 29 7.5% 9 3.4% 2 7.7% 58 6.0% 1D Flakes with distal cortex 6 2.0% 16 4.1% 12 4.6% 0 0.0% 34 3.5% 1A Flakes with entirely cortical surface 11 3.7% 9 2.3% 8 3.0% 1 3.8% 29 3.0% 2: Non-Levallois flakes 180 60.6% 193 50.0% 94 35.7% 11 42.3% 478 49.2% 2A Flakes with multi-directional dorsal scars 65 21.9% 56 14.5% 31 11.8% 4 15.4% 156 16.0% 2B Flakes with unidirectional or bidirectional dorsal scars 52 17.5% 74 19.2% 31 11.8% 4 15.4% 161 16.6% 2C Flakes with convergent dorsal scars 8 2.7% 8 2.1% 6 2.3% 0 0.0% 22 2.3% 2D Flakes from side of débitage surface with steep lateral edges 20 6.7% 17 4.4% 11 4.2% 0 0.0% 48 4.9% 2E Flakes from side of débitage surface with steep lateral cortical edge 8 2.7% 24 6.2% 12 4.6% 0 0.0% 44 4.5% 2F Thick side-struck flakes 27 9.1% 14 3.6% 3 1.1% 3 11.5% 47 4.8% 3: Probable Levallois products 37 12.5% 77 19.9% 81 30.8% 6 23.1% 201 20.7% 3A Levallois flakes with prepared platforms and straight profile (from center of Levallois débitage surface, possibly preferential removals) 17 5.7% 39 10.1% 38 14.4% 5 19.2% 99 10.2% 3A1 with unidirectional/bidirectional scars (these differ from blades based on elongation and fragments based on having expansive sides, rather than parallel or convergent) 6 2.0% 16 4.1% 13 4.9% 3 11.5% 38 3.9% 3A2 with multidirectional scars 3 1.0% 8 2.1% 12 4.6% 1 3.8% 24 2.5% 3A3 with convergent scars 6 2.0% 15 3.9% 11 4.2% 0 0.0% 32 3.3% 3A4 with other or indeterminate scar pattern 2 0.7% 0 0.0% 2 0.8% 1 3.8% 5 0.5% 3B Levallois flakes with prepared platforms and curved profile (possibly core shaping and maintenance flakes, or flakes from recurrent cores) 9 3.0% 24 6.2% 26 9.9% 1 3.8% 60 6.2% 3B1 with mainly uni/bidirectional flake scars (may have some prep scars along one edge) 2 0.7% 8 2.1% 13 4.9% 0 0.0% 23 2.4% 3B2 with multidirectional scars 4 1.3% 13 3.4% 10 3.8% 1 3.8% 28 2.9% 3B3 with subradial scars (possibly reflects the scar of one preferential removal) 3 1.0% 1 0.3% 3 1.1% 0 0.0% 7 0.7% 3B4 other scar pattern or indeterminate 0 0.0% 2 0.5% 0 0.0% 0 0.0% 2 0.2% 3C Debordant flakes including lateral striking platform surface 9 3.0% 12 3.1% 17 6.5% 0 0.0% 38 3.9% 3C1 Debordant flakes with mainly uni/bidirectional flake scars 2 0.7% 7 1.8% 4 1.5% 0 0.0% 13 1.3% 3C2 Debordant flakes with multidirectional flake scars 6 2.0% 5 1.3% 10 3.8% 0 0.0% 21 2.2% 3C3 Debordant flakes with other or indeterminate flake scar 1 0.3% 0 0.0% 2 0.8% 0 0.0% 3 0.3% 3D Overshoot flakes from Levallois reduction including distal striking platform surface 2 0.7% 2 0.5% 1 0.4% 0 0.0% 5 0.5% Shaping and Thinning Flakes 41 13.8% 55 14.2% 53 20.2% 6 23.1% 155 15.9% 4A general (broad, thin and acute platforms) 33 11.1% 51 13.2% 45 17.1% 6 23.1% 135 13.9% 4B final (small, thin and acute platforms) 8 2.7% 4 1.0% 8 3.0% 0 0.0% 20 2.1%

127

Table 17 Summary of stratum 4a Levallois flake metrics for all Levallois flakes from squares F23 and F21 (n=113). Length (mm) Width (mm) Thickness (mm) L/W n=72 n=108 n=113 n=69 Mean 55.9 42.0 8.43 1.4 Min 24.7 17.0 3.4 0.6 Max 117.9 85.5 21.7 2.0 St. dev. 17.6 13.7 3.1 0.3

5.2.2.2 Core preparation

Early stage flakes make up 14.2% of the flakes (Table 16) and 2.3% (138/6052) of the entire lithic assemblage. The presence of these early stage flakes with cortex indicates that raw material was transported to the site mainly as unprepared nodules.

Dorsal surfaces of Levallois cores seem to have been prepared using a variety of strategies, including unidirectional, bidirectional, radial, and less often, convergent, removals. This argument is made based on the various dorsal scar patterns on Levallois flakes (Table 16) and the scar patterns on the upper surface of Levallois cores.

The upper surfaces of Levallois cores at KP1 were prepared with appropriate distal and lateral convexities for ‘predetermined’ removals. Lower surfaces were usually heavily prepared so that they could serve as striking platforms.

5.2.2.3 Core maintenance

Débordant flakes with various dorsal scar patterns and a preserved lateral platform surface make up 3.9% of the flakes (Figure 14a) and may have been used to rejuvenate Levallois core edges

5.2.2.4 Core exploitation: directionality and recurrence

There is evidence for both preferential and recurrent Levallois reduction, based on core and Levallois flake dorsal scar patterns. Of the Levallois cores, 44.8% (39/87) exhibit a single large preferential removal parallel to the plane of intersection on the core surface, and 55.2% (48/87) exhibit multiple large removals parallel to the plane of intersection. Some cores were definitely not prepared between successive removals, whereas others were discarded after a single large removal took off the majority of the upper surface. The presence of Levallois flakes with straight

128 profile and multidirectional dorsal scars (10.2% of flakes) further suggests that preferential Levallois reduction occurred, which requires re-preparation of core surfaces in between each removal. The presence of subradial Levallois flakes (0.9% of flakes) with large dorsal scars from previous predetermined removals (Boëda, 1995; Van Peer, 1992) is consistent with recurrent exploitation.

Recurrent Levallois reduction either occurred unidirectionally or centripetally, based on the presence of Levallois cores and flakes with these dorsal scar patterns.

5.2.2.5 Platform and dorsal surface preparation

Platforms and dorsal surfaces of Levallois flakes were highly prepared. Most platforms were faceted immediately prior to flake detachment (54.1% faceted platforms with bulbs). Dorsal surfaces usually exhibit extensive preparation with small hinged removals or small more elongated dorsal scars (62.1%).

5.2.2.6 Discarded Levallois core morphology

Some core reductions at KP1 conform to the Levallois Volumetric Concept as defined by Böeda (1995). Levallois cores consist of two asymmetrical convex surfaces with roles that are irreversible and the flaking surface is maintained in a way that created lateral and distal convexities to guide the shock wave of each predetermined blank. The two surfaces meet at a plane of intersection, and extractions from the upper surface, which carries less volume and is less convex than the lower surface, are removed parallel to the plane of intersection.

It was already mentioned above that some cores that were initially used for blade extraction might have been exploited in a different way for flake production just prior to discard. However, Levallois flake production and blade production seem to have had mainly separate trajectories. Some blade cores may have been transformed into Levallois flake cores later in the reduction sequence, but if this occurred, it did not occur regularity. If blade cores were regularly transformed into Levallois flake cores, we would expect significant differences in size between the two core types, with Levallois cores exhibiting smaller maximum dimensions than blade cores. Levallois cores do not have significantly smaller maximum dimensions than blade cores (Figure 21, unequal variance t-test, t=1.141, p=0.265).

129

Figure 21 Box plot of core size for KP1 stratum 4a assemblage.

5.2.3 Non-Levallois reduction

The majority (58.7%) of discarded cores at KP1 were used for flake removals and were not organized by the Levallois Volumetric Concept (Table 15). There are a variety of core types in this category, including cores with multiple platforms, minimal cores with just a few removals, and centripetally exploited, or radial cores.

Some of these cores could represent blade and Levallois cores at the end of their life history, with the final extractions removing evidence for Levallois or blade reduction strategies. As a group, the non-Levallois flake cores are significantly smaller (Figure 21) than blade (t=3.975, p<0.001) and Levallois cores (t=3.134, p=0.003).

5.2.4 Retouched pieces

Retouched pieces represent 12.1% of the analyzed KP1 stratum 4a sample (Table 12). There are a variety of categories represented (Table 18).

130

Table 18 Summary of retouch pieces types, counts, and frequencies from KP1 stratum 4a squares F23, F21, C21, C23. Pieces with edge damage of unknown origin (rolled or heavily weathered) are not included in these counts.

F23 F21 C21 C23 Total

n % n % n % n % n % Retouched pointed converging edges shaped into a forms (including point (retouch is almost always 42 23.9% 51 22.9% 55 28.4% 1 5.6% 149 24.4% fragments) unifacial, can be along one or both lateral edges) Denticulates and retouch must occur inside the notches notches to qualify, or denticulation 36 20.5% 42 18.8% 36 18.6% 5 27.8% 119 19.5% must be exaggerated and regular Side continuous retouch along a portion 27 15.3% 36 16.1% 32 16.5% 2 11.1% 97 15.9% one lateral edge End scraper continuous retouch that includes 16 9.1% 15 6.7% 19 9.8% 1 5.6% 51 8.3% distal end Double scraper continuous retouch along two edges 16 9.1% 34 15.2% 9 4.6% 3 16.7% 62 10.1% Scrapers with three or - 12 6.8% 11 4.9% 11 5.7% 0 0.0% 34 5.6% more edges retouched Minimally retouched retouch is limited or edge damage 12 6.8% 16 7.2% 10 5.2% 0 0.0% 38 6.2% may have resulted from use Miscellaneous retouched pieces that do not fit any 8 4.5% 6 2.7% 11 5.7% 2 11.1% 27 4.4% of the above categories Broken retouched fragment of a retouched piece that pieces cannot be assigned to above 7 4.0% 12 5.4% 11 5.7% 4 22.2% 34 5.6% category Total 176 223 194 18 611

Both blades and flakes served as blanks for retouched pieces, but blades may have been preferentially selected. Table 19 presents a comparison of retouched types on flake and blade blanks in square F23. All categories of retouched pieces were made on both flake and blade blanks, but there is a pattern indicative of preferential selection for blades. Blades are overrepresented in the retouched pieces component of the assemblage compared to the debitage. There is a 1 to 1 flake to blade ratio within the retouched pieces for which the blank could be determined (Table 19), even though there are 2.7 to 1 flake to blade ratio in the debitage. The difference between the flake and blade counts in the debitage and in the retouched pieces is statistically significant (Fisher’s exact test, p<0.01).

Unifacially worked pointed forms (Figure 12h,j,k) are the most striking category and make up 33.3% of the retouched pieces (Table 18). Unifacially retouched pointed forms exhibit the strongest pattern of selection for blade blanks. Of the 31 pointed forms for which the blank could be determined, more were manufactured on blades (n=19) than on flakes (n=12, Table 19). Most of the pointed forms that were manufactured on blades were manufactured on B1 blades,

131

Table 19 Percent frequency of retouched pieces manufactured on flake, blade, and B1 blade blanks in square F23. Only retouched pieces for which the blank could be confidently determined are included in those columns.

On Flake Blanks On Blade Blanks On B1 Blanks n % n % n %

Retouched pointed forms (including fragments) 12 20.0% 19 31.6% 16 33.3%

Denticulates and notches 15 25.0% 12 20.0% 10 20.8%

Side scraper 8 13.3% 11 18.3% 8 16.7%

End scraper 5 8.3% 5 8.3% 4 8.3%

Double scraper 8 13.3% 4 6.7% 3 6.3%

Scrapers with three or more edges retouched 4 6.7% 2 3.3% 2 4.2%

Minimally retouched 5 8.3% 4 6.7% 2 4.2%

Miscellaneous 3 5.0% 3 5.0% 3 6.3%

Broken retouched pieces 0 0.0% 0 0.0% 0 0.0%

Total 60 60 48

produced during optimal phase of debitage. A functional analysis of these points is presented in Chapter 5.

5.2.5 Percussion technique

A bulb and platform analysis (Soriano et al., 2007; Villa et al., 2010), was used to determine the percussion technique for blade and flake production. Table 20 summarizes the results and Figure 17 provides examples of some of the features. The high frequencies of large prominent bulbs, thick platforms, shattered bulbs (Figure 17f), and platforms with a broken delineation (Figure 17e) are consistent with direct hard hammer percussion that is away from the core edge (i.e. internal, not marginal). These observations contrast with those observed for the levels attributed to the Howiesons Poort (a subphase of the MSA, 65-60 ka, Jacobs et al. 2008) at Rose Cottage Cave and Klasies River Mouth where the authors argue that marginal percussion with a soft stone hammer was the technique used for blade production (Table 20).

An experimental sample (n=147) of debitage from hard hammer percussion of banded ironstone cores local to the Kathu region was also subjected to the same bulb and platform analysis. This experimental sample yielded trait frequencies similar to the KP1 blades, demonstrating higher frequencies of thick and wide quadrangular platforms and shattered bulbs, but very low frequencies of lipped platforms without a bulb (Table 20). An interesting observation is that

132

Table 20 Percent frequencies of bulb and platform attributes at KP1 (square F23, stratum 4a), compared to an experimental hard hammer percussion sample (n=147), and Howiesons Poort assemblages at Rose Cottage Cave (RCC, Soriano et al., 2007) and Klasies River (KRM, Villa et al., 2010) where direct soft stone hammer percussion may have been used. Traits that exhibit divergence between KP1 and the Howiesons Poort, but similarity between KP1 and the experimental hard hammer sample are bolded. For details about the different stratigraphic units consult Soriano et al. (2007) and Villa et al. (2010). A list of all observed attributes is provided in Table 11. Soft Stone Hammer KP1 KP1 Experimental RCC RCC RCC RCC KRM KRM KRM Blades Flakes Hard Hammer EMD MAS HP HP HP n=129 n=116 N=147 ETH+SUZ BYR+THO Lower Middle Upper Dorsal preparation (trimming of the edge on the 48.8 53.4 54.8 87.7 85.3 62.7 42.9 58.8 56.5 16.2 exterior core surface) Abrasion of the edge 14.7 32.8 19.2 65.3 42.3 26.9 11.0 24.6 24.4 8.1 Faceted platforms (with 36.4 31.0 17.1 5.5 2.0 20.0 24.7 18.4 15.9 46.2 bulbs) Platform<2mmthick 14.7 2.6 2.1 80.3 78.8 60.2 60.9 38.9 51.2 15.4 Platform>5mmthick 82.9 62.9 76.0 0.0 1.0 3.5 6.5 12.1 4.7 28.2 Platforms<3mmwide 0.7 0.0 0.0 52.0 37.8 27.6 17.0 4.4 17.1 2.6 Impact near the lateral edge 23.3 37.9 32.2 16.3 16.7 15.3 10.4 of the platform Platforms with a visible 73.6 79.3 89.0 67.5 72.9 78.4 81.3 impact point Straight or curved platform delineation (not 28.7 13.8 14.4 62.3 51.6 39.4 33.7 49.7 45.5 23.7 overhanging) Overhanging curved 28.7 16.4 20.5 10.1 18.3 18.3 21.3 14.1 15.9 23.7 platform Overhanging with bulb in clear relief (broken 12.4 40.5 32.9 8.7 3.2 21.2 29.2 11.9 13.6 31.6 delineation) Oval or narrow triangular 32.6 34.5 34.9 54.4 43.4 37.4 28.3 36.2 38.6 28.2 platform Quadrangular or wide 45.7 37.9 43.2 0.0 5.1 23.4 27.2 28.1 22.7 33.3 trapezoidal platform Lipped without a bulb 0.0 1.7 0.7 8.5 8.6 4.4 7.8 18.1 6.8 5.6 Lipped with or without a 32.0 38.8 42.5 bulb Prominent bulb with or 80.6 63.8 56.2 47.6 47.3 53.5 55.6 29.0 52.3 52.8 without lipping Weakly developed bulb with 15.5 32.8 41.8 42.7 41.9 40.4 35.6 47.1 34.1 33.3 or without lipping Platform with impact point 2.3 5.1 0.7 0.0 0.0 0.0 1.1 0.0 0.0 0.0 contoured by a fissure Platform with partly 7.8 3.4 6.2 6.0 16.5 12.9 18.3 27.0 25.6 13.5 contoured impact point Platform with contoured 6.2 5.2 10.3 1.2 3.1 2.6 0.0 0.5 4.7 0.0 Hertzian cone Platform with shattered 32.6 56.9 42.5 14.3 8.0 9.4 11.5 5.6 13.3 10.0 bulb

133

>30% of platforms on KP1 blades and flakes were lipped or had a slight lip (Figure 17c), which is a trait generally associated with soft hammer or indirect percussion for MSA assemblages (e.g., Wurz, 2000, 2002). Of the hard hammer experimental debitage sample, a large proportion were also lipped or had a slight lip, consistent with the interpretation that direct hard hammer percussion was used to detach blades at KP1.

5.2.6 LCTs

LCTs are 0.3% of the stratum 4a assemblage (Table 12). These LCTs are variable in size and shape, and exhibit large invasive flake removals with large negative bulbs of percussion. They are generally not finely-shaped (Porat et al., 2010). When blank could be determined, LCTs were manufactured on nodules, not flakes. There is no technological link between the LCTs and the rest of the stratum 4a assemblage, because the LCTs do not appear to be manufactured on any of the byproducts of blade or Levallois reduction. In some later Acheulean assemblages, LCTs are manufactured on large flakes removed from “giant cores” (Sharon, 2009), but that is not the case for the KP1 handaxes. Interestingly, the original definition of the Fauresmith included handaxes made on flakes (Goodwin and Van Riet Lowe, 1929:72-85), and in that respect KP1 does not conform to the original definition of the Fauresmith. Side-struck flakes (technical category 2F) and shaping and thinning flakes (technical category 4) in the KP1 assemblage are consistent with LCT manufacture, but not diagnostic of it. Those types of flakes could also be the by-products of blade and Levallois core preparation. In contrast, other aspects of the KP1 assemblage are technologically linked; retouched pointed forms are manufactured on both blade and Levallois flake blanks, for example.

Table 21 presents a summary of KP1 LCT metrics. Fauresmith handaxes are described as small relative to Acheulean handaxes (Goodwin and Van Riet Lowe, 1929). However, a comparison of strata 4a and 4b handaxes at KP1 reveals that the mean length differs only by 12 mm, and the difference is not significantly different (T-test, t=1.344, p=0.190). Variance is also not significantly different (F-test, F= 1.455, p=0.477). KP1 LCTs from both stratum 4a and 4b have a mean length shorter than the mean length of 108 mm reported for Fauresmith handaxes by Goodwin and Van Riet Lowe (1929).

134

Table 21 Summary of LCT metrics. All LCTs from stratum 4a of four square units, F23, F21, C23, C21 (n=20), and two levels of stratum 4b (F21, levels 160-170, 170-180, n=14) Stratum 4a Stratum 4b Length Width (mm) Thickness Length Width Thickness (mm) (mm) (mm) (mm) (mm) n 18 19 19 11 11 11 Mean 75.3 47.3 23.9 87.1 54.9 25.1 Min 51.2 31.7 14.4 57.6 34.9 18.7 Max 137.2 71.8 43.7 132.6 80.9 35.8 St. dev. 21.2 10.9 8.6 25.6 16.1 5.4

5.2.7 Vertical distribution of artifact classes

All aspects of the 4a assemblage are distributed relatively evenly throughout the stratum 4a levels in square F23, except for LCTs, which only occur in the lower levels (Figure 22). All LCTs in the four sampled squares come from depths greater than 100 cm, and in the only the lowermost levels of stratum 4a, with the exception of a quartzite LCT that could be alternatively interpreted as a core. Because KP1 was excavated by large arbitrary spits (10 cm for F23, 20 cm for some other squares), the association of handaxes with the rest of the KP1 assemblage could be the result of excavation procedures. As discussed in Chapter 4, the 4a LCTs are also more weathered than the rest of the 4a assemblage, suggesting that they may have been subjected to longer surface exposure or higher energy transport. Several lines of evidence suggest that non- behavioral processes explain the association of the LCTs with the blades, Levallois products, and points of stratum 4a. Further excavations that focus on site formation processes could address the relationship of the bifaces to the remainder of the assemblage from stratum 4a.

5.3 Discussion

The stratum 4a lithic assemblage from KP1 represents a blade- and flake-based industry that can be summarized as follows:

 Diverse strategies that included Levallois and non-Levallois methods were employed for blank production.

135

Figure 22 Vertical distribution of artifact types within stratum 4a, square F23 (n=1936). Dating samples (see Chapter 4) were recovered from ~80-110 cm below the top of stratum 3.

 Blades were systematically manufactured on bifacial hierarchical cores with steep lateral edges that were prepared and maintained with flake removals.

 Nodules of raw material were transported to the site as complete or minimally worked cores.

 Sometimes core surfaces and platforms were extensively prepared for the production of pre-determined blanks, sometimes they were not.

 Direct hard hammer percussion was the technique used.

136

 Both flake and blade blanks were selected for retouch into a variety of forms, the most prominent form being the retouched points.

 LCTs are rare in the assemblage and could possibly be intrusive.

Thus, based on the emphasis on flakes and blades from prepared cores, the high frequency of both convergent and retouched points, and the rarity of LCTs that are not technologically linked to the remainder of the assemblage, I hold that the KP1 stratum 4a assemblage is more consistent with an MSA designation than an ESA designation.

5.3.1 KP1 blade production in context

KP1 is one of only two African localities with blade production dated to the early Middle Pleistocene. In East Africa, the earliest occurrences of blades and blade cores date to 500-550 ka in the Kapthurin Formation (Johnson and McBrearty, 2010). Blade production at KP1 is distinct from blade production at the Kapthurin sites (Table 22). The Kapthurin blade cores are generally pyramidal or flat and blades generally follow the long edge of one preparatory flake scar with little to no core preparation. In contrast, blade cores are extensively prepared at KP1. Johnson and McBrearty (2010:195) describe the Kapthurin blade reduction strategy as similar to the Hummal volumetric concept (c.f., Boëda, 1995), but KP1 blade production is clearly related to Levallois reduction strategies. Blades are much less frequent at the Kapthurin Formation sites, which exhibit a flake to blade ratio of 36.6 to 1, compared to 2.7 to 1 at KP1. The Kapthurin sites may demonstrate a less regular emphasis on blade production than KP1, though the low frequency of blades at the Kapthurin sites could in part be related to the ephemeral, open-air occupations that the sites represent.

The only other early Middle Pleistocene site with laminar technology is , Israel, where the earliest occupation has been dated to 320±30 ka (Gopher et al., 2010). KP1 blade production is also distinct from Qesem Cave blade production (Table 22). At Qesem Cave, blades are detached from minimally prepared flat nodules, reduction starts along one nodule edge, and core maintenance is minimal (Shimelmitz et al., 2011). When preparation or maintenance was required, edges were often crested, resulting in a fairly high frequency of crested blades (n=199, 2.2% of the debitage and shaped items, Tables 1 and 9 in Shimelmitz,

137

Table 22 Comparison of KP1 blade production to published results for the Kapthurin Formation (Johnson and McBrearty, 2010) and the Amudian industry as represented at Qesem Cave, Israel (Barkai et al., 2009; Shimelmitz, 2009; Shimelmitz et al. 2011). Kapthurin Formation Sites KP1 GnJh-42 and GnJh-50 Qesem Cave South Africa Kenya, East Africa Israel, Age 464 ±47 ka <545±3 ka 320±30 ka 542+140/-107 ka >509±9 ka 300±10 ka > 300±30 ka Dating Method OSL/ESR 40Ar/30Ar 230Th/234U

Minimum- 682-435 ka 548-500 ka Earliest occupation 290-480 ka maximum range Industrial Fauresmith Acheulean Amudian designation Context/site type Open-air, pan Open-air Cave Analyzed sample 6052 862 19,167 size Raw material Mainly large cobbles of banded Mainly fine-grained phonolite, small Small flat nodules, 1-5 km ironstone/chert, local sources mainly < cobbles, local sources 1-5 km away away 4 km away Core organization Levallois-like with steep lateral "Hummal-like”, mainly unidirectional Minimally prepared flat nodules, convexities, intersection of surface not a or "centripetal, one core is reduction starts along nodule plane, uni or bidirectional bidirectional, initation follows long edge, core maintenance minimal edge of flake scar Platform faceting Faceting common Faceting rare Faceting rare Technique Direct hard hammer Direct hard hammer Direct hard hammer Marginal or Internal Internal Internal internal percussion Flake:blade ratio1 2.7:1 36.6:1 1.6:1 Mean length (mm) 69.7 (n=113, sd=19.3) 50 (n=4, sd=17.4)* Blade 51.2 (n= 433, sd=12.7) PE blade 53.7 (n=385, sd=12.0) NBK 52.5 (n=418, sd=10.9) Mean width (mm) 28.2 (n=511, sd=9.2) 20.2 mm (n=15, sd=6.2)* Blade 20.9 (n= 429, sd=5.5) PE blade 21.5 (n=385,sd=5.6) NBK 20.8 (n=420, sd=5.3) Mean length to 2.5:1 (n=92, sd=0.4) 2.7:1 (n=4, sd=0.05)* Blade 2.5:1 (n= 422, sd=0.4) width ratio PE blade 2.6:1 (n=380,sd=0.5) NBK 2.6:1 (n=415, sd=0.5) Retouched pieces Unifacial pointed forms, None Laterally and distally retouched with blades as denticulates/notches, scrapers blades, no points blanks

1(flakes+flake fragments):(blades+blade fragments). *based on ”maximum length” for complete blades and “maximum breadth” in Johnson and McBrearty 2010 supplementary information

2009). Core edges were not crested at KP1, with only 3 pieces classified as first generation crested blades (Table 14).

Though KP1 blade production exhibits some similarities with later MSA strategies (see Table 4, Section 3.3.2), the combination of traits characteristic of KP1 blade production is not currently known from other African sites. At KP1, blade core organization is Levallois-like, with two hierarchical surfaces, an upper surface with steep lateral edges, core preparation with centripetal removals, and the upper surfaces are exploited recurrently. In these respects, KP1 core reduction

138 is similar to that observed for the HP levels at Klasies River (Villa et al., 2010). However, there are important differences between the KP1 and HP at Klasies River. First, the blades at KP1 are much larger. Second, they are exploited bidirectionally from core surfaces, rather than unidirectionally. Third, they are detached using direct hard hammer percussion, not soft stone hammer as suggested for the Klasies River HP assemblage.

There are similarities between KP1 blade production and two refitting core sequences from Taramsa; Taramsa 112 Refit 6 and Refit 3 (Van Peer, 1992:101-112). These refitting cores produced laminar endproducts using longitudinal guiding ridges produced from two opposing platforms, and are described as similar to, but departing from the Levallois concept. Van Peer (1992) argues that the two Taramsa core reductions differ from a strict definition of Levallois in three ways. First, the upper surfaces have more exploitable volume than the lower surface and more pronounced lateral convexities. The purpose of the more pronounced lateral convexities is to limit width of the endproducts without sacrificing too much thickness. Second, the distal platform is plays an integral role and is modified throughout the reduction sequence. Detached pieces from these distal platforms are elongated and have prepared butts and may have been endproducts themselves, or served to create longitudinal guiding ridges on the upper surface. Third, some endproducts are removed more tangentially with respect to the plane of intersection in Refit 3.

KP1 core reduction strategies and the Taramsa refitting sequences are not identical. Both the Taramsa refitting sequences eventually bring down the upper volume so that it does become less than lower volume before discard. The Taramsa discarded cores in most respects look like Levallois cores. They have more acute upper surface convexities compared to the lower surface convexities. In contrast, the KP1 discarded blade cores have more volume on the upper surface and less acute upper surface convexities compared to the lower surface convexities. It is not that KP1 blade cores are discarded at earlier stages of reduction than the Taramsa cores. There are very small KP1 blade cores that exhibit the blade core geometry with steep lateral edges and large upper surface volume, suggesting that the KP1 blade core concept is maintained through the reduction sequence. Furthermore, as discussed above, KP1 Levallois flake cores are not smaller than KP1 blade cores.

139

KP1 blade production adds to the known variants of laminar technology in the early Upper and Middle Pleistocene. This pattern of diversity in MSA blade production strategies seems to contrast with the Upper Paleolithic record of Europe, where prismatic blade production is ubiquitous across Europe for ~20-30 thousand years (Bar-Yosef and Kuhn, 1999). Differences between KP1 and the Kapthurin sites indicate that diversity was probably present at the onset of blade producing behaviors ~500 ka and blade production was invented multiple times in multiple places throughout the Middle and Upper Pleistocene (see also Wilkins and Chazan, 2012).

The meaning of the diversity in blade production behaviors prior to the Upper Paleolithic is a subject worth further consideration. Blade production prior to the Upper Paleolithic has been described as temporally and spatially patchy. What was unique about the Upper Paleolithic according to (Bar-Yosef and Kuhn, 1999), were the social conditions, such as large population sizes and extensive social networks, that favored the widespread adoption and use of blade technology across large geographical space and for a long period of time. Whether or not this pattern is lacking prior to the Upper Paleolithic should remain a matter of debate, however. At this point, the archaeological record of the Middle and early Upper Pleistocene is patchy, but that does not necessarily mean that the behaviors were patchy. Archaeological visibility, dating limitations, and history of research are all factors that contribute to the particular slices of time in the particular regions of Africa that we can access. There is some evidence that blade-producing behaviors persisted for long periods in the early Middle Pleistocene. Blades have also been recovered from more recent contexts , as young as 285 ka, in the Kapthurin Formation (Deino and McBrearty, 2002), and at Qesem Cave the Amudian Industry is present in multiple strata and persisted at the site until at least 220 ka (Gopher et al., 2010). Future research might elucidate the significance and chronological limits of Fauresmith-designated assemblages in southern Africa, but at this point is impossible to evaluate the duration of the so-called Fauresmith Industry (see Section 3.1.2).

5.3.2 Diverse core reduction strategies at KP1

The KP1 assemblage also documents technological variability at a local scale. Above I presented evidence that I think supports the existence of multiple core reduction strategies within the KP1 assemblage. Blade production was one goal at KP1, and there is good evidence that the same method was used for multiple blade core reductions. However, flake production was an

140 additional goal of core reduction at KP1, and there were strategies employed independent from blade production for the purpose of manufacturing flakes. At KP1, both Levallois and non- Levallois methods were used for flake production based on the presence of Levallois and non- Levallois cores, which is an observation common for MSA assemblages. For example, an analysis of the MSA levels from Porc-Epic Cave, Ethiopia emphasizes the contemporaneity of four main schema opératoires (Pleurdeau, 2006). Pleurdeau (2006) suggests this indicates that the MSA toolmakers at the site were flexible to the demands of flake production and capable of employing various and novel behaviors.

Even within the Levallois category of core reduction, there were multiple strategies employed at KP1, including preferential, recurrent unidirectional, and recurrent centripetal. This is also consistent with other late ESA and MSA sites; a variety of exploitation methods was employed during the late ESA and MSA occupations of Kudu Koppie (Wilkins et al., 2010). During the ESA occupation there, both the lineal and recurrent methods were employed, though the lineal method of prepared core exploitation clearly dominates the assemblage. A greater variety of exploitation methods was recognized in the MSA levels of Kudu Koppie. The lineal method was still the most prevalent, but recurrent centripetal, recurrent unidirectional and recurrent bidirectional cores also formed a significant component of the assemblage. There is some evidence at other sites for the diversification of core reduction strategies through time across the ESA and MSA boundary. At Koimilot, Kapthurin Formation, Kenya the MSA is characterized by the diversification of prepared core methods compared to the ESA (Tryon, 2006; Tryon and McBrearty, 2006; Tryon et al., 2005).

A problem with any Pleistocene archaeological assemblage is the potential for palimpsests. While the lithic evidence for these multiple reduction strategies were all buried at the same time ~500 ka, they could represent non-concurrent behaviors; multiple visits to the same locality for the same or different purposes across time. However, the observation that retouched points and other retouched tools were manufactured on both blade blanks and Levallois flake blanks is strong support that the multiple strategies were part of a shared technological repertoire of a group or groups of Middle Pleistocene hominins.

141

5.3.3 Summary

The stratum 4a assemblage at KP1 represents a mainly flake and blade-based industry that employed multiple strategies to produce blanks that were retouched into a variety of forms, including unifacial points. Diversity at KP1 suggests that there was flexibility in the way in which flakes and blades were manufactured. That both types of blanks are selected for each category of retouched tool implies that multiple strategies were used to obtain the same goal. Diverse core reduction strategies might indicate that KP1 hominins were flexible to the specific demands of local raw materials and capable of employing new behaviors. While in some cases they followed social convention, using the same method to obtain the same goal, in other cases, they used different methods to obtain the same goal. This is the kind of technological behavior that would result in invention and the cultural transmission of novel behaviors between individuals and through time. The theoretical link between diversity and innovation and implications for human evolution are discussed in Chapter 8.

142

6 Functional Analysis of Stratum 4a KP1 Points

While sorting the retouched pieces into typological categories, it was realized that retouched points (Figure 23) were the most striking and one of the most abundant categories of tools. These points are usually retouched unifacially on the dorsal surface of both lateral edges. Retouch sometimes is distributed across the entire lateral edge, but more often it is concentrated near the distal end. If there is retouch in the ventral side, it is minimal. In all respects, these points are similar to those recovered from typical MSA and MP assemblages. As discussed in Chapter 3, many MSA and MP points show evidence that they were hafted and used as spear tips. However, some studies of MSA and MP points also highlight their role as multi-purpose tools involved in cutting, scraping, and/or piercing tasks, demonstrating that it is invalid to assume point function based on morphology alone. Furthermore, post-depositional processes are known to mimic some traces on stone tools linked to weapon use, so it is necessary to rule out non-use related wear when determining tool function.

KP1 provides the earliest chronometrically dated assemblage of points and as such, could provide some of the earliest evidence for hafted hunting technology. For reasons detailed in Chapters 2 and 3, the origin of hafted hunting technology has evolutionary implications. For example, the attachment of stone tools to handles may be linked to changes in cognition and social learning, and stone-tipped armatures may have increased hominin foraging returns, with important implications for Homo life history, such as decreased child and adult mortality, and human sociality.

In this chapter, I present the results of a study that employed multiple methods to test the hypothesis that the KP1 points were used as spear tips. Some of this research has been published in Science (Wilkins et al., 2012), and is reprinted here with permission from AAAS. I start by presenting a description of each of four methods used to assess point function, where I also lay out the expectations for the assemblage of KP1 points if points were used by the KP1 hominins to tip hunting weapons. I follow with the results of each of these four methods when applied to the KP1 point assemblage. I conclude with an interpretation of how well the results fit with the expectations for hafted hunting weapons and discuss methodological considerations. In Chapter 8, I discuss the evolutionary implications for the evidence of hafted hunting technology at KP1.

143

6.1 Methods

All points (n=210, Table 23), including retouched points (Figure 23), convergent flakes and blades, irregular pointed forms and fragments (Figure 24) from the analyzed stratum 4a sample were considered in the functional analysis presented here. Preliminary observations indicated that some points exhibited diagnostic impact fractures and basal modifications, and the large size of the sample was suitable for edge damage distribution analysis and quantitative considerations. Both retouched and non-retouched points were considered because both types exhibited diagnostic impact fractures. Here, I will report the results of four aspects of the functional study: A. edge damage distribution analysis, which was a collaborative effort with Benjamin J. Schoville (Arizona State University) and Kyle S. Brown (University of ) and includes an experimental component7 B. diagnostic impact fractures C. basal modification D. point metrics

Of the six types of evidence used by other researchers to determine MSA point function -- microscopic use wear, macroscopic use wear, edge damage distribution, residue, retouch and shaping, and geometric morphometrics (Table 5, Section 3.4.2) -- four were applied to the KP1 points to establish whether any or all of them may have been used as tips. Three of those are presented in this chapter. Point metrics were further used to compare the KP1 points to MSA points and consider the feasibility of their use as spear tips. Traditional microscopic use- wear analysis (e.g., Rots, 2003; Rots, in press; Rots et al., 2011) was not conducted. Patination resulting from the chemical alteration of lithic surfaces can obscure microscopic traces of use such as polishing and striations (Keeley, 1980; Levi Sala, 1986) and most of the KP1 points are on patinated banded ironstone and not suitable for traditional microwear analyses, though there may be potential for microscopic use wear analyses on some of the unpatinated chert points in the future. Residue analysis is unlikely to yield fruitful results due to both the antiquity of the assemblage and its context within an artesian system with a rising and falling water table.

7 Contributions: JW: data collection and analysis of KP1 and experimental points, methodological design , data interpretation, participation in experiments; BJS: methodological design, data interpretation, design and implementation of experiments; KSB: methodological design, design and implementation of experiments

144

Figure 23 Examples of KP1 complete retouched points. All banded ironstone except A, G, M, O, T, W, AD (black chert), C and V (quartzite).

145

Figure 24 Examples of other KP1 point types. A, D) non-retouched convergent flakes, B, C, E) non-retouched convergent blades, F-G) retouched point tips, H-J) retouched point bases, K-L) irregular retouched pointed form. All on banded ironstone except for A (quartzite) and G (black chert). (from Wilkins et al., 2012, supplementary material).

Table 23 Point sample for functional analysis. All retouched points (including fragments), and non-retouched convergent flakes and blades (including fragments) from four square units of stratum 4a. Square Depth below Retouched Retouched Retouched Retouched Not- Not- Irregular Total top of Point Point Tip Point Point Base retouched retouched Retouched stratum 3 Fragment Convergent Convergent Pointed (cm) Flake Blade Form 40-60 5 2 1 2 1 11 60-80 7 1 1 5 2 1 17 C21 80-100 12 1 1 1 4 3 22 100-120 7 3 1 4 3 6 24 C23 60-80 1 1 40-60 2 5 1 3 2 1 3 17 60-70 2 1 2 2 3 10 70-80 1 1 80-90 4 1 5 90-100 3 1 1 3 2 10 F21 100-110 2 1 2 5 110-120 2 1 1 2 2 3 11 120-130 3 3 2 3 1 12 130-140 2 1 2 1 6 140-150 1 1 60-70 2 1 2 1 2 4 12 70-80 5 3 1 3 1 3 1 17 80-90 1 2 1 2 6 F23 90-100 4 1 1 1 2 1 10 100-110 4 2 6 110-120 1 2 1 4 120-130 1 1 2 Total 69 22 4 25 28 35 27 210

146

Geometric morphometrics was also used to as assess the function of KP1 points (Wilkins et al., 2012), but that analysis was conducted primarily by Benjamin J Schoville, so the results are not presented in this dissertation chapter. While it is possible that the points may have served multiple purposes as has been observed for some MP assemblages (e.g., Beyries and Plisson, 1998), addressing that question is not one of the goals of the present research. The focus of this study was to determine whether the manufacture of hafted hunting weapons was part of the technological repertoire of KP1 hominins ~500 ka, and given the nature of the KP1 assemblage, the multiple methods employed here were the best for addressing this question.

6.1.1 Edge damage distribution

The first method used to determine the function of the KP1 points quantifies the distribution of fractures along the lateral edges of the point (Bird et al., 2007; Schoville, 2010; Schoville and Brown, 2010). There are some advantages to this type of analysis over other types of functional analyses. First, the distribution of fractures along tool edges is a quantifiable trait that lends itself to objective statistical tests. Second, edge damage distribution analyses makes statements about the distribution of fractures at the level of the assemblage, rather than the individual artifact, providing increased analytical power for quantitative analysis. Third, the method does not require attributing causation to individual fractures, which can be subjective (Shea and Klenck, 1993).

Multiple processes can result in minor damage and scarring along the edges of lithic tools. Small fractures can form on a tool edge when it is utilized, but post-depositional processes such as trampling and agitation of a flake in water and/or sediments can also fracture tool edges (Tringham et al., 1974). Some use wear analysts maintain that it is possible to distinguish post- depositional fractures from use related fractures because post-depositional fractures are generally isolated or discontinuous, elongated, of variable size and direction, and randomly distributed along the edge (e.g. Grace, 1989; Odell and Odell-Vereecken, 1980:96; Tringham et al., 1974:192). However, blind tests show that qualitative assessments of use wear are prone to error and that wear on non-utilized flakes can be misidentified as use wear (e.g. Odell and Odell- Vereecken, 1980:114; Shea and Klenck, 1993). Fortunately, the distribution of fractures along tool edges is a trait that can be quantified and statistically tested against an equal probability (“random”) distribution, that may be a good proxy for post-depositional processes (Bird et al.,

147

2007; Schoville, 2010; Schoville and Brown, 2010). Tools used ephemerally may not develop extensive wear, making it difficult to determine function confidently for a single artifact. The method used here pools edge damage distribution data for a large sample of artifacts, making it possible to determine where on point edges damage is most likely to occur. Because there is no attribution of causation to individual fractures, the method employed here is less affected by the kind of subjectivity inherent in other types of use wear studies. It is also more suitable for the KP1 assemblage, because of the patinated condition of most of the pieces, as discussed above.

Edge damage distribution was quantified using methods described by (Schoville, 2010) with some minor modifications. The sample (n=106) consisted of only the complete retouched points (n=69), and complete non-retouched convergent flakes and blades (n=37) from stratum 4a in squares F21, F23, C21, C23.

The dorsal and ventral side of each point was photographed on a grid with 1cm by 1 cm divisions using a Nikon D5000 with 60mm macro lens to minimize image distortion. The digital images were georectified in ESRI ArcGIS 10 using the grid as landmarks for the appropriate coordinates. A polyline shape file was created for each point and used to trace the perimeters (Figure 25). The use of a polyline rather than a polygon shape file (c.f. Schoville, 2010) simplifies the mapping process. Tracing started at the edge of the platform at the base of each lateral edge, so that the platform was excluded from the outline. While being traced in ArcGIS, point edges were observed for fractures. Only fractures visible to the naked eye were mapped, but low-power microscopy (10-50x) was used as an aid to confirm the presence and nature of the damage. An individual line represented each homogenous zone of the edge, and each line was coded by side (L= left, R=right) and the following types of fractures: “PED” (potential edge damage), “post” (post-patination) and retouch (Figure 25, Table 24).

PED is used as the descriptor for damage of unknown origin following Bird et al. (2007). These are scars or snaps visible to the naked eye that occurred before patination (i.e., they are the same color as the rest of the surface of the tool). These represent potential use damage, but could also have resulted from post-depositional processes. Causation is not attributed to individual PED scars.

“Post-patination” is a descriptor for a scar or snap that exposes ‘fresh’ raw material that is a different color from the patinated surface of the tool. This kind of damage certainly occurred

148

Figure 25 Edge damage distribution quantification methods. A. Screen shot of digitized image with polyline shape file. Post-patination scars on the right edge are highlighted. Attribute table shows calculated line lengths. B. Resulting output showing location of scars on edge scaled to 100. Base =1, Tip=99. C. Resulting edge damage distribution for ventral right edge of one point, calculated as number of damage occurrences at each location/total number of damage occurrences on that edge. (from Wilkins et al., 2012, supplementary material).

149

Table 24 Types of edge fractures identified on KP1 points. Type of Edge Code Definition Example Potential PED Damage of unknown causation edge damage

Post- post Damage that exposes fresh raw material underlying patinated patination surface, which clearly occurred after discard and is not related damage to utilization. A proxy for post-depositional processes.

Retouch retouch Continuous and relatively invasive flake scars most likely related to shaping or resharpening of the edge

Edge edge Undamaged flake edge

after deposition, and after enough time had passed to patinate the surface of the tool. Some or all of these scars could have occurred during excavation and/or box storage. The patination of the KP1 points provides an advantage in this case. Because post-patination scars are easily identified on many points (Table 24), the KP1 points provide an opportunity to determine the actual distribution of non-use related fractures. The post-patination distribution on the KP1 points can be tested against a random distribution, and also used to test the null hypothesis that pre- patination fractures (PED) resulted mainly from post-depositional processes rather than use.

Continuous zones of large and invasive patinated flake scars are identified as retouch. Retouch was coded so that point edges with retouch could be separated from point edges with no retouch. PED fractures were never identified within retouched zones. Because PED scars cannot confidently be identified within zones of retouch, retouch on edges will affect the resulting distributions of PED (i.e., PED scars will appear to be absent from zones with retouch only

150 because they could not be identified). For that reason, when distributions of PED are examined, only non-modified (i.e., not retouched) point edges are included. For example, if a point has retouch on both dorsal laterals, then only the ventral edges are used to analyze frequencies and distributions of PED. It is however, still possible to distinguish post-patination scars within these zones of retouch.

Once damage was digitized in ArcGIS, line lengths were calculated (Figure 25A) and exported to Excel. Excel was used to calculate total edge length and scale it to 100. Total edge length was scaled to remove the effect of size, so that the relative location of each scar with respect to the point base and tip could be calculated (Schoville, 2010:383-384). The resulting data matrix consists of each point edge (i.e. specimen number 3353, ventral, right) in rows and 100 locations (expressed as percent of total edge length) in columns. The presence or absence of each damage type is expressed as either “1” where there is damage, or “0” where there is no damage. For example, the frequency of PED at location 1 (i.e., 1% of total edge length) is equal to the sum of points with damage at that location. The resulting output expresses the relative location of each ‘scar’ as a percent of the total edge length (Figure 25B), and data for all points was pooled to determine assemblage-level frequency and distribution (Figure 25C) patterns.

Edge damage frequency for each location on a given edge is calculated as the number of points with damage at that location divided by the total number of points with relevant data for that edge (i.e. dorsal left, ventral right, etc.). For example, assume there are 16 points that exhibit PED at location 75 on the dorsal left edge and there are 52 points that have unmodified (not retouched) dorsal left edges. The relative frequency of damage at location 75 is 16/52 (30.1%). In other words, 30.1% of points exhibit PED scars at location 75 on the dorsal left edge. Calculating relative frequency in this way accounts for the fact that, because of retouch, there are a different number of points with relevant data for each edge. This frequency is calculated for each location along each edge. Mann-Whitney U tests are used to establish whether damage occurs more frequently on one edge over the other. The Mann-Whitney U test is a non- parametric statistical test that uses ranks to determine whether one of two samples tends to have larger values than the other.

Edge damage distribution on a given edge is calculated as the frequency of damage at each location divided by the total frequency of damage for that edge (i.e. dorsal left, ventral right,

151 etc.). Continuing with the example above, assume that there are 16 occurrences of PED at location 75 on the dorsal left edge and 893 total PED locations on unmodified dorsal left edges. The distribution is expressed as 16/893 (1.8%). In other words, of all the PED scars that occur on the dorsal left edge, 1.8% occur at location 75. These data were transformed into cumulative distributions and following (Schoville, 2010), subjected to the Kolmogorov-Smirnov (KS) test to establish whether damage is distributed differently along different edges. The KS test is commonly used in archaeology to compare the cumulative distributions of two samples and determine whether they may have been drawn from populations with the same distributions (Shennan, 1997).

If PED distributions are significantly different from post-patination distributions based on a KS test, then we can argue that PED distribution represents damage from utilization, rather than post-depositional processes.

If the PED frequencies and distributions represent damage from utilization and the points were used mainly as spear tips, then we would expect the following: 1. The frequency of damage on the left and right sides may be similar and a Mann-Whitney U test on the damage frequencies may indicate that any differences were not statistically significant. Theoretically, forces exerted on a hafted spear point resulting from impact and penetration would not favor one lateral edge over the other at the assemblage level. Experimental research using quartzite point replicates (Schoville and Brown, 2010), but not banded ironstone point replicates (see below) supports this assertion. 2. The left and right distributions would have similar distributions and a KS test between the left and right distributions would indicate that the distributions were not significantly different. Logically, forces exerted on a hafted spear point resulting from impact would not influence one lateral edge differently than the other at the assemblage level, and damage would be spread similarly along each edge. Experimental research using quartzite point replicates (Schoville and Brown, 2010) and banded ironstone point replicates (see below) supports this assertion. 3. Damage would be concentrated at the tips of the points where the point makes first contact with the target. Experimental research supports this assertion (Schoville and Brown, 2010, and this study).

152

4. Damage may also be concentrated near the proximal end of the point, where the point contacts the haft and the binding material. Experimental research supports this assertion in some cases (Schoville and Brown, 2010), but not in the experimental study that is described below.

If the PED frequencies and distributions represent damage from utilization and the points were used mainly as cutting tools, then we would expect the following: 1. If one edge was favored for cutting, then the frequency of damage on the left and right sides may differ and a Mann-Whitney U test would indicate that the differences were significant. 2. The left and right distributions may differ and a KS test between the left and right distributions would give significantly different results. Points from the MSA cave site of PP13B exhibits this pattern and are interpreted as mainly cutting tools (Schoville, 2010). 3. Damage would NOT be concentrated at the tips of the points, but rather along most of the length of the lateral edges, where the points would be making contact with the flesh, hide, or vegetal material that was being cut. PP13B points exhibit this pattern and are interpreted as mainly cutting tools (Schoville, 2010) 4. If the cutting tools were hafted, damage may also be concentrated near the proximal end of the point, where the point contacts the haft and the binding material.

6.1.1.1 Experimental study

An experimental study was conducted to expand previous work on edge damage distribution patterns for spear tips (Schoville and Brown, 2010) and to create experimental patterns of edge damage using banded ironstone, which was the most frequently used raw material for the KP1 points. The experimental study permitted further evaluation of and justification for the above hypotheses regarding edge damage distributions for thrusting spear tips. Experimental retouched points and convergent flakes and blades similar to those recovered from KP1 were replicated by Kyle S. Brown using banded ironstone (n=32). Each point was hafted to a wooden dowel using a combination of Acacia karoo mastic and cow tendon (Figure 26A). Fire was used to aid in the drying process (Figure 26B). A calibrated crossbow designed after (Shea et al., 2001) was used to deliver and control the draw force (Figure 26C,D). Two (Antidorcas marcupialis) carcasses culled from a nearby ranch served as the target. Each surviving point was thrust until

153

Figure 26 Experimental methods. Hafted banded ironstone replicates on wooden dowel ‘foreshaft’, Acacia resin mastic and cow tendon bindings. B. Hafted ‘foreshaft’ replicates drying. C. Calibrated crossbow and springbok carcass target. D. Hafted replicate in metal coupling that links ‘foreshaft’ to main shaft. (from Wilkins et al., 2012, supplementary material)

there was visible damage, which sometimes occurred after a single trial, and up to a maximum of nine trials. The same methods for determining edge damage distributions on the KP1 points were applied to the experimental points.

6.1.2 Diagnostic impact fractures The second method used to determine the function of the KP1 points was an examination for diagnostic impact fractures (DIFs). DIFs are types of macrofractures that occur on experimental weapon tips due to the longitudinal forces exerted on points during impact (e.g. Fischer et al., 1984; Lombard, 2005c; Lombard and Pargeter, 2008) and when found on archaeological specimens are used to support the interpretation of their function as weapon tips (e.g. Lombard, 2005a, c, 2007, 2011; Lombard and Pargeter, 2008; Lombard et al., 2004; Lombard and

154

Phillipson, 2010; Villa et al., 2009a; Villa et al., 2005; Villa and Lenoir, 2006; Villa et al., 2009b; Villa et al., 2010).

There are four DIF types: 1. Step-terminating bending fractures: Step-terminating fractures end abruptly and meet the surface of the flake at a right angle (Figure 27(Hayden, 1979)). Bending fractures initiate without the formation of a Hertzian cone and consequently lack a negative bulb of percussion (Figure 27). In one set of experimental studies (Fischer et al., 1984), step- terminating bending fractures occurred commonly on arrowheads and spear tips used on animal targets (46%), but not when trampled (Table 25). Trampling experiments conducted by Pargeter (2011a) showed that step-terminating bending fractures do occasionally occur on un-used and trampled artifacts, but infrequently (<1%, Table 25) 2. Spin-off fractures > 6 mm: Spin-off fractures are cone fractures that initiate off bending fractures (Figure 27). Spin-off fractures greater than 6 mm in maximum length occur on 23% of experimental arrowheads and spear tips, but not on trampled points (Fischer et al. 1984, Table 25), whereas smaller spin-off fractures (<6 mm) do occur quite frequently on trampled points (Fischer et al., 1984:26). Pargeter (2011a) showed that large spinoff fractures > 6mm can occur on non-used artifacts, but very rarely (<1%). 3. Bifacial spin-off fractures: The presence of multiple spin-off fractures that initiate off both faces of a bending fracture occur on 19% of experimental arrowheads and spear tips, but not on trampled points (Fischer et al., 1984:26; Pargeter, 2011a). 4. Impact burinations: Impact burinations resemble a blow occurring along either one of the lateral edges, but lack the negative bulb of percussion common to deliberate burination (Lombard, 2005c). This kind of fracture is observed in experimental (Barton and Bergman, 1982; Bergman and Newcomer, 1983) and spear (Lombard et al., 2004) studies. Pargeter (2011a) showed that impact burinations can result from trampling and knapping, but rarely (<2%).

Multiple variables influence DIF formation. DIFs can occasionally result from a variety of other activities such as trampling and knapping, though in these cases, DIFs usually occur in low frequencies (Pargeter, 2011a). Site type may also have an influence on the frequency of impact fractures. Kill sites/open air sites appear to have higher DIF frequencies (≥ 40 %) than residential

155

Figure 27 Macrofracture type after Fischer et al. (1984). Step-terminating bending fractures and spin-off fractures >6 mm in maximum length or occurring on both faces of a bending fracture are considered DIFs.

Table 25 Summary of diagnostic impact fracture counts and frequencies from experiments reported by Fischer et al. (1984), Pargeter (2011), and Lombard et al. (2004). Step- terminating Spin-off Bending Fractures Bifacial Spin- Impact Total points Sample Fractures >6mm off Fractures Burinations with DIFs Size n(%) n(%) n(%) n(%) n(%)3 (Fischer et Arrows* 107 47(44) 24(22) 19(16) n/a 48(45) al., 1984) Spears 11 7(64) 3(27) 3(27) n/a 6(55) Total weapons 118 54(46) 27(23) 22(19) n/a 54(46) Trampling1 304 0 0 0 n/a 0 (Pargeter, Cow trampling 250 2(0.8) 0 0 4(1.6) 5(2.0) 2011a) Human 200 2(1) 0 0 1(0.5) 3(1.5) trampling Total trampling 450 4(0.9) 0 0 5(1.1) 8(1.8) No use, no 327 2(0.6) 1(0.3) 0 3(0.9) 6(1.8) trampling2 (Lombard Spears 35 ~14(~40)** n/a ~6(~17) ** n/a 20(57) et al., 2004)

*Only includes those used on whole animals or simulated animal targets made of hide, meat, and bone. 1Includes “walked upon”, “rolled upon by a stone”, and “dropped stones upon”.2Described in original paper as “knapping” damage.3 The data presented here accounts for the fact that individual points may exhibit multiple impact fractures.**Estimated from Figure 3 by (Lombard et al., 2004)

156 sites (~4-30%), probably because broken tools associated with hunting activities would more often be discarded at kill sites (Villa et al., 2009a). Experimental work shows that DIFs can occur on weapon tips of various morphologies and of various raw materials (Fischer et al., 1984; Lombard et al., 2004; Pargeter, 2007) , though how point morphology and raw material may affect the frequency of macrofractures is not well known. The thrusting spear experiments described above have an additional role to play here, providing data on DIF formation on banded ironstone points.

If the KP1 points were used as spear tips, and if banded ironstone points used experimentally as thrusting spear tips exhibit DIFs, then we would expect the KP1 points to exhibit DIFs. The frequency of the DIFs should be greater than 3% (Pargeter, 2011a) to rule out post-depositional processes as the primary cause.

If the KP1 points were used mainly as cutting tools, then we would expect a low frequency (<3%) of DIFs.

6.1.3 Basal modification

MSA points sometimes exhibit evidence of modification near the proximal end of the point that may have facilitated hafting. Removal of the original striking platform and flaking of the ventral surface is a way of accommodating the bases of the points to hafting requirements. At ≠Gi and Aduma, MSA points have bases that are thinned and modified and sometimes the platforms are entirely removed (Brooks et al., 2006). At Sibudu, 25% of points have thinned bases (Villa et al., 2005). At Rose Cottage Cave, 19% (9 out of 47) of points exhibit evidence for basal modification (Villa and Lenoir, 2006).

The KP1 points were also examined for evidence of modifications near the proximal end of the point that may have facilitated hafting. If the KP1 points were hafted, there may be some evidence for basal modification.

6.1.4 Point metrics

If the KP1 points were used as spear tips, then their size and morphology should be consistent with (1) experimental assemblages that functioned well as spear tips and (2) assemblages of MSA points that have been interpreted as spear tips. When considering point metrics, it is

157 important to emphasize that they inform us of potentiality only, and not actual function (Sisk and Shea, 2011).

Calibrated cross bow experiments of (Shea et al., 2001) demonstrate that longer, narrower points fail catastrophically on impact more often than shorter, wider points. These longer points tend to fracture transversely near the midpoint of their length, or break distally removing a significant part of their tip. Of the various point sizes and shapes tested, 28 of the 54 points emerged from multiple uses essentially undamaged, and Shea et al. (2001) suggest that these undamaged points provide the possible dimensions for the optimal design of points intended for use as armatures for thrusting spears (Figure 28). If the KP1 points were used as spear tips, their dimensions should fall within this zone.

Shea (2006) reports tip cross-sectional area (TCSA) values for points from MSA assemblages at Klasies River Mouth, Blombos Cave, Porc Epic, North Africa (tanged Aterian points), and Still Bay points from various sites. TCSA is an important factor influencing penetration depth and thus, the killing power of hunting weapons (Hughes, 1998). Because the TCSA values of these assemblages of MSA points are significantly greater than the means of ethnographic darts and arrows, they were argued to be spear tips, rather than projectile tips (Shea, 2006). The means for these assemblages fall between 100 and 200 mm2, with most points falling below ~420 mm2. The TCSA values for the majority of the experimental points in previous experiments (68%, Shea et al. 2001) fall between 100 and 250 mm2 and their overall mean TCSA is 168 mm2 (Shea, 2006). Additional data is available from Sibudu Cave, where MSA points that also exhibit diagnostic impact fractures have a mean TCSA of 130 ± 60 mm2 (Villa et al., 2009a), consistent with the ranges reported by (Shea, 2006).

Recent research suggests that the metrics of some of the assemblages reported by (Shea, 2006) – from Porc Epic Cave and Aterian sites - are consistent with projectile use rather than spear use, based on tip cross-sectional perimeter (TCSP), which is argued to be a stronger predictor of point function (Sisk and Shea, 2011). For that reason, the KP1 point metrics will only be compared to the MSA points at Klasies River Mouth, Blombos Cave, Sibudu Cave, and Still Bay points. Mean TCSP values for MSA assemblages fall between ~50 and 80 mm, with most points falling below ~120 mm (Sisk and Shea, 2011). For the KP1 points to feasibly serve as spear tips, they should exhibit TCSA and TCSP values consistent with the above studies.

158

Figure 28 Scatterplot of point length and width showing ‘optimal zone’ after Shea et al. (2001). Open squares = points still useable. Open circles=points require some minor repair. Closed circles=points broken and no longer useable.

6.2 Results

The results section is divided into four parts, each focused on one of the four methods for determining point function as discussed above – edge damage distribution, DIFs, basal modifications, and point metrics.

6.2.1 Edge damage distribution

The frequency of PED on each edge compared to the other edges has the potential to inform us of point function. As outlined above, the original expectation for an assemblage of points used mainly as spear tips and not affected by post-depositional processes is for each lateral edge to have a similar frequency of damage. Table 26 presents a summary of PED frequency for all unmodified (non-retouched) edges on the KP1 points.

159

Table 26 PED frequency on unmodified (non-retouched edges) of KP1 points (n=106). All points (n=106) Retouched points (n=69) Non-retouched points (n=37) Edge Number Number Average Number Number Average Number Number Average of PED of edges number of PED of edges number of PED of edges number locations of PED locations of PED locations of PED locations locations locations per edge per edge per edge dorsal left 893 52 17.2 193 14 13.8 700 37 18.9 dorsal right 611 47 13.0 124 9 13.8 487 37 13.2 ventral right 940 100 9.4 648 62 10.5 292 37 7.9 ventral left 1033 101 10.2 594 63 9.4 439 37 11.9 dorsal 1504 99 15.2 317 23 13.8 1187 74 16.0 ventral 1973 201 9.8 1242 125 9.9 731 74 9.9 dorsal 1833 152 12.1 841 76 11.1 992 74 13.4 left/ventral right dorsal 1644 148 11.1 718 72 10.0 926 74 12.5 right/ventral left TOTAL 3477 300 11.6 1559 148 10.5 1918 148 13.0

The dorsal edges exhibit more damage than the ventral edges. When all points are examined in aggregate, an average of 15.2% of each dorsal edge is damaged, whereas only 9.8% of each ventral edge is damaged. The sample size for the dorsal surface is smaller, because retouch occurs much more frequently on the dorsal side than the ventral side. Retouched edges are excluded from interpretation involving PED distribution, because it is not possible to confidently identify PED damage within retouched zones.

One lateral edge of the points (dorsal left/ventral right, 12.1%) exhibits slightly more PED than the other edge (dorsal right/ventral left, 11.1%). Considering just the ventral surface, the ventral left edge has slightly more damage (10.2%) than the ventral right edge (9.4%).

The frequency of edges with PED for each location are presented in Figure 29A. For the graphs, frequency is calculated as the number of points with damage at that location divided by the total number of points with relevant data for that edge. Mann-Whitney U tests (PAST, Hammer et al., 2001) were used on these frequencies to assess whether certain edges of the points preferentially formed PED. A Mann-Whitney U test comparing the frequency of damage on the dorsal and

160

Figure 29 Edge damage frequency on KP1 points , calculated as number of points with damage at each location/total number of points with relevant data for that edge. A. PED. B. Post-patination.

161 ventral surfaces combined indicates that there are significantly more edges with damage on the dorsal side than the ventral side (U=3552, z=3.380, p<0.001). A Mann-Whitney U Test comparing the frequency of damage on either combined lateral edge of the points (dorsal left/ventral right vs. dorsal right/ventral left) indicates that the dorsal left/ventral right edge has significantly more damage than the other (U=3438, z=3.631, p<0.001). The frequency of damage is greater on the ventral left side than the ventral right side, and a Mann-Whitney U test comparing the ventral frequencies indicates the difference is significant (U=4054, z=-2.116, p=0.03).

To summarize, the PED frequencies on the KP1 points indicate that there is more PED damage on the dorsal than the ventral surface, and one edge has more damage than the other does. The ventral left edge has more damage than the ventral right edge. This does not fit the original expectation for hafted spear tips, but the role of post-depositional processes have not yet been considered, nor have the results of the experimental spearing experiment conducted here.

The post-patination damage on the KP1 points serves as a proxy for post-depositional processes and the frequencies can be compared to the PED frequencies in order to test whether the majority of PED scars on the KP1 points are of taphonomic origin. Table 27 presents a summary of post- patination damage frequency for all edges on the KP1 points.

The dorsal surface exhibits more damage than the ventral surface. On average, 11.2% of each dorsal edge is damaged and 10.1% of each ventral edge is damaged. The two laterals of the points exhibit similar frequencies of post-patination damage (10.8% dorsal left/ventral right vs. 10.5 % dorsal right/ventral left). The ventral right edge shows slightly more damage (10.7%) than the ventral left edge (9.6%).

The frequency of edges with post-patination damage for each location are presented in Figure 29B. A Mann-Whitney U test comparing the frequency of damage on the dorsal and ventral surfaces indicates that the difference is statistically significant (U=727, z=10.370, p<0.001). A Mann-Whitney U test comparing the frequency of damage on either side of the points (dorsal left/ventral right vs. dorsal right/ventral left) indicates that there is no significant difference (U=4415, z=1.208, p=0.227). The frequency of damage is greater on the ventral right side than the ventral left side (Table 27), and a Mann-Whitney U test comparing the ventral frequencies indicates the difference is significant (U=3828, z=-2.679, p<0.01).

162

Table 27 Post-patination damage frequency on KP1 points (n=106). All points (n=106) Retouched points (n=69) Non-retouched points (n=37)

Edge Numb Number of Average Number of Number of Average Number of Number of Average er of edges number of post- edges number of post- edges number of post- post- patination post- patination post- patina patination scarred patination scarred patination tion scarred locations scarred locations scarred scarre locations locations locations d per edge per edge per edge locatio ns dorsal left 1169 106 11.0 699 69 10.1 470 37 12.7 dorsal right 1204 106 11.4 751 69 10.9 453 37 12.2 ventral right 1130 106 10.7 717 69 10.4 413 37 11.2 ventral left 1015 106 9.6 591 69 8.6 424 37 11.5 dorsal total 2373 212 11.2 1450 138 10.5 923 74 12.5 ventral total 2145 212 10.1 1308 138 9.5 837 74 11.3 dorsal 2299 212 10.8 1416 138 10.3 883 74 11.9 left/ventral right dorsal 2219 212 10.5 1342 138 9.7 877 74 11.9 right/ventral left Total 4518 424 10.7 2758 276 10.0 1760 148 11.9

To summarize, there is more post-patination damage on the dorsal surfaces of the KP1 points than the ventral surfaces. This is a similar pattern observed for the PED scars, so post- depositional patterns might explain the increased frequency of dorsal PED scars compared to ventral PED scars. The lateral edges have similar frequencies of post-patination damage when dorsal and ventral edges are combined, but there is slightly more damage on the ventral right side than the ventral left side. This is the opposite pattern as observed for the PED, which shows more damage on the ventral left side, implying that post-depositional processes do not explain the PED on the ventral surface of the KP1 points.

Thus far, I have presented results concerning the frequency of damage and comparison of these frequencies between different edges of the points. What follows is a consideration of how damage is distributed along each edge. The expectation for an assemblage of points used mainly as spear tips is for more damage to occur at and near the tips of the points, with less damage along the edges. There may also be a concentration of damage near the proximal end where the haft make contact with the points. Figure 30 presents the edge damage distribution for all unmodified edges on the KP1 points, comparing PED and post-patination damage.

163

Figure 30 Edge damage distribution on KP1 points, calculated as number of damage occurrences at each location/total number of damage occurrences on that edge. A. PED. B. Post-patination.

164

PED on the ventral surfaces is concentrated near the tips, in the last 10% of the point edge (Figure 30A). The dorsal right side also exhibits an increased distribution of damage at the tip, though not to the same degree. The dorsal left side does not show an increased frequency at the tip.

The post-patination damage shows a slight increase in damage near the tip for the ventral right and dorsal right edges, but not the dorsal left and ventral left (Figure 30B), indicating that post- depositional processes may influence damage distributions differently for different edges. To test this idea, KS tests were used to compare the cumulative distribution of post-patination damage between edges, and these tests indicate that each edge has a distribution significantly different from every other edge (Table 28).

Previously, it has been suggested that post-depositional processes should produce non-patterned distributions that are not significantly different from “random” (i.e., an equal probability distribution) (Bird et al., 2007; Schoville, 2010). The post-patination scars on the KP1 points provide an opportunity to assess this assumption. KS tests comparing the cumulative distribution of damage on each edge to an equal probability distribution indicate that the dorsal distributions are not significantly different from random, but the ventral distributions are (Table 29, Figure 31). This suggests that non-use related processes can create patterned distributions on the ventral surfaces of points.

In order to rule out the possibility that the PED on the KP1 points resulted mainly from post- depositional processes, the PED distributions for each edge was compared to the post-patination distribution for the equivalent edge. KS tests indicate that the dorsal left and right PED distributions are not significantly different from the dorsal left and right post-patination distributions (Table 30, Figure 32). The ventral left and right PED distributions are significantly different from the ventral left and right post-patination distributions. Thus, the distribution of dorsal edge damage will not serve as an indicator of tool use, but ventral edge damage will serve as an indicator of tool use. In other words, it is unlikely that the distribution of damage on the ventral surface results solely from post-depositional processes, because the distribution of PED differs significantly from the distribution of post-patination damage.

165

Table 28 KS tests comparing post-patination damage distributions between all edges. dorsal left dorsal right ventral right ventral left dorsal left x 0.059* 0.074* 0.119* dorsal right 0.056 x 0.065* 0.110* ventral right 0.057 0.056 x 0.131* ventral left 0.058 0.058 0.059 x

KS value\max difference * significantly different

Table 29 KS tests comparing post-patination distribution on the KP1 points to a “random” (equal probability) distribution. KS Max Diff Significance dorsal left 0.056 0.045 Not sig different dorsal right 0.055 0.048 Not sig different ventral right 0.057 0.082 Sig different ventral left 0.060 0.089 Sig different

Table 30 KS tests comparing PED cumulative distributions on KP1 points to post- patination damage cumulative distributions on the equivalent edge. KS Max Diff Significance dorsal left 0.06044359 0.03945 Not sig different dorsal right 0.067552803 0.03600 Not sig different ventral right 0.06003727 0.086669 Sig different ventral left 0.060106398 0.114336 Sig different

166

Figure 31 Cumulative percent frequency (using data from Figure 30B) of post-patination damage on the KP1 points compared to a “random” (equal probability) distribution. A. Dorsal. B. Ventral.

167

Figure 32 PED vs. post-patination damage cumulative distributions comparing equivalent edges (using data from Figure 30A and B). A. Dorsal left. B. Dorsal right. C. Ventral right. D. Ventral left.

168

The distribution of PED on the ventral unmodified edges can be used to assess the most likely function of the KP1 points. The ventral right and left distributions are not significantly different based on a KS test of the cumulative distributions (max difference=0.0610, KS value=0.0613, Figure 33). There is an increased frequency of damage at the tip of the points, compared to the rest of the point edge (Figure 30A). These observations are consistent with previous experimental spear tips, which show increased frequency at the tip and similar distributions between left and right sides (Schoville and Brown, 2010).

The distribution of PED on the KP1 points is inconsistent with previous spearing experiments because the distribution lacks what has been described as a “hafting bump” (Schoville and Brown, 2010). The ‘hafting bump’ represents the location where damage from the haft and binding material is concentrated at approximately 15-30% of the total edge length. This inconsistency is addressed further below with the presentation of the experimental results.

The KP1 PED distribution is inconsistent with expectations for cutting tools, where we would expect more damage along the edges and less at the tip (Schoville, 2010; Schoville and Brown, 2010), and with points from PP13B that have been interpreted as cutting tools (Schoville, 2010).

6.2.1.1 Experimental edge damage

Edge damage on the experimental banded ironstone points used as thrusting spear tips can be compared to the KP1 points to further assess point function. The experimental points considered here add to previous experimental work (Schoville and Brown, 2010) and also provide a raw material specific comparative sample. Table 31 presents a summary of damage frequency for all unmodified (non-retouched) edges on the experimental points.

The dorsal and ventral edges of the experimental points exhibit similar amounts of damage. In total, 10.6% of the dorsal edges are damaged, and 11.4% of the ventral edges are damaged. The sample size for the dorsal surface is much smaller, because in the experimental sample retouch occurred much more frequently on the dorsal side than the ventral side and retouched edges are excluded from the analysis. One lateral edge (dorsal left/ventral right, 12.5%) exhibits more damage than the other (dorsal right/ventral left, 10.3%).

169

Figure 33 PED ventral left vs. ventral right cumulative distributions (using data from Figure 30A).

Table 31 Damage frequency on experimental spear tips (n=32). Edge Number of Scarred Locations Number of Edges Average number of scarred locations per edge dorsal left 6 2 3.0 dorsal right 68 5 13.6 ventral right 381 29 13.1 ventral left 293 30 9.8 dorsal total 74 7 10.6 ventral total 674 59 11.4 dorsal left/ventral right 387 31 12.5 dorsal right/ventral left 361 35 10.3 Total 748 66 11.3

170

The frequency of ventral edges with damage for each location are presented in Figure 34. The dorsal surfaces are excluded from this plot because the small sample size distorts the resulting plot. A Mann-Whitney U test comparing the frequency of damage on the dorsal and ventral surfaces (Figure 35) indicates that there is significantly more damage on the ventral side than the dorsal side (U=3518, z=3.470, p<0.001). A Mann-Whitney U test comparing the frequency of damage on the lateral edge of the points (dorsal left/ventral right vs. dorsal right/ventral left) indicates that one edge (dorsal left/ventral right) has significantly more damage than the other (U=3759, z=2.840, p<0.01). The frequency of damage is greater on the ventral right side than the ventral left side (Table 31), and a Mann-Whitney U test comparing the ventral right and ventral left frequencies indicates the difference is significant (U=3828, z=-2.705, p<0.01).

With respect to edge damage frequency, there are both similarities and differences between the PED on the KP1 points and damage on the experimental points. Like the KP1 points, the experimental points show more damage on one lateral edge. Unlike the KP1 points, there is more ventral damage than dorsal damage and the lateral edge (left vs. right) on which the increased damage occurs is different. These differences are probably due to the influence that post- depositional processes had on the KP1 points, and I will discuss this interpretation further in the discussion section below.

Edge damage distribution on the experimental points should show similarities with PED distribution on the KP1 points, if the KP1 points were also used as thrusting spear tips. Figure 36A presents the edge damage distribution for the experimental spear tips. The results presented here focus only on the ventral side, because the dorsal side of the experimental points is too small of a sample to be meaningful and because the ventral side of the KP1 points was shown to be significantly different from that expected for post-depositional processes, and therefore of more behavioral consequence.

There is an increased frequency of damage at the tip of the experimental points, compared to the rest of point edge (Figure 36A). The ventral left and right distributions are not significantly different based on a KS test of the cumulative distributions (max difference=0.0843, KS value=0.1057, Figure 36B). These observations are consistent with previous experimental spear tips, which show increased frequency at the tip and similar distributions between left and right sides (Schoville, 2010; Schoville and Brown, 2010).

171

Figure 34 Edge damage frequency on experimental spear tips , calculated as number of points with damage at each location/total number of points with relevant data for that edge.

Figure 35 Comparison of dorsal and ventral damage edge damage frequencies on the experimental spear tips , calculated as number of points with damage at each location/total number of points with relevant data for that edge.

172

Figure 36 Edge damage distribution on experimental spear tips. A. Percent frequency of damage, calculated as number of damage occurrences at each location/total number of damage occurrences on that edge. B. Cumulative percent frequency of damage, using data from A.

173

The distribution of damage on the experimental KP1 points is inconsistent with previous spearing experiments, because the distribution lacks a “hafting bump” (Schoville and Brown, 2010). Possible explanations for this finding include the hardness or other unique characteristics of banded ironstone compared to quartzite, and these ideas are discussed further below.

PED on the ventral surface of the KP1 points exhibits the following similarities with the ventral surface of the experimental spear tips: 1. The frequency of damage on one ventral edge is significantly greater than on the other ventral edge. 2. The ventral left and right distributions of damage are not significantly different 3. Damage is concentrated near the tip of the point. 4. There is no “hafting bump”. The only difference concerns the edge that exhibits a higher frequency of damage. For the KP1 points, the ventral left edge has slightly more damage. For the experimental points, the ventral right edge has more damage.

6.2.1.2 Summary of edge damage distribution results

The PED fractures on the KP1 points occurred prior to patination, but each individual scar could have resulted from use or non-use (post-depositional) related forces. However, even if each individual fracture has unknown causation, assemblage-level distribution patterns permit, first, an evaluation of whether the main cause of the distribution pattern is related to use or to post- depositional processes, and second, an evaluation of which functions best explain the use-related distribution patterns.

To test whether the PED distributions on the KP1 points were best explained by post- depositional processes or not, the post-patination distributions for each edge were tested against the PED distribution of the same edge. The results showed that the dorsal PED distributions were not significantly different from the post-patination pattern, but the ventral PED distributions were. Thus, the dorsal PED distributions are not likely to reflect patterns of use, but the ventral PED distributions are and can be used to assess point function.

174

The frequency and distribution of ventral damage on the KP1 points fit some of the expectations for spear tips based on previous work. The distributions on the left and right sides are not significantly different, and damage is concentrated at the tips of the points.

There are also two inconsistencies with the original expectations for the spear tips based on previous experimental research. First, the ventral left side exhibits statistically more damage than the ventral right. Second, there is no ‘hafting bump’. However, the banded ironstone experimental spear tips presented here also have a greater frequency of damage on one of the ventral edges, consistent with the PED distribution on the KP1 points, and they did not exhibit a ‘hafting bump’. The implication is that the expectations for hafted spear tips must be modified based on the experimental results of this study.

Table 32 presents a summary of the edge damage frequency and distribution expectations (original and modified) and the KP1 results. The KP1 points fit all modified expectations for spear tips based on the results of the current experimental study. The KP1 points are inconsistent with expectations for cutting tools, for which the left and right distributions should be significantly different from each other and there should be no increase in damage at the tip of the points. This is the pattern observed on points from the MSA site of PP13B (Schoville, 2010; Schoville and Brown, 2010). Based on experimental evidence presented here, the different frequencies of PED between the left and right sides and the lack of a hafting bump neither support nor reject the hypothesis that the points were used as spear tips. However, two important observations are consistent with expectations for the spear tip hypothesis and inconsistent with the cutting tool hypothesis -- PED is concentrated at the tips of the points and PED distributions are similar between the left and right sides

6.2.2 Diagnostic impact fractures

Diagnostic impact fractures (DIFs) are features seen in experiments where weapon tips impact animal targets (e.g. Fischer et al., 1984) and their presence implies that archaeological points were used as weapon tips (Lombard, 2005a, c, 2007; Villa et al., 2009a; Villa et al., 2005; Villa et al., 2009b; Villa et al., 2010). Similar-appearing fractures can result from post-depositional processes, but the frequency within an assemblage is low (<3%, Pargeter, 2011a).

175

Table 32 Summary of expectations based on previous work (Bird et al., 2007; Schoville, 2010; Schoville and Brown, 2010) and KP1 results. Results presented for ventral surface only because dorsal surfaces did not give a pattern significantly different from post- patination distribution. Expectation for Modified Expectation for KP1 results- ventral SPEAR TIPS based expectation for CUTTING TOOLS surface only on previous studies SPEAR TIPS based (Bird et al., 2007; on current Schoville, 2010; experimental study Schoville and Brown, 2010) PED frequency Left and right Left and right Left and right Left and right frequencies NOT frequencies frequencies frequencies significantly significantly significantly significantly different different different different PED distribution Left and right Left and right Left and right Left and right distributions NOT distributions NOT distributions distributions NOT significantly significantly significantly significantly different different different different Damage Damage Damage NOT Damage concentrated at tips concentrated at tips concentrated at tips, concentrated at tips but along length of lateral edges ‘Hafting bump’ No ‘hafting bump’ ‘Hafting bump’ (if No ‘hafting bump’ hafted)

The KP1 points exhibit DIFs (Figure 37). Counts and frequencies for the different DIF types are presented in Table 33. There are a total of 31 DIFs on 29 points and point fragments in the sample of 210. This gives a frequency of points with DIFs of 13.8%, including all points and point fragments and DIFs with distal and proximal initiations. Even calculating DIF frequency more conservatively by only including distal DIFs on complete points and point tips, gives a frequency of 9.9% (16 points and point tips with DIFs out of a sample 162). Both retouched points and non-retouched points exhibit relatively high DIF frequencies (Table 33).

The irregular pointed forms do not exhibit a DIF frequency (3.7%) much greater than expected (~3%) given the known influence of post-depositional processes (Pargeter, 2011a). It is possible that the irregular pointed forms, which were identified prior to examining macrofractures based on their uneven lateral edges, thick edge angles, and/or tips that are not very pointy, exhibit a low frequency of diagnostic impact fractures due to post-depositional processes alone. If the irregular pointed forms are excluded from the counts, than the KP1 points have a DIF frequency of 14.2%.

DIFs were observed on the experimental spear tips. Counts and frequencies are given in Table 34. The experimental points exhibit a DIF frequency of 25.0%.

176

Figure 37 Examples of diagnostic impact fractures on KP1 points. A. distal step- terminating bending fracture on ventral surface of complete non-retouched convergent blade, banded ironstone B. distal step-terminating bending fracture on ventral surface of irregular pointed form, black chert C. distal step-terminating bending fracture on ventral surface of complete retouched point, banded ironstone, also shows some post-patination damage at tip D. distal step-terminating bending fracture on ventral surface of complete non-retouched convergent blade, banded ironstone E. Unifacial spin-off fracture 6.6 mm in length originating off dorsal side of distal snap fracture, banded ironstone, irregular pointed form. F. Proximal impact burination on ventral surface of non-retouched convergent flake, banded ironstone G. distal impact burination on ventral surface of complete non-retouched convergent blade, banded ironstone H. distal impact burination on ventral surface of complete non-retouched convergent flake, banded ironstone I. distal impact burination on ventral surface of irregular pointed form, banded ironstone.

177

Table 33 Counts and frequencies of diagnostic impact fractures on the KP1 points and point fragments (n=210). Not- Not- Retouched Irregular Retouched Retouched Retouched retouched retouched Point Pointed Total Type of Fracture Point Point Tip Point Base Convergent Convergent Fragment Form n=210 n=69 n=22 n=25 Point Blade n=4 n=27 n=28 n=35 step-terminating 7(10.1%) 2(9.1%) 1(25.0%) 0 0 2(5.7%) 1(3.7%) 13 (6.2%) bending facture unifacial spin-off 0 0 0 0 0 0 1(3.7%) 1(0.5%) > 6mm Distal bifacial spin-off 0 0 0 0 0 0 0 0

impact 4(5.8%) 0 1(25.0%) 0 1(3.6%) 2(5.7%) 1(3.7%) 9(4.3%) burination step-terminating 0 1(4.5%) 0 0 2(7.1%) 1(2.9%) 0 4(1.9%) bending facture unifacial spin-off 0 0 0 0 0 0 0 0 Proxi > 6mm mal bifacial spin-off 0 1(4.5%) 0 0 0 0 0 1(0.5%)

impact 0 2(9.1%) 0 0 1(3.6%) 0 0 3(1.4%) burination Total number of points 10(14.5%)* 6(27.3%) 1(25.0%)* 0 4(14.3%) 5(14.3%) 3(3.7%) 29(13.8%) with DIFs

*one point exhibits two DIFs

Table 34 Counts and frequencies of diagnostic impact fractures on experimental spear tips (n=32). n(%) Type of Fracture

step-terminating bending facture 8 (25.0%) unifacial spin-off 0 Distal > 6mm bifacial spin-off 0 impact burination 1(3.1%) step-terminating bending facture 0 unifacial spin-off 0 Proximal > 6mm bifacial spin-off 0 impact burination 0 Total number of points with DIFs (out of 32) 8(25.0%)

178

The presence of DIFs on the proximal ends of points is an observation that may also be consistent with hafting (Villa et al., 2009b: 451), though proximal DIFs were not observed on any of the experimental spear tips (Table 34).

Table 35 presents published frequencies for other archaeological and experimental assemblages. The KP1 frequency is well within the range observed at residential sites from more recent archaeological contexts, and at the high end of the range observed at MSA and MP sites. The DIF frequency is much higher than that observed on trampled and non-utilized assemblages. The frequency of DIFs at KP1 is below the range observed at known kill sites and in experimental assemblages.

Interestingly, the frequency of DIFs on the experimental spear points in this study are below those observed in other experimental assemblages. This may indicate that, while banded ironstone does develop DIFs, DIFs may form less often on banded ironstone than other raw material types. The implications of this are discussed further below.

6.2.3 Basal modifications

Of the KP1 points, 23 exhibit some evidence for minor basal modification (Figure 38), giving a frequency of 13.0% (out of 177 points, excluding distal fragments). These points have between 2 and 7 flake scars usually detached off the ventral face at the base. Occasionally, it appears as though flakes detached from the dorsal face remove the entire bulb.

The frequency of basal modification on the KP1 points is similar to frequencies reported for MSA assemblages at Sibudu and Rose Cottage Cave (25% and 13%, respectively)(Villa et al., 2005; Villa and Lenoir, 2006).

6.2.4 Point metrics and spear tip potentiality

KP1point metrics were recorded and compared to MSA assemblages to further test the feasibility that they could have served as spear tips. Table 36 presents a summary of the KP1 point metrics. Figure 39 shows the lengths and widths of the KP1 points plotted with the experimental data from (Shea et al., 2001). Demarcated zones delineate the ‘spear point optimal zone’ from values. l

179

Table 35 Summary of diagnostic impact fracture frequencies in the literature. Number of Site/Levels Industry Site type Frequency DIFs Reference Mean St Dev Points Experimental spears - - 32 25.0 - - - This study KP1 points Fauresmith/early MSA Open air 210 13.8 - - - Oscurusciuto, , Middle Paleolithic Rockshelter 19 5.3 (Villa et al., 2009a) Unit 1 Oscurusciuto, Italy, Unit Middle Paleolithic Rockshelter 38 7.9 (Villa et al., 2009a) 2 Bouheben, Layers 2 and Middle Paleolithic Open air 113 5.3 (Villa and Lenoir, 2006) 11 Middle Paleolithic Sibudu, unifacial points 7.5 3.4 and MSA sites MSA Cave 101 8.9 (Villa et al., 2009a) from layer RSP Rose Cottage, unifacial points from post-HP MSA Cave 47 4.2 (Villa and Lenoir, 2006) layers Blombos Still Bay, MSA Cave 82 13.4 (Villa et al., 2009b) finished bifacial points , ~2800 Muldbjerg, arrowheads Residential 30 30.0 (Fischer et al., 1984) BC Northern Europe, ~3200 Præstelyng, arrowheads Residential 56 14.3 Fischer et al. (1984) BC Vejlebro, level 8, Northern Europe, ~3500 Residential 24 20.8 Fischer et al. (1984) Residential sites arrowheads BC 14.8 9.4 Vejlebro, level 9, Northern Europe, ~3500 Residential 42 4.8 Fischer et al. (1984) arrowheads BC Bromme, tanged points Brommian, ~13 ka Residential 47 6.4 Fischer et al. (1984) Ommelshoved, tanged Brommian, ~13 ka Residential 88 12.5 Fischer et al. (1984) points Stellmoor, tanged points Ahrensburgian, ~12 ka Kill site 45 42.2 Fischer et al. (1984) Kill sites Casper, Hell Gap 42.8 0.8 Paleoindian, ~10 ka Kill site 60 43.3 cited by (Villa et al., 2009a) bifacial points Experimental - - 107 44.9 Fischer et al. (1984) Experimental arrowheads** 52.2 6.4 weapons Experimental spears - - 11 54.5 Fischer et al. (1984) Experimental spears - - 35 57.1 (Lombard et al., 2004) Cattle Trampling, - - 250 2.0 (Pargeter, 2011a) Experimental Human Trampling, - - 200 1.5 (Pargeter, 2011a) 1.8 0.3 trampling No Trampling or Use - - 327 1.8 (Pargeter, 2011a)

* Lombard reports 42% at Sibudu, but calculation based on selected sample of 50 points with visible wear. **Only including experiments on animal targets.

180

Figure 38 Examples of basal modifications on KP1 points. (from Wilkins et al., 2012, supplementary material)

181

Table 36 Summary of KP1 point metrics. Retouched Retouched Retouched Retouched Not-retouched Not-retouched Irregular All Point Point Tip Point Fragment Point Base Convergent Convergent Pointed Form Flake Blade Mass n 69 0 0 0 20 18 20 127 (g) Mean 35.9 n/a n/a n/a 27.2 35.4 34.6 34.3 Min 10.0 n/a n/a n/a 8.5 6.0 4.0 4.0 Max 121.5 n/a n/a n/a 72.5 111.0 102.0 121.5 Std Dev 20.9 n/a n/a n/a 16.0 29.0 23.8 22.0 Length n 69 0 0 0 20 18 20 127 (mm) Mean 71.0 n/a n/a n/a 58.3 78.2 71.1 70.1 Min 41.3 n/a n/a n/a 39.5 49.3 27.8 27.8 Max 117.8 n/a n/a n/a 98.5 117.7 122.8 122.8 Std Dev 18.5 n/a n/a n/a 15.9 19.0 23.7 19.7 Width n 69 8 3 25 28 30 26 189 (mm) Mean 40.3 31.1 25.0 37.3 38.6 34.3 38.0 37.7 Min 24.7 19.6 21.3 22.6 23.4 22.0 20.6 19.6 Max 62.9 46.3 27.2 49.4 56.1 60.6 59.7 62.9 Std Dev 8.3 8.9 3.2 6.9 8.3 8.8 9.6 8.8 Length/Width n 69 0 0 0 20 18 20 127 Mean 1.8 n/a n/a n/a 1.5 2.3 1.9 1.8 Min 1.2 n/a n/a n/a 1.0 1.7 1.1 1.0 Max 2.8 n/a n/a n/a 1.9 3.0 3.2 3.2 Std Dev 0.4 n/a n/a n/a 0.3 0.3 0.6 0.4 Thickness n 69 18 2 25 28 31 27 200 (mm) Mean 12.1 8.5 7.7 10.7 11.2 10.3 11.9 11.1 Min 5.8 4.9 7.1 7.9 6.4 3.7 5.5 3.7 Max 20.1 14.3 8.2 14.2 16.9 16.5 27.3 27.3 Std Dev 3.1 2.2 0.8 1.6 2.6 3.7 4.6 3.3 TCSA n 69 5 2 25 28 27 26 182 (mm2) Mean 249.0 149.5 92.1 202.9 222.9 200.8 242.2 226.1 Min 86.3 61.7 75.6 94.9 89.9 51.7 63.9 51.7 Max 591.3 331.0 108.7 317.3 414.6 446.3 814.9 814.9 Std Dev 101.3 106.6 23.4 60.9 84.9 100.1 146.8 104.2 TCSP n 69 5 2 25 28 27 26 182 (mm) Mean 87.5 66.9 52.3 80.5 83.4 77.0 83.5 82.8 Min 54.1 50.5 46.9 50.8 51.6 45.9 44.6 44.6 Max 136.2 100.7 57.7 103.8 116.8 127.2 140.6 140.6 Std Dev 17.3 19.7 7.6 13.9 17.2 18.6 20.8 18.2

182

“optimal zone”

Figure 39 Scatterplot of length and width of the KP1 points plotted with experimental data from Shea et al. (2001). Closed black circles = broken after use. Open purple circles = minor damage. Open blue squares = still useable. Red crosses = KP1 points

that resulted in catastrophic damage during Shea et al.’s (2001) thrusting spear experiments. The KP1 points fall mainly within the ‘optimal zone’, with 113 points (89% of 127 measurable points) falling within this ‘optimal zone’. The size and shape of the KP1 points indicates that they have good potential for serving as functional spear tips. The mean TCSA at KP1 is slightly higher than, but within the range of, those published for points from MSA assemblages (Figure 40A). The mean TCSP value is also slightly higher, but falls within the range of other MSA assemblages (Figure 40B). The TCSA and TCSP values of the KP1 points are similar to those of other MSA assemblages, where the points are often assumed to have functioned as spear tips, demonstrating the feasibility of the spear tip hypothesis for the KP1 points.

183

Figure 40 Comparison of KP1 point size to MSA point size. A. Box plots of TCSA values for experimental spear tips and MSA spear tips from Shea (2006) compared to KP1. B. Box plots of TCSP values for MSA spear tips from (Sisk and Shea, 2011) compared to KP1. Outliers are excluded from both plots, including two very large points from KP1

184

Table 37 Experimental point characteristics and results of experimental trials. Experimental Length Width L/W Thickness TCSA TCSP Mass Number of Number of Haft Type of Damage Status after Point (mm) (mm) (mm) (g) Penetrations Failures Experiment Number 1 79.2 38.9 2.0 15.9 309.3 89.1 45.5 6 1 Reusable 2 70.2 33.4 2.1 10.7 178.7 73.1 30.8 7 0 Reusable 3 83.7 43.8 1.9 17.8 389.8 100.2 61.1 3 0 step-terminating bending fracture Reusable 4 65.1 33.2 2.0 8.9 147.7 70.9 23.2 1 1 Reusable 5 67.2 38.0 1.8 12.5 237.5 83.5 36.5 6 0 snap fracture Reusable 6 58.8 36.3 1.6 11.7 212.4 79.5 28.9 4 1 Reusable 7 72.3 40.1 1.8 8.5 170.4 83.7 31.9 4 1 Reusable 8 62.8 28.6 2.2 11.0 157.3 64.7 20.9 2 1 Reusable 9 60.0 30.5 2.0 10.0 152.5 67.0 15.9 5 0 step-terminating bending fracture Reusable 10 57.7 39.7 1.5 8.9 176.7 83.2 18.8 2 0 step-terminating bending fracture Damaged (not-reusable) 11 60.7 29.0 2.1 12.0 174.0 66.6 27.8 1 0 step-terminating bending fracture Reusable 12 55.5 24.7 2.2 12.5 154.4 59.8 18.8 4 0 Reusable 13 68.3 24.8 2.8 11.6 143.8 58.8 22.2 1 0 Reusable 14 62.9 40.3 1.6 10.2 205.5 85.5 37.4 5 1 Reusable 15 54.6 22.0 2.5 8.0 88.0 49.2 11.8 2 1 Reusable 16 64.4 35.8 1.8 11.0 196.9 77.8 28.7 7 0 Reusable 17 64.1 30.8 2.1 11.3 174.0 69.0 25.7 3 0 step-terminating bending fracture Reusable 18 62.6 26.4 2.4 10.0 132.0 59.5 16.5 9 0 step-terminating bending fracture Reusable 19 65.8 27.5 2.4 12.6 173.3 64.8 18.6 6 2 Reusable 20 60.0 29.0 2.1 12.0 174.0 66.6 26.8 4 2 snap fracture Minor repair 21 49.2 33.3 1.5 7.9 131.5 70.2 16.4 2 0 crushing Reusable 22 76.3 25.6 3.0 7.3 93.4 55.1 17.3 1 0 Damaged (not reusable) 23 65.7 24.7 2.7 10.5 129.7 57.1 24.6 1 0 crushing Reusable 24 87.6 39.1 2.2 11.8 230.7 84.8 53.9 1 0 impact burination Reusable 25 77.4 40.6 1.9 15.0 304.5 91.1 53.5 4 0 step-terminating bending fracture Reusable 26 71.0 40.8 1.7 12.7 259.1 88.9 54.4 2 0 Reusable 27 56.2 37.0 1.5 8.9 164.7 78.1 21.9 1 0 crushing Reusable 28 57.8 23.1 2.5 8.6 99.3 51.9 12.2 1 0 Reusable 29 62.1 36.4 1.7 9.4 171.1 77.4 28.5 4 1 crushing Reusable 30 71.9 37.9 1.9 14.6 276.7 85.7 48.5 1 0 crushing Reusable 31 76.7 29.6 2.6 9.6 142.1 64.9 23.0 1 0 step-terminating bending fracture Damaged (not reusable) 32 72.0 38.5 1.9 17.8 342.7 90.9 48.2 5 0 Minor repair Mean 66.2 33.1 11.3 2.1 190.4 73.4 29.7 3.31 0.38 Min 49.2 22 7.3 1.5 88.0 49.2 11.8 1 0 Max 87.6 43.8 17.8 3.0 389.8 100.2 61.1 9 2 Std Dev 8.8 6.2 2.7 0.4 71.5 13.1 13.7 2.24 0.61

185

The experimental study presented here adds further support to the argument that the KP1 points could have feasibly served as spear tips. The replicated banded ironstone points performed well in the calibrated crossbow set up and adequately penetrated the target. The points were shot multiple times until damage was visible or haft failure prevented additional trials. Table 37 presents the results of the experimental trials. Though the experiment was not designed to test for spear tip performance, the following observations were made:

 The 32 points were shot a total of 106 times. Each point was shot an average of 3 times before there was visible damage.

 Only 10 points had visible damage after a single shot, 9 points were shot 5 or more times before developing visible damage.

 Out of the 106 trials there were 12 haft failures.

Figure 41 presents a scatter plot of the experimental banded ironstone replicates against the ‘optimal zone’ reported by (Shea et al., 2001). The scatterplot weakly corroborates previous experimental studies, with two of the damaged/not reusable points falling outside the optimal zone. However, there is also one damaged/not reusable point within the optimal zone, and several ‘still useable’ points (9, 28%) falling outside the zone. The plot does show that experimental spear tips with similar metrics as the KP1 points functioned well as spear tips.

6.3 Discussion

Several lines of evidence – edge damage distribution, DIFs, basal modifications, and metrics -- support the hypothesis that the KP1 points were used as spear tips. Edge damage distributions are inconsistent with expectations for cutting tools and post-depositional processes are ruled out as the only cause of damage on the KP1 points. Consistent with the spear tip hypothesis, the KP1 points exhibit increased frequencies of damage at the tips and similar distributions between the left and right edges of the ventral surface. A relatively high frequency of DIFs also indicate that point tips were subjected to the kind of strong longitudinal forces that experimental hunting weapons experience. While there is no definitive evidence for hafting, the KP1 points exhibit two characteristics that are suggestive of hafting - basal modifications and DIFs originating from

186

“optimal zone”

Figure 41 Scatterplot of length and widths of the experimental points plotted with the ‘optimal zone’ of Shea et al. (2001). Closed black circles = broken after use. Open purple circles = minor damage. Open blue squares = still useable. the proximal ends. The size and shape of the KP1 points indicate that they could feasibly function as spear tips.

There may be other functional explanations for the observed characteristics of the KP1 points. PED distributions on the KP1 points are inconsistent with those observed on MSA points from PP13B that have been interpreted as cutting tools (Schoville, 2010)., and the DIFs observed on the KP1 points were not observed on experimental butchery tools (n=2, Fischer et al., 1984), but the study presented here does not rule out all potential functions. Any function using a ‘piercing’ tool motion – defined by (Shea, 1991:93) as one that moves at roughly normal incidence to an incised surface under dynamic load without a significant degree of rotation about the principal axis of loading – could give a similar pattern. There does not seem to be any criteria that allows one to distinguish between wear resulting from the use of spear and the use of a dagger, which would be used to stab prey at a much closer range than a spear, nor is there a method for distinguishing throwing spears from thrusting spears (Shea, 1991:95). Damage at the point tip

187 could also feasibly be explained by heavy-duty and vigorous butchery, though in that case, one might expect more damage to be distributed along the edges. Future experimental work might help to elucidate some means of distinguishing between these potential functions.

Despite this cautionary note, however, use as hafted spear tips is the most reasonable interpretation of the evidence in light of the substantial body of work focused on the role of MSA and MP points in hunting weaponry (see Section 3.4). The KP1 points are morphologically and technologically similar to MSA and MP points, and exhibit the same kinds of evidence used to support the role of MSA and MP points as tips in composite hunting weapons.

6.3.1 Interpreting the edge damage distribution results and methodological considerations

Patination on the KP1 points provides a unique opportunity to examine the effect of post- depositional processes on edge damage frequencies and distribution, and thus, contributes methodologically to edge damage distribution studies. The formation of the post-patination scars on the KP1 points is probably due to a combination of factors, including fluvial and/or colluvial transport prior to deposition, water and sediment action after deposition, and excavation and post-excavation disturbances (i.e. long-term storage in boxes with other lithics). It is not possible to ascertain whether one of these many processes accounts for more of the damage than another process. However, the post-patination scars on the KP1 points permit a consideration of two assumptions about non-use related, post-discard patterns of edge damage. First, it is assumed that post-depositional damage would be randomly distributed along point edges, which means that any location along a point edge would exhibit an equal probability of having damage as any other location (Bird et al., 2007; Schoville, 2010). The post-patination scars on the KP1 points indicate that this is not true in all cases. The post-patination damage on the ventral surfaces of the KP1 points is significantly different from an equal probability distribution. The post-patination distribution on the ventral edges (Figure 30B) appears to show a slight increase of damage at the tip and more damage in the first ~60% of the total edge length. The increased damage at the tip might be because that area of the point is more fragile and prone to breakage than the rest of the edge. The increased damage in the lower portion of the point edge could be related to point form; the edge near the base tends to be slightly more convex than the distal portion of the edge and convex edges have been suggested to be more prone to damage (Tringham et al., 1974:Figure 3). Alternatively, the patterned damage could reflect retouch distribution on the dorsal surface;

188 perhaps retouch on the opposing face, which tends to increase the edge angle, prevents some damage from occurring on the ventral face. There are multiple variables that could influence the effect of post-depositional processes on edge damage distributions and result in patterned distributions.

The KP1 points also demonstrate that post-depositional processes can affect each edge in a different way. The distributions of post-patination edge damage on each edge are significantly different from every other edge (Table 28). Some of this variability is related to the amount of damage at the tip. The dorsal right and ventral right edges exhibit the greatest concentration of post-patination damage at the tip. These edges are opposed and damage would feasibly give this pattern if the points tended to rotate mainly in one direction (dorsal left over dorsal right) due to post-depositional disturbances (ie. rolling/tumbling due to low-medium energy water transport). This interpretation is reasonable given the secondary context of the KP1 points and their position within the artifact-dense sandy spring vents of stratum 4a (Chapter 3).

Post-depositional processes can also result in different frequencies of damage on different edges. The KP1 points exhibit higher frequencies of post-patination damage on their dorsal faces than their ventral faces. When the edges are combined (dorsal left/ventral right and dorsal right/ventral left), there is no significant difference in the frequency of damage between them. However, as discussed above, the dorsal right and ventral right edges exhibit higher frequencies of damage, and it was suggested that this pattern could result from rolling/turning in the sediments or due to some water transport.

By comparing the PED distribution to the post-patination distribution on the same edge, it was possible to distinguish edge distribution patterns primarily reflecting post-depositional patterns from patterns related to tool function. On the KP1 points, only the ventral edges showed PED distributions significantly different from post-patination distributions. There are a couple potential explanations for why significant differences are detected for the ventral but not the dorsal edges. First, more post-patination damage occurs on the dorsal surfaces of the KP1 points, implying that taphonomic processes preferentially damaged the dorsal surface. Second, more damage occurs on the ventral surface of experimental spear tips in previous studies (Schoville and Brown, 2010) and in the current study (Table 31). Thus, post-depositional processes are

189 more likely to obscure dorsal use-related damage patterns than ventral use-related damage patterns.

One unexpected result of this study that was that both the KP1 points (PED) and the experimental spear tips showed a higher frequency (but similar distributions) of damage on the one ventral edge over the other. For PED on the KP1 points, the ventral left side exhibits statistically more damage than the ventral right. The difference, however, is small degree; an average of 10.2 % of each ventral left edge is damaged compared to an average of 9.4% of each ventral right edge. The location on the points where the difference is apparent is between ~16- 36% of the total point edge (Figure 29). This area is where we would expect hafting damage based on the location of the ‘hafting bump’ in previous experiments (Schoville and Brown, 2010). Perhaps, PED frequencies are greater on the ventral left side because of the configuration of an asymmetrical haft. Another possibility is that rotation of the spear (dorsal right edge over dorsal left) during penetration could result in increased damage on the ventral left surface. Rots et al. (2011: 646) suggest that evidence for rotation on lithics from Sodmein Cave are consistent with twisting of thrusting spears as a means to increase the size and severity of the inflicted wound

The experimental spear tips presented here also have a greater frequency of damage on one of the ventral edges (Table 31), consistent with the PED distribution on the KP1 points. However, the pattern is reversed, with more damage occurring on the ventral right side of the experimental points. If the frequency of damage on opposing ventral edges is influenced by hafting configuration, and/or a twisting action during use, one would not expect perfectly congruent results between the experimental studies and the KP1 points unless the experimental study perfectly replicated past hafting configurations and behaviors, which it certainly did not. Post- depositional processes tend to result in an increased frequency of damage on the ventral right side compared to the ventral left (Table 27), so they cannot be used to explain the increased frequency of PED on the ventral left side.

The PED distribution on the KP1 points lacks the ‘hafting bump’ observed in previous experimental spear tip studies (Schoville and Brown, 2010). However, the experiments conducted here using banded ironstone replicates did not result in a ‘hafting bump’, despite using the same hafting configurations, binding materials, calibrated crossbow, and target (springbok

190 carcass) as previous experiments. Two variables differ between the previous experiments and the one presented here – retouch and raw material. Retouch might influence wear formation on stone tools (Beyries and Plisson, 1998). It may be that retouch on the dorsal surface of the points reduced the fracture frequency on the ventral surface by creating a thicker and more obtuse lateral edge that is less prone to fracture. Additionally, the iron content in banded ironstone could make it a harder raw material less prone to fracture than the quartzite used in previous experiments. Regardless of the reason why there is no ‘hafting bump’ on the experimental banded ironstone points, the main point is that there is not one, even though the experimental points were hafted. The implication is that even if the KP1 points were hafted, there is no reason to expect an obvious ‘hafting bump’ in the ventral edge damage distribution.

6.3.2 Multiple supporting lines of evidence for point function

The presence and frequency of DIFs is consistent with use as hunting weapons. The presence of DIFs is often used as a major line of evidence for tool function as weapon tips, often in the absence of other supporting evidence. However, post-depositional processes can result in low frequencies of DIFs and even non-utilized and non-trampled artifacts can exhibit DIFS, presumably due to the forces exerted during hard hammer percussion (Pargeter, 2011a). Non-use related processes result in very low frequencies of DIFs however, less than ~1% overall, and no more than 4% for certain raw material types under certain conditions (Pargeter, 2011a). The frequency of DIFs on the KP1 points is 13.8%, much greater than that expected for post- depositional processes alone. Even the most conservative calculation of DIF frequency (9.9%), excluding fractures originating from the proximal end and including irregular pointed forms, is greater than expected given the known influence of post-trampling processes. Furthermore, DIF frequencies on the KP1 points are on the high end of the range for MSA sites and MP sites, and within the range for more recent residential sites with points of known function as arrowheads (Table 35). The KP1 frequencies are lower than those observed at kill sites with points of known function and in experimental spear and arrowhead studies.

The experimental work presented here demonstrates that DIFs do form on banded ironstone points used as spear tips, but at a lower frequency than other experimental studies (25% compared to ~40-50%). It is feasible that the iron content in banded ironstone makes it harder than other raw materials and less prone to DIF formation. This is an avenue for future research.

191

If banded ironstone is less susceptible to DIF formation, then the lower KP1 frequencies compared to known kill sites may actually reflect raw material differences rather than differences in site function. Furthermore, because the KP1 material is probably not in primary context, we might be observing a palimpsest of multiple site types, making it difficult to interpret the relationship between DIF frequency and site type.

The KP1 points exhibit a relatively low frequency of minor basal modifications suggestive of hafting. More conclusive evidence for hafting may be provided by future microwear analyses focused on identifying hafting polish/bright spots and binding scars at the level of the individual artifact (Lombard, 2005a; Rots et al., 2011; Van Peer et al., 2008), especially on some of the unpatinated points. Focusing on individual artifacts may also help to elucidate whether the KP1 points may have served other roles, in addition to their role as spear tips, as has been observed for some MP assemblages (e.g., Beyries and Plisson, 1998).

The size and shape of the KP1 points is consistent with experimental studies that hint at an ‘optimal zone’ for point length and width values. In other words, the KP1 points as a group are neither too long nor too wide to function well as spear tips. The experimental study presented here adds further support to the potentiality of the KP1 points as spear tips, demonstrating that point replicates manufactured on the same raw material and with similar dimensions function well as spear tips. Most points were used multiple times while accumulating only minor damage, only two points sustained catastrophic damage. The TCSA and TCSP values for the KP1 points are also consistent with MSA assemblages of points that have been interpreted as spear tips. More work is required to confirm the reality of ‘optimal’ design parameters for spear tips.

Of the six types of evidence used by other researchers to determine MSA point function - three are applied to the KP1 points here and all support their role as hunting weapon tips. The remaining types of evidence were and will be the focus of further collaborative research. A geometric morphometric analysis of the points tested the hypothesis that point symmetry does not change through the reduction sequence of the KP1 points (c.f. Iovita, 2011). The smaller KP1 points are as symmetrical as the larger KP1 points, which supports the spear tip hypothesis (Wilkins et al., in prep). For cutting tools, the expectation is for the smaller points to be asymmetrical compared to the larger points, because they represent later stages of resharpening. In that case, reduction of the point edges would be maximizing edge length (important for

192 cutting) rather than symmetry (important for spear tips). The patinated condition of the KP1 points makes most of them unsuitable for the only other lines of evidence not discussed here, microscopic use wear and residue analysis. However, some of the black chert points that did not patinate may exhibit microscopic use wear traces and provide an opportunity for further collaborative research.

6.3.3 Summary

Several lines of evidence support the hypothesis that the KP1 points were used as spear tips. While it remains possible that some damage on some points was caused by post-depositional processes, or resulted from alternative functions such as cutting, scraping, or piercing tasks, the assemblage-level pattern is one that supports the use of at least some points as weapon tips. Points from MSA and MP sites were often used as weapon tips, and evidence for this behavior can now be pushed back to ~500 ka. The antiquity for hafted hunting technology has important implications for the evolution of the genus Homo that are discussed in Chapter 8.

193

7 Raw Material Analysis

In this chapter, raw material types recognized in hand-specimens and in thin section are described petrographically. I identify potential primary and secondary sources for the raw materials used for lithic reduction at KP1, and quantify raw material variability at secondary sources. Patterns of raw material selection are identified within the KP1 stratum 4a assemblage and compared to a sample from stratum 4b to test for a temporal pattern, and assess whether technological differences between the Fauresmith-designated assemblage and the Acheulean assemblage can be explained by differences in raw material. Lithic raw material studies also provide information on how far and in what direction lithic resources were moved across the landscape, and shed light on hominin foraging strategies. In Chapter 3, I summarized the evidence for predominately local use of raw materials in the ESA, with increased transport distances in the MSA demonstrating long-distance trade and the expansion of social networks. Many MSA sites also indicate that hominins intensified their use of higher-quality fine-grained raw material that required significantly longer search time to acquire. A consideration of raw material foraging strategies represented by stratum 4a permits an assessment of whether there was a significant change compared to the underlying Acheulean assemblage, and whether the foraging strategies are most similar to a typical ESA or MSA pattern. Furthermore, the KP1 lithic assemblages are the first in this region of the Northern Cape that have been subjected to a program of raw material identification, and these data serve as a starting point for future comparative analyses.

7.1 Petrographic Identification of Lithic Raw Material Types

Artifact raw material at KP1 consists mainly of banded ironstone, with low frequencies of black chert, volcanic material, quartzite, quartz, and agate. Banded ironstone (also locally called “jaspilite” or “jasper”), is a sedimentary rock consisting of layered bands of chert or shale and iron-rich magnetite or haematite. Banded ironstone exhibits a high degree of variability, expressed macroscopically in color and structure. Many banded ironstone artifacts from stratum 4a exhibit various shades of patinas including white, grey/blue, and pink. Porat et al. (2010) described the stratum 4a raw material as varied fine-grained rocks, but that designation was before a sample of this material was sectioned. A sample of sectioned flake fragments showed that the majority of lithics were manufactured from either relatively homogenous banded

194 ironstone or homogenous zones of iron-rich chert within the banded ironstone formation. This material is opaque and commonly brown, red-brown, and grey-brown. Black chert differs from the iron-rich zones of chert within the banded ironstone because it is translucent, has a waxy lustre, and does not develop the same kind of white patination. The artifacts made on volcanic raw material are usually heavily weathered. The interior is grey-green in color with barely visible (<1mm) black phenocrysts. The general term ‘volcanic’ is used because of discrepancies between identifications on the geological map (i.e., andesite) and the results of this study (i.e., komatite), discussed below. Quartzite artifacts have characteristic coarse-grained texture and are of a variety of colors. Milky white quartz is rare, but present. Extremely rare examples of white chalcedony/agate were also identified.

To characterize the raw material types, standard petrographic thin sections were prepared from a sample of 14 lithic artifacts. Because this is a destructive technique, only small flake fragments or ‘chunks’ were selected, with permission granted from the McGregor Museum. All lithic artifact characteristics were recorded prior to processing. The sample consisted of 8 banded ironstone variants, 1 black chert, 1 volcanic, 2 quartzite, and 1 quartz. Stephen Wood in the Department of Sciences laboratories at the University of Western Ontario prepared the thin sections. In the absence of chemical analyses, classification of the KP1 raw material is based on textural and structural properties, and observable mineral characteristics. Hand samples and photomicrographs of thin sections are shown in Figure 42 and Figure 43. Table 38 summarizes the macro and microscopic characteristics of each raw material type.

7.1.1 A note on the ‘volcanic’ identification

On the most recent known geological map for this region (1979, prepared by the Geological Survey, Pretoria, published by The Government Printer), the volcanic Hartley and Ongeluk formations are identified as amygdaloidal andesitic lava. More recent works describe the Hartley formation as and basaltic lava (Moen, 2006). In thin section, the archaeological volcanic material exhibits a slightly spinifex texture consisting of long acicular phenocrysts of olivine and

195

Figure 42 Banded ironstone KP1 stratum 4a archaeological hand samples (sectioned flake fragments) and thin sections. A. type 1a, B. type 1b, C. type 1c, D. type 1d, E type 1e, F. type 1f. G. type 1g. Bar in thin sections = 50 μm.

196

Figure 43 Other raw material types KP1 stratum 4a archaeological hand samples (sectioned flake fragments) and thin sections. A. black chert, thin section cross-polarized, B. volcanic, thin section cross-polarized, C. quartzite, thin section cross-polarized, D. quartzite, thin section cross-polarized, E. quartz, thin section cross-polarized. Bar in thin sections = 200 μm.

197

Table 38 Petrographic identification of raw material types in the KP1 assemblage. Type Macroscopic characteristics Microscopic characteristics Banded ironstone colour: brown, red, grey-blue, orange, purple Microcrystalline (~3-10 μm) translucency: opaque silica groundmass with iron luster: dull, iron-rich bands may appear metallic oxides (that differ with texture: microcrystalline respect to size, shape, and structure: banded, mottled, or homogenous concentration between patina: well-developed samples and between different bands or areas of the same sample) Fissures often infilled with crystalline silica Black chert colour: black, dark grey Microcrystalline and translucency: translucent chalcedonic silica luster: dull-waxy groundmass, homogenous, texture: non-clastic, cryptocrystalline Very few and small iron oxide structure: generally homogenous, some banding inclusions and/or mottling patina: not well-developed Volcanic colour: green-blue-grey Spinifex texture consisting of translucency: opaque long acicular phenocrysts of luster: dull olivine texture: aphanitic or porphyritic Some iron oxide inclusions structure: homogenous patina: white-yellow-brown, well-developed Quartzite colour: various, including pink-purple, green-grey, Crystalline quartz with white variable mineral inclusions translucency: translucent including chlorite, feldspar, luster: glassy muscovite texture: granular, medium-grained structure: homogenous patina: not well-developed Quartz colour: white Crystalline-massive quartz, translucency: translucent grains have poorly defined luster: glassy boundaries texture: not granular structure: cleavage planes patina: not well-developed Other includes agates and chalcedonies, very rare

olivine pseudomorphs (Figure 43B), characteristic of komatite. Geological samples collected from outcrops of the Ongeluk Formation (see below) exhibit similar structure. Komatitie is a rare and usually associated with ancient greenstone sequences (McCarthy and Rubridge, 2005:78) . Further investigation that includes chemical analysis is required to identify the volcanic material confidently. For the purposes of this study, the generic term ‘volcanic’ suffices to distinguish it from other available raw material types in the region and link it with potential primary and secondary sources in the region surrounding KP1.

198

7.2 Stratum 4a assemblage composition

7.2.1 Raw material types

Table 39 presents the frequency of the different raw material types in the KP1 stratum 4a assemblage. Banded ironstone is the dominant raw material for all artifact categories. All other raw materials occur in very low frequencies, less than 7.5% overall. For the retouched points, black chert is relatively common (11.1%) compared to the other non-banded ironstone raw material types. This implies that, though black chert was use infrequently, it may have been preferentially used for retouched points when it was used. Volcanic material has relatively high frequencies among the stratum 4a LCTs (15%), implying that hominins may have preferentially manufactured LCTs from volcanic material when it was used. For both the retouched points and LCTs, however, banded ironstone is still the dominant raw material (>80%).

The majority of pieces with cortex that could be defined as either river-rolled (smooth, rounded, pock marks) or outcrop (rough, angular, no pock marks) exhibit river-rolled cortex (89.4%, Table 40). Of the different raw material types, only banded ironstone exhibits some evidence for outcrop exploitation, but at a low frequency (11.6%). Negative evidence for outcrop cortex in other raw material categories could potentially be a factor of small sample size, but other lines of evidence, discussed further below, are consistent with a stronger focus on secondary sources such as stream and river beds, rather than primary sources.

7.2.2 Banded ironstone variability

Ten varieties of banded ironstone were defined to look for patterns of selection within the banded ironstone raw material type. These are normative types identifiable in hand specimens without magnification, defined on the basis of shared attributes such as color and structure. The types are:  1a: mainly homogenous blue/green/brown material (5Y and 2.5 Y hues) with homogenous white-yellow patination, generally fresh edges, may have some banding, but there are thick homogenous zones  1b: mainly homogenous yellow/brown material (10YR hues) with white-yellow patination, generally fresh edges, may have some banding, but there are thick homogenous zones

199

Table 39 Raw material type frequencies in the KP1 stratum 4a assemblage. Sample consists of all bifaces from squares F23, F21, C23, C21, identifiable blades and blade fragments from squares F23, F21, identifiable cores from square F23, identifiable flakes and proximal flake fragments from F23, F21, identifiable retouched points and point fragments from squares F23, F21, C23, C21, identifiable retouched pieces from square F23. Indeterminable identifications excluded. LCTs Blades Cores Flakes Retouched Points Retouched Pieces Total n % n % n % n % n % n % n % Banded Ironstone 16 80.0% 450 96.2% 110 89.4% 638 92.2% 120 83.3% 121 96.0% 1455 92.5% Chert 0 0.0% 5 1.1% 2 1.6% 9 1.3% 16 11.1% 2 1.6% 34 2.2% Volcanic 3 15.0% 3 0.6% 1 0.8% 23 3.3% 3 2.1% 0 0.0% 33 2.1% Quartzite 1 5.0% 8 1.7% 9 7.3% 15 2.2% 4 2.8% 3 2.4% 40 2.5% Quartz 0 0.0% 2 0.4% 1 0.8% 6 0.9% 0 0.0% 0 0.0% 9 0.6% Other 0 0.0% 0 0.0% 0 0.0% 1 0.1% 1 0.7% 0 0.0% 2 0.1% Total 20 468 123 692 144 126 1573

Table 40 Cortex frequencies by raw material type in KP1 stratum 4a. Indeterminable identifications excluded. Banded ironstone Chert Volcanic Quartzite Quartz Total n % n % n % n % n % n % Outcrop cortex 18 11.6% 0.0% 0.0% 0.0% 0.0% 18 10.7% River-rolled cortex 137 88.4% 3 100.0% 2 100.0% 8 100.0% 1 100.0% 151 89.4% Total 155 3 2 8 1 169

 1c: dark brown material with rust or manganese deposits adhering to surface (10YR hues with lower values and chroma than above), sometimes with patination or weathering  1d: burned material that is reddened and exhibits crazing and pockmarks  1e: finely banded material  1f: tiny rounded inclusions that patinate more heavily than the rest of the stone  1g: other types that are less homogenous than a. and b. – mottled, banding, inclusions, various colours, does not fill well in any other category  1h: pieces with higher concentrations of red jasper  1i. pieces with pink patination  1j. highly weathered pieces with very strong white patination

Types 1a, 1b , and 1g are the most common banded ironstone types in stratum 4a assemblage (Table 41). This is true across all artifact categories, with the exception of LCTs, where 1a and 1b are rare. The types 1a and 1b are usually characterized by a mainly homogenous structure, both in hand samples and in cross section (Figure 42). At the microscopic scale, variation between these banded ironstone types can be seen in the size, shape, and density of iron oxide inclusions (Figure 42), which are opaque in regular and cross-polarized light. The different

200 colors of the iron oxide inclusions might represent different iron oxide types – hematite, goethite, magnetite, etc – but geochemical characterization is required to further identify them.

To further explore the influence of raw material characteristics on hominin selection, all lithic artifacts in the stratum 4a sample were assessed for structure and visible flaws. Three types of structure were identified for the banded ironstone artifacts: banded, homogenous, and mottled (Figure 44). Based on informal knapping experience with this raw material, it is easier to control flaking when using the homogenous banded ironstone than when using heavily banded material. Three types of flaws were identified (Figure 45): thin fissures (hair-line cracks), thick infilled fissures (with visible crystal growth), and thick iron-rich bands (dark grey-black bands with metallic luster). Based on informal knapping experience with this material, infilled fissures make it difficult to predict knapping outcomes, and thick iron bands are extremely difficult to knap through. In the KP1 stratum 4a assemblage, lithic artifacts with homogenous structure and no flaws are the most common category at 32.0% (Table 42).

If homogenous, flaw-free banded ironstone has lower frequencies on the landscape than what is observed in the archaeological assemblage, then one could argue that hominins at KP1 were preferentially selecting homogenous, flaw-free banded ironstone. To test this hypothesis, and to examine other aspects of hominin raw material selection and transport, it is necessary to identify and characterize all potential raw material sources near the site of KP1.

7.3 Potential Sources of Raw Material

7.3.1 Primary sources

Primary sources of raw material occur as outcrops and have a fixed location on the landscape. In many regions of the world, water, sand, and geological processes such as volcanic activity, uplift, and erosion can result in landscape changes that cover and expose primary raw material sources, but these processes probably had a minor effect on the landscape during and since hominin occupation at and around Kathu Pan. The landscape around Kathu Pan is not tectonically active. Sedimentation and erosion rates in arid environments such as this region of the Northern Cape are low compared to other environments (Portenga and Bierman, 2011). Water levels certainly fluctuated, but at the scale of analysis conducted here, and given the extent

201

Table 41 Banded ironstone type frequencies in KP1 stratum 4a. Indeterminable identifications excluded. LCTs Blades Cores Flakes Retouched Points Retouched Pieces Total n % n % n % n % n % n % n % 1a 1 6.3% 82 18.3% 24 21.8% 127 19.9% 30 25.0% 33 27.5% 297 20.4% 1b 1 6.3% 154 34.3% 16 14.5% 148 23.2% 39 32.5% 34 28.3% 392 27.0% 1c 1 6.3% 30 6.7% 14 12.7% 46 7.2% 4 3.3% 10 8.3% 105 7.2% 1d 0 0.0% 0 0.0% 0 0.0% 1 0.2% 2 1.7% 0 0.0% 3 0.2% 1e 1 6.3% 2 0.4% 4 3.6% 6 0.9% 3 2.5% 0 0.0% 16 1.1% 1f 1 6.3% 29 6.5% 1 0.9% 33 5.2% 5 4.2% 1 0.8% 70 4.8% 1g 9 56.3% 84 18.7% 40 36.4% 180 28.2% 27 22.5% 35 29.2% 375 25.8% 1h 0 0.0% 9 2.0% 7 6.4% 23 3.6% 5 4.2% 4 3.3% 48 3.3% 1i 0 0.0% 13 2.9% 1 0.9% 15 2.4% 0 0.0% 0 0.0% 29 2.0% 1j 2 12.5% 46 10.2% 3 2.7% 59 9.2% 5 4.2% 3 2.5% 118 8.1% Total 16 449 110 638 120 120 1453

of the outcrops over at least tens of kilometers, the effect of water levels on raw material availability at primary sources was probably negligible. Sand cover would have also experienced shifts over time, but again, the effects of sand cover on primary sources at this scale of analysis are negligible. Water levels and sand cover may have had a more important role in the availability of secondary sources at streambeds in the immediate vicinity of KP1, and this will be discussed further below.

The identification and characterization of primary raw material sources involved infield observations and thin section analysis (Table 43). The Department of Earth Sciences laboratories at the University of Western Ontario thin-sectioned 11 primary source samples collected from road cut exposures. The locations of these samples are marked in Figure 46. Distributions of primary raw material sources are derived from digitized geological maps (Figure 46). Representative hand samples and thin sections are presented in Figure 47 and Figure 48.

7.3.1.1 Banded ironstone

Banded ironstone is available in primary context in the Kuruman Hills to the east of Kathu Pan. These hills belong to the Danielskuil and Kuruman formations of the Ghaap group in the Griqualand West Basin of the Transvaal Supergroup. The nearest marked outcrop relative to Kathu Pan is the northernmost hills of the western flank of the Kuruman Hills near the archaeological sites at Bestwood, 7.3 km away (Chazan et al., 2012b). The archaeological site of Kathu Townlands 6.6 km away is located on top of banded ironstone outcropping, and the site

202

Figure 44 Examples of banded ironstone structure types.

Figure 45 Examples of banded ironstone flaw types.

203

Table 42 Frequency of different banded ironstone structure and flaw types in the KP1 stratum 4a assemblage. Indeterminable identifications excluded. No flaws Fissures Thick infilled fissures Thick iron bands Total n 326 145 23 2 496 Banded structure % 22.9% 10.2% 1.6% 0.1% 34.9% n 455 161 19 635 Homogenous structure % 32.0% 11.3% 1.3% 0.0% 44.7% n 178 92 20 290 Mottled structure % 12.5% 6.5% 1.4% 0.0% 20.4% n 959 398 62 2 1421 Total % 67.5% 28.0% 4.4% 0.1% 100.0%

has been interpreted as quarry location based on the abundance of flaking debris (Beaumont, 2004a). The presence of outcrop cortex on a low frequency of artifacts at KP1 indicates that at least some raw material from primary contexts were transported to the site. The Kathu Townlands site is the nearest known source.

Banded ironstone, in the form of massive and banded ‘jasper’, also outcrops 26 km to the southwest as the Voelwater Formation of the Oliphantshoek Supergroup. The distance of the this outcrop makes it a less likely candidate for the majority of outcrop banded ironstone at KP1, because there are much closer primary sources. Furthermore, the thin section sample from this outcrop also exhibits unique characteristics in thin section (abundant very angular rhomboid-like iron oxide inclusions, Figure 47), that were not observed in other geological or archaeological sample of banded ironstone. If the Voelwater Formation was used as a source of raw material for the KP1 artifacts, it was probably to a much lesser degree than the Danielskuil and Kuruman Formations.

7.3.1.2 Chert

The black chert described here differs from the iron-rich chert in banded ironstone in thin section because of its cryptocrystalline, sometimes chalcedonic texture and lack of iron-oxide inclusions (Figure 43A, Figure 48A). It has been observed as seams within the banded ironstone formation near the archaeological site of Wonderwerk Cave, and is associated with the Danielskuil and Kuruman Formation (Figure 46). Survey in the hills near the Bestwood sites and Kathu Townlands did not result in locating primary sources of black chert closer to KP1, though it is possible that there are nearer primary sources somewhere in the Kuruman Hills that occur at a

204

Table 43 Geological formations near KP1, their location relative to KP1, and macroscopic and microscopic characteristics (based on thin section samples collected for this analysis). Taxonomy based on Moen (2006) and Eriksson et al. (2006). Supergroup Group Subgroup Formation Lithology Distance and Label Macroscopic characteristics Microscopic characteristics direction from on KP1 Figure 3 Griqualand Ghaap Danielskuil banded 7 km east A Banded ironstone Banded ironstone West Basin of Hills and ironstone colour: brown, red, grey-blue, orange, Microcrystalline (~3-10 μm) silica groundmass with iron Transvaal Kuruman with black purple oxides (that differ with respect to size, shape, and Supergroup chert translucency: opaque concentration between samples and between different seams luster: dull, iron-rich bands metallic bands or areas of the same sample) texture: microcrystalline Fissures often infilled with crystalline silica structure: banded, mottled, or homogenous Chert Chert colour: black, dark grey Microcrystalline and chalcedonic silica groundmass, translucency: translucent homogenous, Very few and small iron oxide inclusions luster: dull-waxy texture: non-clastic, cryptocrystalline structure: generally homogenous, some banding and/or mottling Postmasbur - Ongeluk volcanic 21 km east B colour: green-blue-grey Spinifex texture consisting of long acicular phenocrysts g (minor outcrops translucency: opaque of olivine as close as 14 luster: dull Some iron oxide inclusions km), southwest texture: aphanitic or porphyritic 16 km structure: homogenous Olifantshoek - - Gamagara quartzite 13 km south C - - (Mapedi) - - Voelwater banded 26 km southwest D colour: red, Two intermingled structures: (1) microcrystalline ironstone translucency: opaque groundmass with large (~50 um) rhomboid black iron (massive luster: dull oxides and (2) microcrystalline groundmass with and banded texture: microcrystalline laminations of small globular iron oxide (this iron oxide red structure: banded, mottled, or homogenous, is yellow/brown in thin section, red on the slide) “jasper”) semi-circular iron rich inclusions - - Lucknow quartzite/ 32 km southwest F colour: white, blue Crystalline quartz, variability in size (~50 to > 500 um), quartz translucency: translucent variable mineral inclusions including chlorite, feldspar, luster: glassy muscovite texture: granular, medium-grained structure: homogenous - - Hartley volcanic 29 km southwest E - - Volop - Matsap and quartzite 32.0 km F colour: brown, purple, grey, white Crystalline and microcrystalline quartz, variable mineral Brulsand southwest translucency: translucent inclusions including chlorite, feldspar, muscovite luster: glassy texture: granular, medium-grained structure: homogenous

205

Figure 46 Map of primary sources of raw material in the Kathu region showing thin section sample locations. Geological formations and rivers redrawn from 1979 geological map, prepared by the Geological Survey, Pretoria, published by The Government Printer. A. Danielskuil and Kuruman Formations, banded ironstone, B. Ongeluk Formation, volcanic, C. Gamagara Formation, quartzite, D. Voelwater Formation, banded ironstone, E. Hartley Formation, volcanic, F. Lucknow, Matsap, and Brulsand Formations, quartzite.

206

Figure 47 Geological samples and thin sections of primary and secondary sources of banded ironstone. Locations mapped in Figures 5 (primary sources) and 8 (secondary sources). A. Potential primary source at Bestwood in Kuruman Hills, B. Potential primary source at Kathu Townlands. C. Additional sample from Kuruman Hills in eastern flank nearer to Wonderwerk Cave, showing gradation into iron-oxide free chert-like material, D. Sample from Voelwater Formation, E. Sample from potential secondary source at Vermeelsgleete River, F. Sample from Location 1, Gamagara River. All thin sections in normal light

207

Figure 48 Geological samples and thin sections of primary and secondary sources of other raw material types. Locations of samples mapped in Figures 5 (primary sources) and 8 (secondary sources). A. Black chert, possible secondary source at Vermeelsgleete River, thin section in cross-polarized light, B. quartzite from Matsup Formation, thin section in cross-polarized light, C. quartzite from Oliphentsloop River in Langeberg Hills, thin section in cross-polarized light, D, volcanic from Vermeelsgleete River, thin section on left in normal light, thin section on right in cross-polarized light, E. volcanic from Ongeluk Formation, thin section on left in normal light, thin section on right in cross-polarized light.

208

Figure 49 Secondary sources of raw material in the Kathu region showing thin section sample locations and quantitative survey locations. Geological formations and rivers redrawn from 1979 geological map, prepared by the Geological Survey, Pretoria, published by The Government Printer.

209

minimum distance of ~7 km from KP1. The geological samples of black chert for petrographic identification were recovered in secondary context only from the upper Gamagara (Figure 49, Loc 1). Black chert occurs infrequently in the stratum 4a assemblage (Table 39). Of the three chert artifacts with identifiable cortex, all exhibited river-rolled cortex (Table 40), suggesting that outcrops of black chert were not the main source of raw material for stone tool production at KP1. An exception to this generalization that will be discussed below concerns the retouched points.

7.3.1.3 Volcanic

The volcanic material requires further petrographic analysis to identify it confidently as basalt, andesite, or komatite (see above). There are two volcanic formations that outcrop in the Kathu region – the Ongeluk Formation of the Postmasburg Group, and the Hartley Formation of the Oliphantshoek Supergroup (Figure 46). The Ongeluk Formation is exposed between the two flanks of the Kuruman Hills about 21 km east of KP and there are small outcrops as close as 14 km. About 16 km southwest of KP1 there is another outcropping of the Ongeluk Formation. Hand samples and thin sections from the latter exposure of the Ongeluk Formation (Figure 48E) exhibit similarities with the archaeological samples (Figure 43B). Due to accessibility limitations to private and government-owned land, it has not yet been possible to collect geological samples from the Ongeluk Formation outcrops between the Kuruman Hills that are located closer to KP1. The Hartley Formation is exposed 29 km southwest of KP1, and it was not possible to sample that exposure.

Volcanic material makes up a minor proportion of the KP1 assemblage. Two pieces with identifiable cortex exhibited river-rolled cortex (Table 40), suggesting that primary sources of volcanic material were not an important source of volcanic material at KP1.

7.3.1.4 Quartzite

Quartzites of the Oliphantshoek Supergroup make up the Langeberg Hills west of KP1. The nearest point of these hills to KP1 is approximately 26 km southwest. A nearer outcrop of the quartzitic Gamagara Formation is located 13 km south. These quartzites exhibit diverse macro and microscopic characteristics, but with similar mineral inclusions present in the geological and archaeological samples.

210

The distance of the primary sources of quartzite, together with the observation that all of the lithic artifacts with identifiable cortex were river-rolled (Table 40), suggest that primary sources of quartzite were not a major source of lithic material at KP1.

7.3.1.5 Quartz

Milky white quartz similar to the few archaeological specimens in the KP1 assemblage were observed as gravels on top of the Lucknow and Voelwater Formations in the Langeberg Hills. Quartz also occurs as veins in the banded ironstone formation of the Kuruman Hills. Primary sources of significant size were not located in the Kuruman Hills, but quartz pebbles were present in colluvial and alluvial settings there (see below).

Only a single quartz artifact in the KP1 assemblage had identifiable cortex, and it was river- rolled. Primary sources of quartz probably did not contribute significantly to the KP1 lithic assemblage.

7.3.2 Secondary sources

Based on the frequency of river-rolled cortex, secondary sources seem to have played a more important role than primary sources. Stream and river beds are dynamic aspects of the landscape and their location and visibility depends on largely on current weather patterns that affect water levels and sand cover. Figure 49 shows the current location of streams and rivers in the Kathu regions. These waterways are intermittent at best, and many of them may be completely inactive. Running water has not been observed in any streambeds during fieldwork in the dry season. However, the presence of large cobbles (>200 mm) indicates that they were at some point high energy transporters of raw material. Substantial gravel beds are currently visible along the majority of the main Gamagara River (Figure 49).

The location of modern streams and riverbeds may not correspond exactly with their location during the early Middle Pleistocene occupation of KP1. However, based on evidence that landscape around Kathu is relatively stable with little to no recent volcanic activity, and experiences relatively low sediment and erosion rates, it is reasonable to assume that the major drainage basins today correspond well with the major drainage basins during the early Middle Pleistocene, even if the exact location of streams and river channels do not. Today, the headwaters of the Gamagara River begin in the Kuruman Hills and on the dolomitic Maramane

211

Dome just south of Sishen Mine. The Gamagara flows north where it meets the Kuruman River. The Veermulsleegte tributary, which also originates in the Kuruman Hills, meets the Gamagara about 30 km north of Dibeng, though there are also minor tributaries from the Kuruman Hills that today occur within 3.5 km of KP1. From the west, the Oliphentsloop, Ga-mmatshephe, and Dooimansleegte tributaries originate in the Langeberg hills and flow into the Gamagara (Figure 49).

7.3.2.1 Raw material availability at secondary sources

Characterizing raw material availability in secondary sources can be methodologically challenging. In a recent study, Goldman-Neuman and Hovers (2012) examined raw material selectivity of secondary sources represented by the Oldowan sites of A.L. 894 and A.L. 666 in the Makaamitalu Basin, Hadar, Ethiopia. They identified conglomerate less than 400 m away as potential source for raw material and analyzed a random sample of 208 clasts > 5cm in maximum length from two areas within this formation. The variables they recorded were rock type, phenocryst size and abundance, groundmass texture, metrics (length, width, thickness, mass), cobble shape based on six categories (highly angular, angular, sub-spheroid, spheroid, elongated disks, rounded disks), and qualitative ranks of raw material knapping quality. Petrographic thin-section analysis was used to confirm rock type assignments, which included mainly of different types of volcanic rocks, though chert and quartz were also present. The frequency of the variables in the conglomerate was then compared to the archaeological frequencies, to demonstrate which variables influenced Oldowan selectivity. I employ a similar methodology here, but with attention to variables relevant to the locally available raw materials, and using a more structured sampling strategy and are greater number of sampling locations. My sampling locations are located much farther from the archaeological site, and as discussed below, are not meant to represent the exact location where KP1 hominins were selecting raw material, but to stand as proxies for the kinds of areas that might be exploited.

To examine raw material availability at secondary sources near KP1, the Gamagara drainage basin was divided into three zones (Figure 50). The concept of ‘zones’ serves to address the issue that the streams and riverbeds are not in the exact location today as they were in the early Middle Pleistocene. Zone A includes the headwaters of the Gamagara River and the eastern tributaries, where water flows through the Danielskuil, Kuruman, and Ongeluk Formations - the primary

212 sources of banded ironstone, black chert, volcanics, and quartz. Zone B includes the western tributaries that originate in the Langeberg Hills carrying quartzites and volcanics to the Gamagara River. Zone C includes the main Gamagara River after it crosses the quartzitic Gamagara Formation.

7.3.2.2 Methods

A quantitative sampling strategy was carried out in each of these three zones. Zone A includes three sampling locations (Loc 1, 3 and 6). One location was sampled in Zone B (Loc 2). Two locations were sampled in Zone C (Loc 4 and 5). The exact locations for sampling depended on largely on visibility (where extensive gravel beds were observable), and access (where we did not require or could arrange permission).

All cobbles within a 25 by 25 cm sampling square with a maximum length of at least 50 mm were documented (Figure 51). Multiple sampling squares were set up at each location, with distances of 10 to 20 m between each square, depending on the extent of the gravel exposure and vegetative cover. Lithology, structure, flaws, and metrics were recorded for all materials within each square. Each cobble was broken open using a geological hammer so that characteristics of the interior could be recorded. Data were collected infield; material was not collected. Each cobble was photographed. The sampling occurred over a period of four days and with the assistance of three students.

7.3.2.3 Results

Table 44 presents the raw material type frequency at each location. The presence/absence of specific raw material types is consistent with expectations based on the distribution of primary rock sources relative to modern drainage patterns.

7.3.2.3.1 Zone A

Location 1 is at the upper Gamagara between the two flanks of Kuruman Hills. The sample transects the modern riverbed as well as a gravel bed that probably represents an ancient river terrace. Cobbles of banded ironstone are the most abundant type, but black chert, volcanics, and white quartz are present. Quartzite is not present.

213

Figure 50 Secondary source zones. See text for discussion.

214

Figure 51 Methods for quantification of secondary sources of raw material. A. Overview of gravel bed at Location 1. B. Close-up of 25 cm by 25 cm sampling square. C. Banded ironstone cobble on scale. D. Measuring between sampling squares.

Location 3 is a gravel bed located 3.7 km east of the modern Gamagara. There is a low to medium density of non-diagnostic lithic artifacts consistent with an ESA or MSA designation on the surface of the gravel bed, and the site had been previously recognized during an impact assessment in the area (David Morris, personal communication). Some of the gravel bed has been mined by the landowners. Location 3 exhibits the highest density of cobbles out of all of the sampled locations (Table 44). Cobbles of banded ironstone are the most abundant type, but black chert is present. Volcanics, quartzite, and white quartz are not present. The absence of volcanics implies that the tributary that resulted in the gravel deposit had headwaters in the western flank of the Kuruman Hills and did not cut through the Ongeluk volcanic exposure between the

215

Table 44 Raw material type frequencies at six sampled secondary source location in three

zones (see text for discussion).

)

2

)

2

Location Zone Numberof squares Samplearea (m Numberof cobbles5cm > (n/m Density Banded ironstone Black chert Volcanic Quartzite Quartz

Loc 1 A 7 1.75 91 52.0 85.7% 1.1% 11.0% 0.0% 2.2% Loc 2 B 4 1 20 20.0 0.0% 0.0% 0.0% 95.0% 5.0% Loc 3 A 3 0.75 89 118.7 95.5% 4.5% 0.0% 0.0% 0.0% Loc 4 C 5 1.25 71 56.8 23.9% 2.8% 32.4% 22.5% 18.3% Loc 5 C 5 1.25 54 43.2 53.7% 1.9% 22.2% 11.1% 11.1% Loc 6 A 4 1 47 47.0 100.0% 0.0% 0.0% 0.0% 0.0% Zone A Loc 1, 3+6 14 3.5 227 64.9 92.5% 2.2% 4.4% 0.0% 0.9% Zone C Loc 4+5 10 2.50 125 50.0 36.8% 2.4% 15.2% 17.6% 28.0%

two flanks. A single cobble of agate was encountered during informal survey across this gravel bed, suggesting that agate, though extremely rare, does have a local origin.

Location 6 is near the archaeological site of Bestwood 1 (Chazan et al., 2012b). Mining operations there had stripped off several meters of sand, exposing artifacts lying flat on top of a buried gravel layer composed entirely of banded ironstone (Table 44).

7.3.2.3.2 Zone B

Location 2 is located at the upper Oliphentsloop in the Langeberg Hills. Only quartzite and rare white quartz are available at this location (Table 44). Based on the distribution of primary sources of within Zone B, the lower Oliphentsloop is likely to also contain Ongeluk and Hartley volcanics and banded ironstone from the Voelwater formation.

7.3.2.3.3 Zone C

Locations 4 and 5 are located at gravel beds along the main Gamagara River. There are numerous gravel beds identified along the Gamagara. Based on the distribution of primary sources, one would expect the main Gamagara, after crossing the quartzitic Gamagara Formation and collecting the tributaries from the Langeberg Hills, to contain banded ironstone, black chert,

216 volcanics, quartzites and quartz. The sampling strategy at Locations 4 and 5 confirm this (Table 44).

7.4 Raw Material Selection at KP1

7.4.1 Raw material type availability and frequency at secondary sources

Kathu Pan 1 is located in Zone A, where tributaries carry banded ironstone, black chert, volcanics, quartz, and very rarely, agate. Modern tributaries in Zone A run within 3.5 km of KP1, though it is possible that during the early Middle Pleistocene, KP1 hominins had access to streams and gravel deposits even closer to KP1 that are now buried by sand.

Within Zone A, the gravel bed at Location 3 is probably the best approximation of raw material availability for the secondary sources exploited by hominins in the immediate vicinity of KP1. Based on where it is situated, the gravel bed at Location 3 may have been deposited by an ancient channel of a stream that runs just 3.5 km west of KP1 (Figure 49). The upper part of this stream could not be accessed for survey. The modern Veermulsleegte, 5 km east of KP1, was assessed for survey, but sand cover was too extensive and there were no visible gravel deposits.

Quartzite is not available within Zone A, and it would not have been at any point in the past. The nearest source of quartzite is the Gamagara River bed in Zone C 11 km west of KP1. All other raw material types used in the KP1 stratum 4a assemblage are also found in Zone C.

The lithic artifacts at KP1 from secondary sources were most likely manufactured on materials foraged from both within Zone A and Zone C. Exploitation in Zone B is less likely, because of its location beyond Zone C, but remains a possible source. Based on a comparison of the raw material type frequencies in Zones A and C to the KP1 assemblage (Figure 52), it seems more likely that Zone A served as the main supply of secondary source raw materials. However, the frequency distribution for the combined locations in Zone A is significantly different from the KP1 frequency distribution (χ=20.3, df=3, p<0.001), indicating that some selection of raw material types occurred. The biggest difference is that volcanic material is underrepresented in the KP1 Stratum 4a assemblage compared to its availability within Zone A, indicating that there was preferential selection against volcanics for lithic manufacture during the stratum 4a occupation. It is known that hominins must have also accessed Zone C, because that is the nearest secondary source of quartzite.

217

Figure 52 Comparison of raw material type frequencies in KP1 stratum 4a to secondary sources in Zone A and Zone B.

7.4.2 Banded ironstone variability

To address the question of whether banded ironstone varieties with certain structure and without flaws were preferentially selected from secondary sources, the frequency of these traits in the KP1 assemblage (Table 42) can be compared to the frequency of these types in the secondary sources within Zones A and C (Table 45). Figure 53 presents a summary comparison. Homogenous structure is overrepresented, and banded structure is underrepresented in the KP1 stratum 4a assemblage compared to the availability of these types of banded ironstone in Zones

218

Table 45 Frequency of different banded ironstone structure and flaw types at each of the six secondary source sampling locations. No flaws Fissures Thick Thick iron Weathering Total infilled bands interior fissures Loc 1 Banded structure n 9 8 2 14 2 35 % 12.0% 10.7% 2.7% 18.7% 2.7% 46.7% Homogenous n 17 2 2 21 structure % 22.7% 2.7% 2.7% 0.0% 0.0% 28.0%

Mottled structure n 7 7 5 19 % 9.3% 9.3% 6.7% 0.0% 0.0% 25.3% Total Loc 1 n 33 17 9 14 2 75 % 44.0% 22.7% 12.0% 18.7% 2.7% 100.0% Loc 3 Banded structure n 10 11 3 7 1 32 % 11.9% 13.1% 3.6% 8.3% 1.2% 38.1% Homogenous n 11 3 5 19 structure % 13.1% 3.6% 6.0% 0.0% 0.0% 22.6%

Mottled structure n 11 14 6 2 33 % 13.1% 16.7% 7.1% 2.4% 0.0% 39.3% Total Loc 3 n 32 28 14 9 1 84 % 38.1% 33.3% 16.7% 10.7% 1.2% 100.0% Loc 4 Banded structure n 2 2 3 1 8 % 11.8% 11.8% 17.6% 0.0% 5.9% 47.1% Homogenous n 3 1 1 5 structure % 17.6% 5.9% 5.9% 0.0% 0.0% 29.4%

Mottled structure n 2 2 4 % 11.8% 0.0% 11.8% 0.0% 0.0% 23.5% Total Loc 4 n 7 3 6 1 17 % 41.2% 17.6% 35.3% 0.0% 5.9% 100.0% Loc 5 Banded structure n 7 4 1 2 2 16 % 24.1% 13.8% 3.4% 6.9% 6.9% 55.2% Homogenous n 3 1 4 structure % 10.3% 0.0% 3.4% 0.0% 0.0% 13.8%

Mottled structure n 2 5 2 9 % 6.9% 17.2% 6.9% 0.0% 0.0% 31.0% Total Loc 5 n 12 9 4 2 2 29 % 41.4% 31.0% 13.8% 6.9% 6.9% 100.0% Loc 6 Banded structure n 7 12 20 39 % 14.9% 25.5% 0.0% 42.6% 0.0% 83.0% Homogenous n 3 1 4 structure % 6.4% 0.0% 0.0% 2.1% 0.0% 8.5%

Mottled structure n 2 2 4 % 0.0% 4.3% 0.0% 4.3% 0.0% 8.5% Total Loc 6 n 10 14 23 47 % 21.3% 29.8% 0.0% 48.9% 0.0% 100.0% Total All Locations n 94 71 33 48 6 252 % 37.3% 28.2% 13.1% 19.0% 2.4% 100.0%

219

Figure 53 Comparison of structure and flaw type frequencies in KP1 stratum 4a to secondary sources in Zone A and Zone B. A. Structure. B. Flaws.

A and C. Chi-square tests between KP1 and Zone A (χ=39.3, df=2, p<0.001), and KP1 and Zone C (χ=11.8, df=2, p=0.003) indicate that the frequency distributions or structure types are significantly different. Also, artifacts without flaws are overrepresented in the KP1 assemblage compared to the availability of cobbles without flaws at secondary sources in Zones A and C. Artifacts with thick infilled fissures and thick iron bands, are underrepresented in the KP1 assemblage. Chi-square tests between KP1 and Zone A (χ=11355.1, df=3, p<0.001), and KP1 and Zone C (χ=7494.2, df=3, p<0.001) indicate that the frequency distributions of the flaw types are significantly different. Together, the evidence indicates that KP1 hominins were selecting banded ironstone cobbles from secondary sources with homogenous structure and no major flaws.

7.4.3 Nodule size

The maximum length of complete flakes, blades, and cores represents the minimum length of the original nodule. In order for the cobbles available at secondary sources to have served as the main source for lithic reduction, there must exist cobbles that exceed the maximum length of the archaeological remains. For all raw material types except for chert, the maximum cobble size at

220 secondary sources exceeds the maximum length of the archaeological remains (Figure 54). There are numerous chert retouched points that exceed the maximum dimensions of chert cobbles at secondary sources. All chert nodules in the sampled squares were less than 70 mm in maximum length, and there were no large cobbles of chert discovered during informal surveys of secondary sources in the Kathu region. Thus, it is possible that sources of chert not available at nearby secondary sources were exploited to manufacture points, 4 of which exceed 70 mm in maximum length.

7.4.4 Raw material selection for blade production at KP1

If certain raw material types were selected for blade manufacture as opposed to flake manufacture at KP1, then there would be different raw material type frequencies between blades and Levallois flakes in the Stratum 4a assemblage. The frequencies are similar (Table 46) and a Chi square test indicates that the differences are not statistically significant (χ2=2.312, df=4, p=0.679).

There is evidence for selection of homogenous banded ironstone for blade production. Table 47 presents the percent frequency of different structure types observed on blades and Levallois flakes. More blades than Levallois flakes were manufactured on homogenous banded ironstone. A chi square test indicates that the Levallois flakes and blades exhibit statistically different frequencies (χ2=22.5, df=2, p<0.001).

Blades were also preferentially manufactured on banded ironstone without flaws (Table 48). Levallois flakes were more frequently manufactured from nodules with fissures or thick infilled fissures. The difference in flaw type frequencies is statistically different between blades and Levallois flakes (χ2=12.6, df=2, p=0.002).

7.4.5 Direction of flaking axis with respect to banding

An additional attribute observed on the KP1 lithic artifacts was the direction of flaking axis with respect to banding. This attributed was recorded for each lithic artifact within the sample that was manufactured on banded ironstone with a banded structure. It was decided to measure this attribute after comments made by Kyle S. Brown during the replication of the retouched points for the experimental study described in Chapter 6. During knapping, it was noted that the ‘grain’

221

Figure 54 Comparison of KP1 stratum 4a artifact length (complete pieces) and maximum length of cobbles at secondary sources for each raw material type. Among the black chert pieces, retouched points are highlighted as red points.

222

Table 46 Comparison of percent frequency of raw material types for blades and Levallois flakes in stratum 4a assemblage at KP1. Blade Levallois Flake n=468 n=104 Banded Ironstone 96.2% 97.2% Black chert 1.1% 0.0% Volcanic 0.6% 0.9% Quartzite 1.7% 0.9% Quartz 0.4% 0.9%

Table 47 Comparison of frequency of banded ironstone structure types between blades and Levallois flakes in stratum 4a KP1 assemblage. Blade Levallois Flake n=461 n=107 banded 31.5% 30.8% homogenous 56.2% 42.1% mottled 12.4% 27.1%

Table 48 Comparison of flaw type frequencies between blades and Levallois flakes in the stratum 4a assemblage of KP1. Blade Levallois Flake n=463 n=105 No flaws 70.6% 61.0% Fissures 27.6% 33.3% Thick infilled fissures 1.7% 5.7%

and banding direction influenced knapping decisions and that it was more difficult to remove blades against the grain than with the grain Figure 55 presents the definitions for the terms used here: with, through, diagonal, and against. For cores, the direction was recorded for the last removal.

Of all the raw material categories, blades and retouched points (which were often manufactured on blade blanks, see above) exhibit the highest frequencies of pieces that are with the banding and the lowest frequencies of pieces that are against the banding (Figure 56). A chi-square test comparing the frequencies of the different directions between blades and flakes indicates that

223

Figure 55 Definitions for direction of flaking axis with respect to banding.

differences are statistically significant (χ2=301.8, df=3, p<0.001). In other words, when banded material was used for blade production, blades were usually removed parallel to the direction of the banding, rather than against it.

224

Figure 56 Comparison of flaking direction with respect to banding for different artifact types in KP1 stratum 4a assemblage.

7.4.6 Stratum 4a (Fauresmith-designated) vs. Stratum 4b (Acheulean)

Do the raw material foraging strategies represented by stratum 4a represent a change compared to the earlier Acheulean occupation represented by stratum 4b? To address this question, a comparative sample from one square unit (n=758, square C20, levels 280-300 cm) of stratum 4b was analyzed for raw material type. The frequencies are similar between Stratum 4a and 4b (Table 49), with banded ironstone dominating both strata, suggesting that raw material type selection and transport differed little between stratum 4a and 4b hominins.

225

The 4b sample does exhibit slightly higher frequencies of quartzite and quartz (Table 49). A chi- square test indicates that the frequency distributions are significantly different (χ2=27.1, df=4, p<0.001). While quartzite also plays a minor role in the stratum 4b assemblage, the higher frequency does suggest that stratum 4b hominins selected quartzite for reduction and transport to KP1 more often than stratum 4a hominins. Given that the nearest source of quartzite is 11 km from KP1, this might have implications for hominin mobility, which will be discussed further below.

There are differences between stratum 4a and 4b with respect to what kind of raw material was used for LCT manufacture. As presented above (Table 39), the stratum 4a bifaces were manufactured on banded ironstone (80.0%), volcanic material (15.0%), and quartzite (5.0%). In the stratum 4b sample analyzed here, all 14 LCTs and LCT fragments were manufacture on banded ironstone. The different frequencies are not significant (χ2=3.2, df=3, p=0.366), and one quartzite handaxe was observed in the unanalyzed stratum 4b collection. Still, there is a suggestive pattern that the stratum 4a LCTs may have been manufactured on a greater diversity of raw materials and in that way, and in other respects (Porat et al., 2010), differ from the stratum 4b LCTs.

To assess additional raw material characteristics, a smaller sample from stratum 4b (n=155, square F21, levels 160-180 cm) was subjected to the same analysis as the stratum 4a sample (Table 49). Of the stratum 4b artifacts with identifiable cortex (n=17, 10.9%) all have river- rolled cortex. Secondary sources were the main source of raw material for stratum 4b, but given the small sample size, it is not possible to rule out some input from primary sources.

The results indicate that stratum 4b hominins also selected homogenous banded ironstone with no flaws (Table 50). The frequencies of different structure types are not significantly different between stratum 4a and 4b (χ2=3.8, df=2, p=0.152). The frequencies of different flaw types are significantly different (χ2=14.8, df=3, p=0.002), with the stratum 4b assemblage exhibiting slightly higher frequencies of artifacts with no flaws and lower frequencies of artifacts with fissures (Figure 53). This data might suggest that selection against flawed banded ironstone was even greater for stratum 4b hominins than stratum 4a hominins.

There are significant differences in the direction of flaking with respect to banding between the stratum 4a and stratum 4b assemblages (Figure 57). There are significantly higher frequencies of

226

Table 49 Comparison of raw material type frequencies in stratum 4b sample compared to Stratum 4a sample. Stratum 4b sample consists of all lithic artifacts from square C20, levels 280-300 cm. Stratum 4a assemblage includes all bifaces from squares F23, F21, C23, C21. Identifiable blades and blade fragments from squares F23, F21. Identifiable cores from square F23. Identifiable flakes and proximal flake fragments from F23, F21. Identifiable retouched points and point fragments from squares F23, F21, C23, C21. Identifiable retouched pieces from square F23. Indeterminable identifications excluded. Stratum 4b Stratum 4a Banded Ironstone 679 89.6% 1455 92.5% Black Chert 14 1.8% 34 2.2% Volcanic 16 2.1% 33 2.1% Quartzite 40 5.3% 40 2.5% Quartz 9 1.2% 9 0.6% Other 0 0.0% 2 0.1% Total 758 1573

Table 50 Frequency of different banded ironstone structure and flaw types in the KP1 stratum 4b assemblage. Sample from square F21, levels 160-180 cm. No flaws Fissures Thick infilled fissures Total Banded structure n 46 14 0 60 % 29.7% 9.0% 0.0% 38.7% Homogenous structure n 64 7 2 73 % 41.3% 4.5% 1.3% 47.1% Mottled structure n 17 4 1 22 % 11.0% 2.6% 0.6% 14.2% Total n 127 25 3 155 % 81.9% 16.1% 1.9% 100.0%

pieces that were flaked with banding than against banding in the stratum 4a assemblage compared to the stratum 4b assemblage (χ2=11.6, df=3, p=0.009).

The maximum length for all stratum 4b artifacts fall below the maximum length of cobbles observed at secondary sources. There are no stratum 4b chert artifacts greater than 30 cm in maximum length, and these pieces could have been detached from the small chert cobbles that are available in the stream and riverbeds of Zones A and C. This observation contrasts that for stratum 4a, where it appears that, based on their size, that at least some of the chert artifacts were manufactured from primary sources, or from cobbles much larger than those available in Zones A and C.

227

Figure 57 Comparison of flaking direction with respect to banding between strata 4a and 4b.

7.5 Discussion

The stratum 4a assemblage at KP1 was manufactured mainly on cobbles of banded ironstone collected from nearby secondary sources. Today these secondary sources run within 3.5 km of KP1, but there may have been ancient stream channels exposed within the immediate vicinity of KP1 that are today covered by sand. Stratum 4a hominins preferentially selected banded ironstone cobbles that were homogenous in structure and had no major flaws. Even though the overwhelming pattern is one focused on immediate raw material resources, there is some evidence that raw material from more distant (but still local) sources were transferred to KP1. Based on the presence of outcrop cortex on some pieces, primary sources of

228 banded ironstone must have been exploited. The nearest primary source of banded ironstone is 6.6 km away at the Kathu Townlands archaeological site. This site is extremely rich in lithic artifacts and has been described as an Acheulean quarry site (Beaumont, 2004a). Recent informal surveys at Kathu Townlands identified the presence of blades and Levallois cores, which is consistent with utilization by Fauresmith hominins. Further work at Kathu Townlands is required to establish the role of the site and whether it may be penecontemporaneous with occupations at KP1. Additional artifact scatters have been located on the top of the two adjacent hills just east of Kathu Townlands near the archaeological site of Bestwood 1. Survey of these hills revealed that very high quality homogenous flawless banded ironstone is abundant at these locations and would have been a good source of raw material for Middle Pleistocene hominins. Research at these sites is ongoing.

The presence of quartzite artifacts with river-rolled cortex indicated that the main Gamagara River (Zone C) was also utilized as a raw material source and at least some material from there was transported to KP1. The nearest point of the modern river is 11 km from KP1. Substantial gravel terraces are located all along the Gamagara and would have served as good sources of all the raw material types that are observed at KP1.

There is also evidence that primary sources of black chert were utilized; there are numerous black chert artifacts that are significantly larger than the available cobbles in secondary sources. Black chert occurs as seems in the Danielskuil and Kuruman Formations that start to outcrop at Kathu Townlands 6.6 km from KP1. Outcrops of black chert were not observed at that location, nor at the Bestwood sites, however, suggesting that the cherts may have a source slightly beyond these locations.

At KP1, there is currently no evidence for long distance (>100km) transport of raw materials, which characterizes some younger MSA assemblages.

7.5.1 Foraging strategy and site function

The pattern of raw material use during the Fauresmith-designated occupation of KP1 indicates that multiple sources were utilized. Many of these are within the daily foraging radius of modern hunter-gatherers. However, the archaeological record is mute on how resources were transported within this foraging area, and it is not possible to determine whether this represents an embedded

229 strategy of raw material procurement that is linked to other subsistence activities (Binford, 1979), or whether raw material was collected during logistical forays for the purpose of raw material acquisition, or some other type of foraging strategy without modern analogues.

The richness of the KP1 assemblage and the abundance of flaking debris indicate that the location was occupied for longer periods, or was revisited numerous times. If modern hunter- gatherer foraging strategies are used as an analogue for KP1, this could suggest that KP1 served as a residential camp (Binford, 1980). The frequency of DIFs on the retouched points is consistent with known habitation sites in the Holocene (Chapter 6), however, the interpretation of DIF frequency depends on a variety of factors on which research has just begun. Because of the unknown variables that affected site formation at KP1, it remains difficult to assert site function confidently. Since the KP1 material is in secondary context, it could represent a palimpsest of different site locations and different site types if the area around Kathu Pan was used for different purposes through time. Further research is required to establish how the site of KP1 fits into landscape use patterns for foraging hominins.

7.5.2 Comparing strata 4a and 4b

The small sample from stratum 4b indicates that there was no major shift in raw material use between the Acheulean and Fauresmith-designated occupations at KP1. During both occupations, there was preferential use of banded ironstone, with infrequent use of black chert, volcanics, quartzites, and quartz. Based on the presence of river-rolled cortex in both strata, stream and river beds were the main source.

The stratum 4b Acheulean assemblage does exhibit a higher frequency of quartzite, the nearest source of which is 11 km from KP1. These observations may have implications for hominin mobility and raw material selectivity. The Acheulean occupants may have preferred quartzite for particular tasks, preferentially selecting it during special logistical forays to Zone C. Alternatively, the increased frequency of quartzite could reflect an embedded foraging strategy that brought the stratum 4b occupants to Zone C more often than the stratum 4a occupants. Interpretations about the variability between strata 4a and 4b should be regarded with caution given the small sample from stratum 4b. The main conclusion is the assemblages were manufactured on the same raw materials, and further work is required to establish the nature and meaning of the very minor differences in raw material types.

230

There also is evidence that the occupants represented by stratum 4a intentionally exploited the natural banding in banded ironstone to produce elongated products. In stratum 4a, the axis of flaking is more often with and through the banding for blades than against it, and the opposite pattern is observed on flakes. One might speculate that the stratum 4a toolmakers recognized that banding could influence endproduct dimensions, and they prepared cores in a manner that would most easily permit blade removals. Blades are not observed in the Acheulean assemblage and detached pieces more often exhibit banding that goes against the axis of flaking, suggesting that the Acheulean occupants did not recognize this characteristic, and/or did not care to make elongated products.

Stratum 4a at KP1 is not associated with increased raw material selectivity, which is a pattern observed for many MSA sites (see Section 3.5). This contrasts with evidence from the Kapthurin Formation, where the early MSA (i.e. ‘transitional’) site of Koimilot exhibits a greater range of fine-grained lavas compared to the nearby Acheulean sites (Tryon et al., 2005). The KP1 raw material pattern is consistent with evidence from Casablanca, North Africa, however, where the same raw material types are utilized in the Acheulean and ‘transitional’ assemblages (Raynal et al., 2001). Together, the archaeological record implies that there is no universal pattern of raw material use across the ESA-MSA boundary. Some sites document clear change in raw material exploitation patterns, some do not.

Local factors would certainly play a role in whether raw material exploitation patterns change or not. For the Kapthurin sites, Tryon et al. (2005) suggest that the increased use of fine-grained raw material in the early MSA reflects a need for sharp edges that are potentially less durable. Based on informal experiment knapping and using the material, banded ironstone flakes have sharp and durable edges. Further research on the mechanical properties of banded ironstone might help to elucidate whether its unique properties explain the different pattern at KP1 compared to the Kapthurin sites. Banded ironstone is also the most abundant raw material available on the landscape surrounding KP1 and high quality, flawless, homogeneous cobbles of appropriate size are available locally and are easy to locate. The acquisition of banded ironstone, which appears to have been the most preferred raw material for both the Acheulean and Fauresmith-designated occupations at KP1, would not require significant search time.

231

The most important implication for the lack of raw material differences between the Acheulean and Fauresmith-designated assemblages at KP1 is that the technological changes reflected by the Fauresmith-designated assemblage occurred independently of raw material differences. For the most part, the same raw materials were exploited for the core and handaxe-based Acheulean occupation, as for the blade and flake-based Fauresmith-designated occupation. However, the same raw materials were exploited in different ways. The stratum 4a occupants seem to have exploited the natural bedding in the banded ironstone to produce elongated products, but the Acheulean occupants did not. If the Fauresmith-designation for the stratum 4a assemblage is accepted than raw material characteristics do not appear to explain the technological differences between Fauresmith and Acheulean assemblages and Humphreys’ (1970) argument to the contrary is not supported. The data regarding direction of banding do support the idea that certain raw materials might be more conducive to blade production than others (Clark, 1980:46). However, while it could explain why blade production occurred in and around the Kuruman Hills and not in all parts of Africa ~500 ka, it cannot explain why systematic blade production did not occur prior to ~500 ka in this area.

7.5.3 Summary

Results are consistent with hominin use of multiple local sources of raw material, including stream cobbles in the immediate vicinity of Kathu Pan, outcrops ~7 km from the site, and stream cobbles at a minimum distance of 11 km. Hominins preferentially selected banded ironstone cobbles with homogenous structure and no flaws. Based on comparisons with a sample from the underlying stratum 4b Acheulean assemblage, the stratum 4a assemblage does not exhibit major changes in the kinds or quality of raw material exploited. If minimum raw material transport distances are taken as a rough proxy for foraging radius, and similarities between the stratum 4a and 4b occupations of Kathu Pan are taken to reflect similarities in how hominins exploited the landscape, then one could argue that the Fauresmith and Acheulean hominins employed similar foraging strategies. Raw material use during the stratum 4a occupation of Kathu Pan, which is focused on abundant locally available resources, is consistent with the general pattern for ESA foraging strategies, and differs from evidence at some MSA sites that show longer transport distances and intensification of certain high-value materials. The lack of change across the stratum 4b/4a boundary at KP1 may also have important implications for hominin social behaviors that will be discussed further in Chapter 8.

232

8 Discussion 8.1 Summary of results

The analysis of stratum 4a at KP1, which has been dated to ~500 ka, reveals systematic blade production, multiple Levallois reduction strategies, the manufacture of points, and the use of points as hafted spear tips. Together, the technological and functional evidence indicates that that MSA technologies date to ~500 ka, approximately 200 ka older than generally recognized. LCTs are also present in the assemblage, but the analysis cannot confidently assert the nature of the association. The LCTs may be intrusive, because they are a minor component of the assemblage and are not technologically linked through chaînes opératoires to the remainder of the assemblage, unlike many Acheulean sites where LCTs are manufactured on large flakes from Levallois cores (Sharon, 2009). The KP1 LCTs are also more weathered than the majority of other lithic artifacts and are concentrated in the lowermost levels of stratum 4a.

The technological changes documented in stratum 4a are not associated with a major change in raw material use or selection, based on a comparison with the underlying stratum 4b Acheulean assemblage. All raw materials at KP1 have local sources and are within the daily foraging radius of modern hunter-gatherer bands, approximately 8-12 km. Most raw material came from an immediate secondary source, no more than 4 km away. Strata 4a and 4b show similar selection patterns with respect to raw material type and banded ironstone variability. There are two implications of this finding. First, raw material characteristics do not explain the observable differences between the Fauresmith-designated and Acheulean assemblages, and changes in raw material use are ruled out as a casual factor for technological change. Second, if raw material use and discard tracks foraging strategies and landscape use, stratum 4a documents little change in these aspects compared to the Acheulean.

8.2 Synthesis and Interpretation

8.2.1 The onset of the MSA by 500 ka

KP1 provides the earliest known evidence for many MSA innovations, but this finding is not unexpected or contradictory in relation to current evidence. To push the onset of the MSA to ~500 ka is reasonable, based on chronometric evidence. Archaeological assemblages dated to the

233 early Middle Pleistocene record are uncommon and there have been few opportunities for assessing hominin technological behavior during this time. In Chapter 3, I summarized the available chronological data for the late ESA and early MSA and highlighted three observations: 1. KP1 and the Kapthurin sites are the only well-documented and reported African contexts that we know have archaeological units dated to between ~500 and 300 ka. Like KP1, the assemblages at the Kapthurin sites contain LCTs, and blades, Levallois, or points (Johnson and McBrearty, 2010), and are what I label as ‘transitional’ assemblages. 2. Acheulean assemblages lacking blades, Levallois, and/or points that we have reliable chronometric information date to ~500 ka or more (including Cave of Hearths, Herries, 2011). Many reported ages younger than ~500 ka for other Acheulean assemblages are most likely minimum age estimates (Herries, 2011). 3. After 300 ka, there are both MSA assemblages (lacking LCTs) and ‘transitional’ assemblages. LCTs may persist until ~125 ka in some areas.

The limits of chronometric dating, together with weak archaeological visibility during the early Middle Pleistocene probably explain why MSA innovations have not been reported from contexts >300 ka prior to now. The paucity of early MSA sites in southern Africa could in part be due to the reality that higher sea stands before OIS 5 cleared out coastal cave sites of older materials, and coastal MSA occupations during lower sea level times are currently submerged on the coastal shelf (Fisher et al., 2010; Marean, 2010). The interior of South Africa has so far yielded few stratified and datable sites. In East Africa, there are fewer reported MSA sites associated with datable volcanic tuffs than ESA sites, and MSA research has generally been overshadowed by earlier time periods. Constructing a chronology for the early MSA, which is beyond the limits of , is hampered by the reliability of alternative dating techniques. Paleomagnetic dating has limited utility for sites younger than 780 ka, which all have normal polarity. Useful methods for this time period, including U-series, ESR, and OSL have only undergone significant improvements within the last two decades (see Jacobs and Roberts, 2008). As radiometric dating methods improve and are applied to more sites, we may find that the African Acheulean in its strictest sense ended around 500 ka, even if LCT tools themselves persisted for many more thousands of years.

234

There are multiple explanations for why assemblages between 500 and 300 ka contain both LCTs and MSA technologies. Perhaps even after the adoption of new technologies focused on flake and blade production and/ point production, hominins continued to manufacture and use LCTs, and the archaeological contexts from which we recover these tool types represent the accumulation of tools used for different extractive tasks. Alternatively, we could be uncovering small-scale palimpsests of the toolmaking behaviors of different taxonomic groups or species – those that adopted the new MSA technologies and those that did not. Or, the palimpsests could be larger in scale, with older artifacts (LCTs) that have accumulated on the surface becoming buried with younger artifacts. I suggest that the latter could be the explanation at KP1, but every ‘transitional’ assemblage will not necessarily have the same explanation.

8.2.2 The status of the Fauresmith Industry is still contentious Despite the use of the term to describe several assemblages in southern Africa, the Fauresmith industry has never been clearly defined and assemblage characteristics used to identify the Fauresmith have been inconsistent and often contradictory. Nonetheless, the term was and probably still is a convenient label for assemblages with technological elements consistent with both the ESA and MSA. We do know that several assemblages in southern Africa have been recovered that contain tool types consistent with both an ESA and MSA designation. Some of these assemblages are from excavated contexts that stratigraphically situate them between more typical ESA assemblages and MSA assemblages. Many of these sites have been designated as Fauresmith, and the term seems to have been used as shorthand for assemblages with LCTs and Levallois, blades, and points, even though no thorough definition of the Fauresmith was ever offered. Most of the Fauresmith-designated assemblages are concentrated in central South Africa, but there are some occurrences, though arguably more debatable, in Limpopo, Eastern Cape, Western Cape, and Zambia. KP1 provides the oldest age estimate for a Fauresmith- designated assemblage, dating to at least 435 ka (Porat et al., 2010). OSL age estimates at two other sites, Bundu Farm and Rooidam 1, suggest an age of at least 150 ka (Kiberd, 2006; Szabo and Butzer, 1979). The duration of the Fauresmith could potentially be ~300 ka, but current chronological information is inconclusive with respect to the younger age limit.

At this point, there are no detailed typological and technological descriptions of Fauresmith- designated assemblages. Except on a superficial level, we do not know how one Fauresmith- designated assemblage compares to another. We do not know the meaning of the association of

235

LCTs with MSA tools – whether this represents a range of tool-making activities of one group of hominins, or whether we are observing archaeological palimpsests. We do not know the meaning of the superficial similarities between the sites or the meaning of variability between sites. We do not know the true distribution, whether these Fauresmith-designated assemblages are actually contemporary, or how long the so-called Fauresmith Industry persisted. While the lack of secure knowledge regarding the Fauresmith poses problems for studying it, it also highlights the necessity for thorough descriptions of Fauresmith-designated assemblages. The analysis presented here, parts of which have been published elsewhere (Wilkins and Chazan, 2012), has resulted in the first detailed technological description of a Fauresmith assemblage. Because detailed descriptions of other Fauresmith assemblages are not available, the data here can only serve as an early step towards defining and understanding the significance of the Fauresmith. If the methods utilized here are applied to other Fauresmith-designated assemblages, robust inter- site comparisons could lead to a clearer definition and evaluation of the Fauresmith Industry.

8.2.3 Reconsidering the significance of technological and morphological change in the Middle Pleistocene

Here, I consider KP1 in evolutionary context and in light of the interpretations of early Middle Pleistocene hominin behavioral and morphological change presented in Chapter 2.

8.2.3.1 Hominin speciation and dispersal

Like most Stone Age sites, KP1 yields no hominin remains and consequently, exactly what hominin manufactured the assemblage is a question that will never be confidently answered. Even in cases where archaeological material and hominin remains come from the same stratigraphic units, there is uncertainty on whether the hominin deposited there necessarily was the maker of the stone artifacts. The deposition of the KP1 assemblages coincides with a period of time when H. heidelbergensis s.l. inhabited the African landscape. H. ergaster/erectus had been manufacturing LCTs and Acheulean assemblages for ~1.0 Ma. After 600 ka, hominins with larger brains and more human-like morphological features appear and by 500 ka, assemblages with blades, Levallois technology, and points appear. At KP1, there is evidence that points were hafted and used as spear tips.

Some current models of hominin evolution in the Middle Pleistocene emphasize the correlation between H. helmei ~300-259 ka and the origin of Mode 3 technologies (Foley and Lahr, 2003;

236

Foley and Lahr, 1997; McBrearty and Brooks, 2000). The chronological and technological evidence from KP1 is inconsistent with this interpretation. The lithic assemblage at KP1 reflects a clear technological break from the strategies used in the Acheulean and predates the earliest evidence for the group of hominins sometimes lumped together as H. helmei (or Group 2 in a gradational scheme, McBrearty and Brooks, 2000). The fossil record is incomplete, and it is possible that new discoveries and advancements in chronometric dating may find that we are underestimating the antiquity of H. helmei. Conversely, H. heidelbergensis s.l. could be responsible for the technological innovations characteristic of Mode 3 technologies in Africa.

In Africa, localities where H. heidelbergensis s.l. specimens have been recovered near or in association with lithic artifacts give no more than equivocal support to the idea that the origins of Mode 3 could be linked to H. heidelbergensis s.l.. Table 1 in Chapter 2 summarizes the archaeological associations with each of the African specimens. Two sites could suggest a link between H. heidelbergensis s.l. and MSA or ‘transitional’ assemblages. Excavations at Broken Hill (Kabwe), where the holotype of H. rhodesiensis was recovered, yielded stratified deposits of Acheulean, Sangoan, and MSA assemblages (Clark, 1959a). It remains unknown how these deposits correlate to the original context of the 1921 discovery of the hominin remains, however, Clark (1959a:223) notes that within ‘Bone Cave’, where the hominin remains were supposedly recovered, only MSA-type artifacts are present. The Kapthurin specimens are recovered from the Middle Silts and Gravels Member (K3) ~500 ka, which also contains handaxes, Levallois technology, and blades (Deino and McBrearty, 2002; Johnson and McBrearty, 2010; McBrearty and Tryon, 2006; Tryon, 2006; Tryon and McBrearty, 2006; Tryon et al., 2005). Nevertheless, most African sites are actually more suggestive of a link between H. heidelbergensis s.l. and Acheulean technology. The Tigenhif (Ternifine) remains are associated with Acheulean technology (Geraads et al., 1986). In the Middle Awash, the was recovered from a sedimentary unit that also contained Acheulean handaxes. The Middle Awash Homo specimens are associated with assemblages designated as Oldowan and Acheulean, the coexistence of which might reflect different activity facies within the Acheulean (Clark et al., 1994). “Finely- made” handaxes sometimes likened to the Fauresmith industry were recovered from the Cutting 10 site located near the location where the Elandsfontein cranium was recovered (Klein, 1988). The Ndutu cranium is probably associated with Acheulean handaxes (Clarke, 1990; Mturi, 1976).

237

This evidence can be interpreted many ways. In overview accounts of hominin evolution, H. heidelbergensis s.l. is generally linked to Mode 2 (e.g. Foley and Gamble, 2009; Foley and Lahr, 2003), and the evidence just presented can be used to support that, because handaxes are present at or near all localities with H. heidelbergensis s.l.. Alternatively, there could be a temporal component to the associations (i.e., Acheulean assemblages with earlier specimens, “transitional” assemblages with later specimens), but the quality of the available chronological data makes it difficult to establish whether this is true. On the other hand, most of these associations described here are extremely problematic and together they may tell us little about hominin-technology relationships.

In addition to demonstrating that the appearance of H. helmei and Mode 3 technologies is not coeval, the KP1 assemblage poses another problem for the speciation and dispersal model proposed by Foley and Lahr (Foley and Lahr, 2003; Foley and Lahr, 1997). KP1 indicates that the origins of Mode 3 technology in Africa predate that in Europe by ~200 ka. There is no apparent correlation between the shift to Mode 3 technology in Africa and the shift to Mode 3 technology in Europe and simple dispersion model cannot explain the current pattern. Supposing the earliest Mode 3 assemblages in Africa in the early Middle Pleistocene ~500 ka were manufactured by H. heidelbergensis s.l., and the same species is present in Europe, then the European counterparts used a different technology. Based on current evidence, Mode 3 technology does not appear in Europe until ~300 ka. In Europe, there are robust associations between ~500 ka H. heidelbergensis-designated specimens and Acheulean technologies at sites like Sima de los Huesos and Boxgrove. The appearance of Mode 3 technology in Europe could still be attributed to population dispersal from Africa, but if this happened, it may have happened after hominins in Africa were using Mode 3 technologies for ~200 ka. The ability to evaluate this model is further complicated by the debate about whether the Sima de los Huesos hominin remains are younger than 500 ka and should actually be classified as H. heidelbergensis or H. neanderthalensis (Stringer, 2012). In either case, current evidence cannot be resolved into a simple model using the current classification schemes linking early Middle Pleistocene technology (i.e. Modes 2 and 3) and morphology (i.e. H. heidelbergensis s.l.) across the Old World.

While in some ways KP1 seems to complicate the relationship between biological change and technological change for the Middle Pleistocene, in others it adds clarity. The innovations

238 represented by the KP1 stratum 4a assemblage (blade production, Levallois, technology, and hafted spear tips) reflect behaviors shared with our sister lineage, Homo neanderthalensis. Genetic studies date the Neanderthal-human divergence to the early Middle Pleistocene, either 700-800 ka (Noonan et al. 2006, Green et al. 2010), 500 ka (Green et al. 2006, but cf. Wall and Kim 2007), or 400 ka (Endicott et al. 2010). The KP1 assemblage could predate the divergence, be coeval with it, or post-date it. Either way, 800-400 ka is an important time for the divergence of the Neanderthal and modern human lineages and it is likely that the behaviors represented by the KP1 assemblage roughly approximate our shared capacities. The rough temporal correlation of the appearance of blade, Levallois, and hafted spear technologies with the divergence of the Neanderthal and human lineages gives a parsimonious explanation for similarities between some MSA assemblages attributed to modern humans and some Middle Paleolithic assemblages attributed to Neanderthals. The end product of this correlation is not, however, a simple pattern of speciation and dispersal for the Old World, as discussed above. This is probably in part due to the limitations we impose on ourselves when we try to classify fossil specimens and archaeological assemblages using a sparse and poorly dated record. The real situation involves more variables than two types of archaeological assemblages (represented by our dichotomy of Mode 2 and Mode 3 assemblages) and four populations of hominins (represented by the species H. heidelbergensis s.l., H. helmei, H. neanderthalensis and H. sapiens).

8.2.3.2 Modern human behavior

The KP1 assemblage has implications for the ‘gradualist’ model for the origins of modern human behavior (McBrearty and Brooks, 2000). As mentioned above, KP1 indicates that a shift toward Mode 3 technologies occurred ~200 ka before the appearance of H. helmei (or ‘Group 2’ hominins in a gradistic scheme) based on the current fossil record. It implies that many of the cognitive abilities represented by the appearance of Mode 3 technologies may have also been shared with H. heidelbergensis s.l.. It also indicates that there may have been an even longer period (~400 ka?) for the accumulation of traits associated with behavioral modernity. If the MSA represents the expansion of a shared body of knowledge that accrued over time as suggested by the gradualist model for modern human behavior, than this shared body of knowledge has great time depth, and crossed at least two potential ‘speciation’ events. Alternatively, one could use the behavioral similarities to suggest close phylogenetic relationships between all African Middle Pleistocene hominins.

239

In many ways, the KP1 assemblage shows consistency with the gradualist model (and other models) for modern human origins with respect to what it does not demonstrate. There are clear differences between the KP1 assemblage and more recent MSA assemblages. KP1 does not document the kind of changes in raw material use – intensive selection or long distance transport of fine-grained raw materials – that characterize MSA assemblages in the Late Pleistocene. Nor is there evidence for worked bone tools, microliths, long-distance projectile technology such as darts or arrows with assisted leverage, or symbolic items such as incised pieces or representational images. Despite many technological similarities between the ~500 ka KP1 assemblage and Late Pleistocene MSA assemblages, there are still important distinctions.

8.2.3.3 Behavioral variability and adaptation

Shea’s (2011b) argument for ‘human-like’ behavioral variability can probably be pushed back to ~500 ka. By this time, lithic assemblages in Africa contain elements of Modes 1-4. The KP1 assemblage itself contains elements diagnostic of Modes 1-3. Blade production at KP1 is not prismatic, which is a defining characteristic for Mode 4 technology. However, at roughly the same time in the Kapthurin Formation, hominins were manufacturing blades using different core reduction strategies than at KP1 (Johnson and McBrearty, 2010). Also, in the Levant ~400 ka, hominins used prismatic core Mode 4 reduction strategies to produce blades (Shimelmitz et al., 2011). There was variability in the technological strategies of early Middle Pleistocene hominins. With respect to blade core reduction strategies, there is no diachronic pattern of change in the way that cores were organized, prepared, and exploited. Instead, characteristics come and go in the archaeological record of the African and Levantine Middle and Late Pleistocene. Variability in blade technology during this time could reflect behavioral adaptations to local conditions, and this kind of technological flexibility is consistent with what modern humans do. The archaeological record of the early Middle Pleistocene is inconsistent with Shea’s (2011a) argument that there is a correlation between the appearance of the earliest anatomically modern humans and modern-behaving people ~200 ka. The kind of lithic variability that Shea emphasizes has greater antiquity than 200 ka.

The KP1 evidence can also be considered with respect to the new Modes defined by Shea (in press) and summarized in Table 2. KP1 contains elements consistent with Modes A (stone percussors), C (pebble cores/non-hierarchical cores), D1 (retouched flake-tools), D2

240

(backed/truncated flakes), D3 (burins), E1 (large cutting tools), F1 (preferential bifacial hierarchical cores), F2 (recurrent bifacial hierarchical cores), and G1 (platform cores). There are no G2 blade cores of the type described by Shea (in press), but there are recurrent bifacial hierarchical cores used for blade production, and these differ from the recurrent bifacial hierarchical cores used for flake production. The co-occurrence of these Modes in the KP1 assemblage ~500 ka is inconsistent with a couple of patterns observed by Shea (in press) for the Levantine Paleolithic record. In the Levantine archaeological record, recurrent bifacial hierarchical cores with preferential removals (F2) only appear after ~250 ka, but they seem to have greater antiquity in Africa. Unifacial hierarchical cores (G1, what I called ‘Large Surface Exploitation’ cores) are also present in the KP1 assemblage, even though they are rare before ~250 ka in the Levant (Shea, in press). The implications of this observation could be that the shift to smaller flake tools over the use of heavy core tools occurred earlier in southern Africa than in the Levant, reflecting either an earlier development of the kind of cognitive capacities required for technologically-mediated adaptive strategies (i.e., behavioral variability), or the existence in Africa of some sort of environmental impetus ~500 ka for a shift to smaller flake tools that was absent in the Levant until ~250 ka.

The KP1 assemblage may represent a new degree of behavioral flexibility and a greater range of behavioral responses to different environmental conditions compared to earlier periods. Foley and Lahr (2003) argued that Mode 3 marked the beginning of human levels of variation because it resulted in a diversity of flakes that can be shaped into a diversity of tool forms. The KP1 assemblage may represent a new degree of behavioral flexibility and a greater range of behavioral responses to different environmental conditions. Potts (1998) pointed out the last ~700 ka experienced increased climatic variability compared to earlier time periods -- the fluctuations in global temperature are more extreme and they occur over shorter time periods. The origins of Mode 3 technologies might reflect an adaptive response by hominins to this increased climatic variability. Fluctuating climatic conditions may have selected for hominin cognitive capacities that encouraged innovation so that hominins could occupy diverse environmental conditions. The correlation between increased climatic variability in the early Middle Pleistocene and the origins of Mode 3 is loose, but suggestive. However, further research is required to solidify the correlation and establish whether climatic variability is the best explanation for the technological shifts that characterize Mode 3 technologies.

241

8.2.3.4 Cognition

The fossil record shows that there was a fairly significant increase in brain size among African hominins ~600-400 ka (Rightmire, 2004). Part of what makes human cognition unique is our ability to hold in attention multiple tasks and conduct goal-oriented behaviour. The concept of 'working memory' has been used to highlight this capacity (e.g. Wynn and Coolidge, 2011). Evidence at KP1 for hafted spear tips may indicate an enrichment of working memory capabilities ~500 ka. The manufacture of hafted technologies is one type of behaviour that requires working memory because it requires the collection, preparation, and combination of different kinds of resources -- wood, stone, and binding material. Wynn and Coolidge argue that simple hafting procedures that would require only a single day’s work do not reflect the kind of 'enhanced' working memory that multiday hafting procedures would, and that multiday hafting procedures are only represented in the archaeological record ~70 ka based on evidence for complex resin recipes (Wadley et al., 2009). It is not clear, however, why adding one or two additional ingredients to resin would add an entire day’s work or indicate such an increase in cognitive capacities. Also, I would argue that even 'simple' hafted spears could represent multiday procedures. Resins or other materials used as adhesives could take several days to dry, and controlled application of heat may be required to process and dry the resin. Regardless, the construction of hafted spears could reflect increased capacities for working memory in hominins compared to previous non-hafted Mode 2 technologies. Wynn and Coolidge suggest that simple hafted spears are closer to the ‘maintainable’ side of the maintainable-reliable continuum for technology, especially compared to multi-component bows and arrows. However, compared to wooden spears without lithic tips, or unhafted handaxes, hafted spears would fit closer to the reliable side of the continuum. Also, many factors influence whether modern hunter-gatherers employ reliable or maintainable technologies, and many modern hunter-gather groups depend on tool kits that are not complex (Oswalt, 1976). Perhaps the more complex technologies of later hominins represent more 'enhanced working memory' capacities, but there are other explanations for more complex technologies in the later MSA, that I will discuss below.

Another way to approach hominin cognition relevant to technological change is the concept of ‘constructive memory’. Ambrose (2010) argues that the human capacity for imagining future scenarios and planning for them is more important than working memory, which is focused on immediate tasks. Hafting represents an advance in hominin constructive memory because it

242 involves the completion of multiple subgoals (shaft manufacture, point manufacture, resin manufacture) and final assembly occurs later. It also represents a substantial amount of investment ahead of time with the final goal of securing game in the future. Evidence for hafted spear tips at KP1 indicates a cognitive shift in the Homo lineage that may have permitted more complex goal-oriented behavior. If one also accepts the proposed linkage between composite technology and grammatical language (Ambrose, 2001; Ambrose, 2010), then grammatical language may have been in place by ~500 ka.

8.2.3.5 Social networks and ‘demographic packing’

Foley and Gamble (2009) suggested that one key behavioural change associated with the origins of hafted spears could have been a shift in community and family organization, where larger communities maintained relations even when separated by time and distance. They identify this as 'exploded' fission-fusion dynamics. New hunting technologies may have improved foraging returns and resulted in increased populations and 'demographic packing', but for the most part, KP1 remains mute on the issue of population expansion and demographic packing. Based on a single site it is impossible to ascertain whether there was population increase in the Northern Cape ~500 ka. There are methods for estimating population size. For example, space-time density maps based on the number and distribution of contemporary sites across western Europe were used to estimate population sizes in the Upper Paleolithic (Bocquet-Appel et al. 2005). The frequency and distribution of sites in and near the Kuruman Hills that are penecontemporaneous with KP1 are currently unknown variables. Further survey is required to map out the localities and technological nature of the multiple lithic scatters through the region.

If there was a significant population expansion in the Kuruman Hills and surrounds ~500 ka, then we would have two expectations for the archaeological record of stratum 4a. First, there may be evidence for changes in the types of raw materials used compared to the earlier occupation, because larger populations could result in increases in territoriality and impact the ability of hominins at KP1 to access certain raw material sources. Second, population expansion could potentially stress the availability of preferred raw material types in the immediate vicinity of the site, and result in evidence for increased lithic raw material conservation and/or the use of less abundant or lower-quality materials. The raw material study presented here (Chapter 7.4) indicates that raw material use between the Acheulean and stratum 4a assemblage is nearly

243 identical, and thus, there is no evidence for population expansion ~500 ka at KP1. However, the ability to address the issue of population expansion is limited by the nature of raw material abundance and distribution in the Kuruman Hills, as well as the design of the current study. The question of population expansion in this region during the Middle Pliestocene is one that warrants further investigation.

8.2.3.6 The social brain hypothesis

Because neocortex size and group size are correlated in primates, it may be reasonable to argue that an increase in brain size in the Homo lineage reflects an increase in group size as well, and that larger brains were selected for to deal with the costs of living in larger groups (Dunbar, 1998; Dunbar, 2003; Gamble et al., 2011). If this logic is applied to the Middle Pleistocene, then increases in brain size associated with the appearance of H. heidelbergensis s.l. suggest that group size also increased in the early Middle Pleistocene. As with population size, group size is a difficult variable to observe in the archaeological record and with the limited amount of information available for the early Middle Pleistocene in Africa, this argument is difficult to properly evaluate.

There are some established quantitative methods for determining group size and home range using the archaeological record. The ethnographic record has been used to develop models relating site area to the number of individuals that occupied the site (Naroll, 1962; Wiessner, 1974). This method cannot be applied to KP1, because it is in secondary context and the horizontal distribution of the artifacts is not related to hominin discard patterns. Furthermore, there are conceptual problems with assuming a priori that Plio-pleistocene archaeological localities represent more than the accumulation of multiple discard events at central places on the landscape (Binford, 1981).

Grove (2010) uses the spatial distribution of sites within ‘site clusters’ to estimate group size, and this would be a useful method for estimating group size at Kathu Pan if an extensive survey program was conducted to characterize the location and distribution of sites in and around the Kuruman Hills. Grove’s (2010) group size estimates are based on the range area (estimated from the distances between sites) and “population density” (estimated from hominin body mass indices). Grove presents results relative to the current discussion. Applying this method to the ‘site clusters’ at Olduvai (~1.5 Ma), (~1.5 Ma), Boxgrove (~500 ka, roughly

244 contemporary with KP1), and Paris Basin (), Grove (2010) found that Boxgrove predates any major shift in group size, demonstrating values similar to Olduvai and Koobi Fora. This is interesting in light of the social brain hypothesis, because it refutes the notion that increases in brain size during period between ~1.5 Ma and 500 ka correlate with increases in group size.

We can also consider the KP1 evidence in the model put forth by Gamble et al. (2011). They suggested the lack of significant technological change during the period between ~1.5 Ma and 400 ka may indicate that changes in ‘emotional’ resources (the immaterial aspects of hominin behavior) were more important during this period than material resources. However, KP1 seems to indicate that there were significant changes in the way that material resources were exploited by 500 ka. Hominins at KP1 manufactured blades, prepared cores using a diversity of Levallois strategies, retouched blades and Levallois flakes into points, and used these points to tip wooden spears. None of these behaviors are observed in the underlying Acheulean occupation or at any other known sites > 500 ka. At the few other known sites in Africa and the Levant that are penecontemporaneous with KP1, there are also blades, Levallois technology, and/or points. These new technological strategies might have a wider distribution than is currently recognized because of the problems with archaeological visibility. The end of this ‘movement’ ~600-400 ka as defined by Gamble et al. (2011) was actually witness to huge changes in the way that lithic material resources were exploited. Hafted spear tips may have also impacted hunting strategies – the kinds of prey that were targeted, the way that prey was targeted, who and how many individuals were involved, as well as the success and fail rates, and thus, this ‘movement’ was also witness to changes in the way the meat resources were exploited. The technological changes marked by KP1 ~500 ka could have had important implications for how hominin groups distributed these meat resources, with important implications for group dynamics, cooperation, sharing, and reciprocity. In this way, evidence for hafting at KP1~500 adds further support to Barham’s (2010) suggestion that the origins of hafting ‘fixes’ what might called “Dunbar’s dilemma”, or the apparent disjunction between encephalization in the early Middle Pleistocene and significant behavioral change. The archaeological record at KP1 does not confirm however, that these significant behavioral changes were necessarily associated with increases in group size or interaction.

245

8.2.3.7 Cooperation and ultrasociality

In Chapter 2, I presented the model put forth by Hill et al. (2009) focused on the evolutionary development of uniquely human levels of cooperation and social learning. Based on their conclusions, at some point in the archaeological record of the Middle Pleistocene we would expect archaeological evidence for an increased reliance on hunting and high quality food resources such as large game. We should also see evidence for shared intentionality, if possible, and cumulative cultural capacities in the way that stone tools are manufactured.

KP1 documents an important innovation in hunting technologies ~500 ka. Based on ethnographic evidence of modern hunter-gatherers, the use of spears is generally directed toward large-bodied animals and used to wound or kill prey that is in a disadvantaged position (Churchill, 1993). A disadvantaged position can be forced on the prey by driving them into water, mud, or other impediments that would allow a hunter to gain extra time and access. Disadvantaging is dependent upon physiographic features for success, requiring natural or man- made barriers, and in the Northern Cape, natural features like narrow streambeds, or other bodies of water such as pans, could serve as barriers. Disadvantaging and ambushing with spear tips is also strongly associated in the ethnographic record with cooperative drive techniques (Churchill, 1993). The new hunting technology represented by KP1 ~500 ka may also signal new hunting strategies with impacts on how hominin groups organized their hunting efforts. One could argue that cooperative hunting requires shared intentionality -- the ability to participate in collaborative activities with shared goals and coordinated action roles. This could reflect higher-order intentionality among early Middle Pleistocene hominins compared to earlier ones, and is consistent with the suggestion that estimates of frontal lobe volume for early Middle Pleistocene hominins indicate levels of intentionality approaching those in modern humans (Dunbar, 2003).

Lithic tipped spears ~500 ka could have also impacted hominin life history patterns. Stone tips probably increase a spear’s penetration and cutting ability, and thus its killing power, potentially reducing the risk of an unsuccessful hunt. Reducing daily variance in foraging may have profound life-history changes associated with the amount and regularity of resources adults can contribute and result in reduced juvenile mortality (Kaplan, et al. 2000). Experimental work suggests that stone-tipped armatures require fewer trials than untipped spears to produce fatal hemorrhaging (Hughes, 1998). Therefore, stone-tipped technology reduces the need for

246 prolonged proximity to dangerous prey. This could reduce adult mortality and increase the average adult lifespan. Increasing the length of the juvenile period, higher female fertility, and pair-bonded cooperative breeding all may be explained in part by reduced variability in successful resource capture (Hill, et al. 2009).

Boyd and Richerson (2005:16) suggest that hafted spears represent cumulative culture change. Hafted spears are the combination of multiple innovations to the lithic point and the shaft. A hafted spear is something unlikely to result solely from individual learning in the course of one individual’s lifetime. Rather, social learning mechanisms passing through multiple generations are required to explain the regular manufacture of points and their use as armatures on spears, and this technology builds on aspects of Mode 2 technology – flake production, platform and dorsal surface preparation, shaping and retouch. Wood-working could also have origins in the Acheulean based on plant residues recovered from the surface of Acheulean tools at Peninj, Tanzania (Dominguez- Rodrigo et al., 2001). Also, hunting large game, which is typically the kind of prey targeted with spears among modern hunter-gatherer groups, requires high levels of knowledge and skill that would have been passed down through generations. Meat consumption was an important aspect of hominin adaptation prior to the origins of lithic-tipped spears, and this new type of hunting technology could reflect an improvement to previously learned hunting strategies.

Presumably, KP1 hominins learned these technological strategies via cultural transmission from others in their social group. There is no archaeological evidence, however, that this social group necessarily extended beyond the family. Nonetheless, the practice of big game hunting is suggestive of cooperation beyond the level of the family, because large packages of high-quality resources, like meat from large game, are generally shared beyond the nuclear family in modern hunter-gatherer societies (Kaplan et al., 2000).

Boyd and Richerson (2005) posits that at some point in the Pleistocene, humans developed the psychological machinery required to participate in interfamilial institutions, because large cooperative groups out competed smaller groups and genes favoring prosocial behavior were selected for. KP1 does shed some light on this hypothesis. If hafted hunting technology reflects a shift to larger game, or a shift toward more reliable access to large game, then it might also reflect increased sharing and cooperation with nonkin, marking the beginning of the kind of

247 prosocial behaviors that form the foundation for interfamilial institutions sensu Boyd and Richerson (Boyd and Richerson, 2005). In other words, hominins with the psychological machinery for sharing meat, and hominin groups that cooperated among themselves with respect to how high-quality large package food resources were distributed, would have been selected for. The evidence for hafted spear tips at KP1 ~500 ka, together with evidence for big game hunting at ~500 ka at Boxgrove (Roberts and Parfitt, 1999), ~400 ka at Schoningen (Thieme, 1997), among other sites, may situate the origin of this kind of prosocial behavior in the early Middle Pleistocene. Hunting and/or early access to meat certainly has greater antiquity than the early Middle Pleistocene (Braun et al., 2010; Domínguez-Rodrigo et al., 2005; McPherron et al., 2010), but the appearance of new hunting technologies in the Middle Pleistocene could signal the increased importance of hunting or a shift in the way that hunting was accomplished (Marean and Assefa, 1999).

8.2.3.8 The ratchet effect and imitation

Among extant primates, the human capacity for cumulative culture change is unique. Evidence at KP1 for multiple innovations compared to the Acheulean occupation involves the way that flakes are produced, the kinds of flakes that are produced, and hafted hunting technology. These are all aspects of technological behavior that reflect cumulative culture change. These innovations could be coeval with significant neurological changes that permit certain enhanced capacities for social learning, but the behaviors themselves are not likely innate, instinctual, or genetic. There is significant variability across time and space in the ways that lithic technology is used to exploit and extract resources. There is evidence for working memory, constructive memory, higher- order and shared intentionality, and perhaps nonkin cooperation in the early Middle Pleistocene. All these traits are shared with modern hunter-gatherer populations, and based on similarities between the Middle Stone Age and Middle Paleolithic archaeological records, with Neanderthals as well.

Imitation is considered an integral component of social learning in humans, where imitation differs from emulation because it focuses on process, the way of doing something, the exact manner in which the end goal is accomplished. In contrast, emulation is focused on the end goal and an emulator essentially learns the process of getting there on their own (Tennie et al., 2009). I suggest that the KP1 assemblage demonstrates that both imitation and emulation were

248 important aspects of learning lithic technology, but emulation may have played a larger role during the Middle Pleistocene than in earlier and later time periods. I make this argument based on the diversity of reduction strategies used to manufacture one kind of endproduct at KP1. Different kinds of blanks, blades and flakes, were retouched into points. Points manufactured using convergent flaking that did not require retouch were not retouched. Diagnostic impact fractures occur on all these types of points, indicating that they may have all been used as spear tips. There were diverse processes leading to the same end goal; multiple chaines operatoires were used to manufacture the same tools. KP1 is not the only MSA site that seems to document this kind of pattern; diversity characterizes much of the MSA.

Technological diversity in the MSA seems to exceed the amount of diversity observed between and within Mode 2 assemblages. The persistence of the handaxe form for at least 1 million years across much of the Old World is often cited as evidence for a high degree of technological homogeneity. There is certainly more variability within and between Acheulean assemblages than text books gives them credit for, but there is still a fairly robust pattern for increased behavioral variability represented by the origins of Mode 3 assemblages compared to Mode 2 assemblages. Perhaps the cognitive and/or psychological mechanisms that promoted imitative behaviors in H. erectus (Shipton, 2010), were relaxed in the context of fluctuating environmental conditions known to characterize the last ~700 ka (Potts, 1998).

Mathematical models of culture change demonstrate that diversity increases when innovation rates are high (Kandler and Laland, 2009). In these models, there are two types of innovation – those represented by independent invention and those represented by improvement through modification, and these two types of innovation show slightly different dynamics. Independent invention generally supports higher levels of diversity than improvement through modification (Kandler and Laland, 2009). Diverse core reduction strategies in the Middle Pleistocene could represent multiple episodes of independent invention and demonstrate the increased importance of this kind of learning compared to earlier time periods. If the diversity during this time also indicates a reduced importance for improvements through modification, then one could explain why cumulative culture change does not appear to pick up the pace until the Late Pleistocene.

The behavioral strategies of early Middle Pleistocene hominins differ from those of more recent time periods in three main ways:

249

1. There are no long-distance projectiles (Shea, 2006), or ‘complex’ tools, defined here as those consisting of more than three parts, or technounits (Oswalt, 1976). 2. There appears to be a lesser degree of formal variation across time and space. Point styles in the later part of the MSA exhibit regional variability on a continent-wide scale (Clark 1982, McBrearty and Brooks 2000), and time– restricted patterning (Wadley, 2001). There is general consensus that the spatial and temporal pattern of MSA point styles implies that artifact variability in the MSA is associated with cultural patterns (Clark, 1992; Foley and Lahr, 2003; Foley and Lahr, 1997; McBrearty and Brooks, 2000).The interpretation that variability in point form signifies so-called ‘regional traditions’ is based on the assumption that point form can be used to communicate messages about ethnic or group affiliation (Wiessner, 1983), and/or that points could serve as important symbolic objects in regional trade networks (Wilkins, 2010). It remains to be shown that the lack of patterning in formal variation in the early MSA is not simply a product of the patchy archaeological record, however, based on current evidence it seems probable that lithic material culture in the early Middle Pleistocene did not serve to communicate ethnic affiliation or to create and sustain large social networks. 3. Third, there is little evidence for the use of symbolic resources in the Middle Pleistocene. Pigment processing may be one important exception, known to date to more than ~240 ka based on evidence from the Kapthurin sites (McBrearty et al., 1996) and Twin Rivers, Zambia (Barham, 2000). There are utilitarian explanations for ochre in the MSA, for example, as an agent in compound resin recipes (Wadley et al., 2009), but it appears that the reddest were selected and discarded at MSA sites (e.g., Marean et al., 2007; Watts, 2002), supporting a symbolic interpretation as a skin, hair, or object colorant. Pigment is present in the KP1 stratum 4a assemblage (Beaumont, 1990b; Beaumont, 2004a). Preliminary inspections by Ian Watts confirmed evidence for scraping and/or grinding on some of the ochre pieces and research on these pieces is ongoing. The use of ochre seems also date to the early Middle Pleistocene. However, other types of symbolic resources such as beads, incised patterns on stone and bone, and painted images are restricted to the Late Pleistocene (d'Errico et al., 2001; Henshilwood et

250

al., 2004; Henshilwood et al., 2009; Henshilwood et al., 2001). Part of what makes human behavior so striking compared to extant non-human primates, our enhanced capacities for symbolic representation in material culture, is lacking or is at least of a different character in the early Middle Pleistocene than more recent times.

There are a few potential explanations for this observation. First, the cumulative nature of culture obviously implies that culture needs time to cumulate. The ‘ratchet’ takes time to ‘ratchet up’. The lack of certain behaviors that we identify with modern humans like projectiles, complex technologies, ethnic markers, and externally-stored symbols, in the early MSA does not necessarily reflect the lack of the capacity for these behaviors. Instead, the early MSA may represent a period too soon along the unique historical trajectory that characterizes modern humans for us to expect these kinds of material finds.

Furthermore, preservation and archaeological visibility affect what kind of material culture is recovered from the archaeological record. Chapter 3 highlighted the lack of chronometric knowledge we currently have for the Middle Pleistocene in Africa, in part due to the history of archaeological research, together with the problem of finding good stratigraphic contexts that can be chronometrically dated.

Moreover, it remains possible that some of the technological differences between the early and later MSA could be explained by environmental factors. For example, tool kit ‘complexity’ is known to correlate with latitude (Oswalt, 1976) and toolkit design is affected by environmental predictability, with ‘reliable’ (highly-standardized) toolkits useful in predictable environments, and ‘maintainable’ (flexible, expedient) toolkits useful in unpredictable environments (Bleed, 1986). Much of what we know for the earlier MSA comes from inland sites in the , whereas many later MSA sites with evidence for projectile technology, hafted geometrics, and externally-stored symbolism are neo-coastal cave sites. Because many of these cave sites were washed out by high sea stands in MIS 5, at this point we have no frame of reference for how or if the same kind of coastal landscape was exploited in the early Middle Pleistocene.

Demography is another important factor when considering technological complexity. The persistence of complex technologies requires large populations of people to compensate for loss

251 of skills due to imperfect transmission (Henrich, 2004). Based on genetic studies, Powell et al. (2009) estimate that Africa would not have reached the critical effective population for the appearance of behaviorally modern traits until ~100 ka.

The above points highlight some possible arguments that could underplay the difference between the MSA in the Middle Pleistocene and the MSA in the Late Pleistocene. There remains the possibility that the dissimilarities do demonstrate biologically determined cognitive differences. The lack of symbolic items, with the possible exception of pigments, could imply that early MSA hominins had restricted communicative capacities compared to modern humans. This would be in line with the lack of archaeological evidence for expanded social networks prior to late Middle Pleistocene or Late Pleistocene.

Conformity bias is also an important variable to consider when interpreting the Middle and Late Pleistocene archaeological records. Conformity is represented by a strong bias to adopt the behavioral variants of others because to deviate would expose an individual to third-party punishment (Boyd and Richerson, 2005; Richerson and Boyd, 2005). In mathematical models, the strength of conformity bias has direct consequences for quantitative measures of cultural diversity (Kandler and Laland, 2009). Point styles in the later MSA are consistent with a fairly high conformity bias, if we presume that point form served to communicate group membership and ethnic identity. Individuals conformed to group norms by creating points similar in form to those created by others in their group. However, the earlier MSA, which might lack this kind of temporal and spatial patterning in point form, could reflect a relaxed conformity bias compared to more recent time periods. This would also be consistent with the suggestion above that emulation was more important than imitation with respect to how stone tools were manufactured in the early Middle Pleistocene. Diversity in the chaines operatoires of lithic reduction implies that there was no negative social consequence of emulating, rather than imitating. That is to say, there was no social reason to conform. Relaxed conformity bias would also be consistent with smaller populations and smaller social networks, because conformity is essentially a strategy for establishing, maintaining, and reaffirming bonds among nonkin (Boyd and Richerson, 2005:129).

I suggest, based on evidence for new technological behaviors at KP1 ~500 ka, that the early Middle Pleistocene was witness to a significant increase in what Tennie et al. (2009) describe as

252 the ‘zone of latent solutions’ (ZLS) -- the behaviors that an individual could learn on their own under the appropriate ecological and social conditions. A new way of framing the modern human behavior debate is to ask whether this new ZLS in the early Middle Pleistocene is essentially the same as what we observe for modern humans, or whether there were additional biologically determined developments to human cognition. I would argue it is essentially the same based on technological similarities (i.e., blade production, Levallois technology, diversity in reduction strategies, hafted hunting tools) between earlier and later MSA, though some aspects of cognition that are exclusively linked to social capacities could have had a later appearance. This new capacity for independent invention ~500 ka, together with what might have been a reduced conformity bias, may have served as a pre-adaptation that led to significant changes in later Homo. At the onset of the MSA, innovation rates may have been slower due to smaller population sizes. Lithic-tipped spears may have reduced adult and juvenile mortality with implications for human life history that affect sociality and cooperation with nonkin, leading to increased opportunity for social learning, the extension of social networks, expanding populations, and eventually, to the faster pace of cumulative culture change that characterizes humanity today. Evidence for diverse core reduction strategies and hafted hunting technologies at KP1 suggests that this process began ~500 ka with the onset of the MSA, even if many aspects of modern human sociality, particularly large-scale and long-distance cooperation with nonkin, were lacking.

8.3 Future Directions for Research

The evidence from KP1 raises important questions about the social brain hypothesis. There is fairly robust evidence for a substantial encephalization event ~600 ka with the appearance of H. heidelbergensis s.l. and KP1 pushes back the origins of MSA technology, hafted hunting technology, and human-like technological diversity to roughly this time. These new behaviors could reflect new social structures that affected social learning and cooperation, and as discussed above, they might solve what Barham (2010) described as ‘Dunbar’s dilemma’. However, the technological changes documented for the early Middle Pleistocene seem to mainly reflect adaptations to the ecological, rather than the social realm. A shift in focus from core tools to flake tools could indicate new, perhaps more specialized, functions for stone tools and new strategies for exploiting resources. Points used as spear tips are one example of a new tool type used to adapt to ecological conditions. The kind of technological diversity represented by early

253

Middle Pleistocene stone tool technologies is more consistent with behavioral adaptation to fluctuating local environmental conditions, rather than intentional signaling of style or group membership. The types of raw materials used during the ~500 ka year old occupation of KP1 do not differ significantly from the underlying Acheulean occupation, indicating that there was probably no major shift in the how raw materials were moved across the landscape, and perhaps, no major shift in mobility, no increase in population or group size, and no expansion of the social network. The aspects of the archaeological record that we typically associate with increased interaction and expanded social networks, such as the use of symbolic resources, long-distance exchange of lithic raw material, and spatially and temporally-restricted formal style consistent with ethnic signaling, appear much later in time.

Rather than solving ‘Dunbar’s Dilemma’, KP1 rearticulates it. If humans have large brains mainly to solve social problems, then why do most of the significant behavioral changes correlated with an increase in brain size in the early Middle Pleistocene appear to solve ecological rather than social problems? Why are so many social adaptations absent until the Late Pleistocene?

Preservation issues, as well as the fact that group size and population have not been systematically considered in the early Middle Pleistocene of Africa using the archaeological record, could explain this puzzle. However, the archaeological record can and should contribute to testing the social brain hypothesis. I suggest that future research into the Middle Pleistocene archaeological record should start to investigate the variables of group size, population, and hunter-gatherer land use strategies, to address the question of whether there were significant changes to hominin social life in the early Middle Pleistocene. Did group size increase and was there an expansion in the size and importance of social networks? There are some established quantitative methods for determining group size and home range and these have been discussed above. Grove’s (2010) is one method that seems promising and is applicable to the Middle Pleistocene archaeological record. Extensive survey focused on identifying the location and nature of lithic scatters in the region around KP1 is required to apply this methodology.

A central issue plaguing our ability to interpret the Middle Pleistocene archaeological record is how we interpret lithic variability within and between sites, inter and intraregionally, and beyond. Traditionally, temporal variability in the Middle Pleistocene is interpreted as reflecting

254 cognitive differences, and spatial differences between the roughly contemporary African MSA and European Middle Paleolithic records are generally interpreted as reflecting cognitive differences between Homo sapiens and Neanderthals. Cognition is one possible explanation for variability in lithic technology across time and space, but it is not the only one and should probably not be the default explanation just because we often lack the contextual information required to test other explanations for lithic variability. The environmental context for the Middle Pleistocene remains largely unknown for most of Africa, and in many cases we do not yet have the scale of resolution required to test hypotheses about the relationship between local ecological conditions and technological strategies. Yet, based on what we know about modern hunter- gatherers, environmental contexts probably do account for much of the variability we see in the African MSA, and between the African and European records.

The relationship between lithic technology and the concepts of innovation, and cumulative cultural transmission as envisioned by Boyd, Richerson, Tomasello, and others (Boyd and Richerson, 2005; Richerson and Boyd, 2005; Tennie et al., 2009; Tomasello, 1999; Tomasello et al., 1993) is under-theorized, despite the prominent role that lithic technology plays in building models for the origins and evolution of human culture. We know that behavioral diversity is impacted by rates of innovation and the strength of various social transmission biases (Kandler and Laland, 2009), but there have been no attempts to quantify diversity in the archaeological record of the African MSA. Methods for quantifying diversity in the archaeological record exist (e.g. Conkey et al., 1980; Grayson and Cole, 1998; Shott, 1986) and an interesting investigation would be to compare quantified measures for lithic technological diversity throughout the African Stone Age. We need a framework for how lithic technological diversity changes through time and how it differs across the African continent and in relation to local ecological conditions, in order test hypotheses about innovation and cumulative cultural transmission through the Middle and Late Pleistocene.

8.4 Conclusion

My aim was to describe the technological behaviors represented by the stratum 4a lithic assemblage from KP1 and to situate new evidence from this site into evolutionary context. I argue that KP1 pushes back the origins of MSA technology, and in doing so, poses problems for models that link the origins of Mode 3 technology to H. helmei and dispersals out of Africa

255

(Foley and Lahr, 2003; Foley and Lahr, 1997). Instead, the new timing for the origins of Mode 3 technology is more in line with the appearance of H. heidelbergensis s.l., and provides a parsimonious explanation for many of the technological similarities between Neanderthals and modern humans, who shared a last common ancestor between ~800 and 400 ka based on genetic evidence (Noonan et al. 2006, Green et al. 2006, Green et al. 2010, Wall and Kim 2007, Endicott et al. 2010). The diversity of core reduction strategies represented at KP1, and between early Middle Pleistocene sites across Africa and the Levant is consistent with ‘human-like’ degrees of ‘behavioral variability’ (Shea, 2011b), as well as enhancements in cognitive capacities for working memory (Wynn and Coolidge, 2011) and constructive memory (Ambrose, 2010). The new technological behaviors represented by KP1 seem to be coeval with global-scale increases in climatic fluctuation (Potts, 1998), and increases in brain size (Rightmire, 2001, 2004). In some ways KP1 solves a previously identified problem, that there were no significant behavioral changes visible in the archaeological record when brain size in the genus Homo was increasing the most (Barham, 2010). KP1 indicates that there are significant changes associated with the early Middle Pleistocene ~500 ka. These changes have not been previously recognized because of limits imposed on us by the nature of the archaeological record and chronometric analyses. KP1 also creates a challenge for the social brain hypothesis (Dunbar, 1998; Dunbar, 2003), because many of the new technologies seem to represent adaptations to ecological pressures, rather than social pressures. Obvious evidence for adaptation to social pressures appear later in the MSA, perhaps not until the Late Pleistocene. Further research is required to determine the significance of this observation. At least in part this research must focus on developing a robust framework for interpreting technological variability in the MSA, with emphasis on how ecological conditions affect lithic assemblage variability and how quantifiable patterns of diversity between and within lithic assemblages relates to social learning, innovation, and cumulative cultural change. An appropriate framework would permit confident explanations for the differences between the archaeological records of the early Middle Pleistocene and Late Pleistocene, ultimately establishing when and why humans started to do the things that humans do, and in so many different ways.

256

References:

Aiello, L.C., Dunbar, R.I.M., 1993. Neocortex Size, Group Size, and the Evolution of Language. Current Anthropology 34, 184-193.

Ambrose, S.H., 2001. Paleolithic Technology and Human Evolution. Science 291, 1748-1753.

Ambrose, Stanley H., 2010. Coevolution of Composite-Tool Technology, Constructive Memory, and Language: Implications for the Evolution of Modern Human Behavior. Current Anthropology 51, S135-S147.

Ambrose, S.H., Lorenz, K.G., 1990. Social and ecological models for the Middle Stone Age in southern Africa, (ed. Mellars, Paul) The Emergence of Modern Humans. Edinburgh University Press, Edinburgh.

Andrefsky, W., 1994. Raw-material availability and the organization of technology. American Antiquity 59, 21-35.

Asfaw, B., Beyene, Y., Suwa, G., Walter, R.C., White, T.D., WoldeGabriel, G., Yemane, T., 1992. The earliest Acheulean from Konso-Gardula. Nature 360, 732-735.

Avery, D.M., 2007. Pleistocene micromammals from Wonderwerk Cave, South Africa: practical issues. Journal of Archaeological Science 34, 613-625.

Backwell, L., d'Errico, F., Wadley, L., 2008. Middle Stone Age bone tools from the Howiesons Poort layers, Sibudu Cave, South Africa. Journal of Archaeological Science 35, 1566-1580.

Bar-Yosef, O., 2002. The Upper Paleolithic Revolution. Annual Review of Anthropology 31, 363-394.

Bar-Yosef, O., Kuhn, S.L., 1999. The Big Deal about Blades: Laminar Technologies and Human Evolution. American Anthropologist 101, 322-338.

Barham, L.S., 1998. Possible Early Pigment Use in South-Central Africa. Current Anthropology 39, 703-710.

Barham, L.S., 2000. The Middle Stone Age of Zambia, south central Africa. Western Academic and Specialist Press Limited, Bristol.

Barham, L.S., 2002. Systematic Pigment Use in the Middle Pleistocene of South-Central Africa. Current Anthropology 43, 181-190.

Barham, L.S., 2010. A Technological Fix of 'Dunbar's Dilemma'? Proceedings of the British Academy 158, 367-389.

Barham, L.S., Smart, P.L., 1996. Current events: An early date for the Middle Stone Age of central Zambia. Journal of Human Evolution 30, 287-290.

257

Barnard, W.D., Smit, P., Van Zyl, J.A., 1972. Suid-Afrika: Die Land en sy Streke. Nasionale Opvoedkundige Uitgewers, Cape Town.

Barton, R.N.E., Bergman, C.A., 1982. Hunters at Hengistbury: Some Evidence from . World Archaeology 14, 237-248.

Beaumont, P.B., 1990a. Canteen Koppie (Klipdrift), in: Beaumont, P.B., Morris, D. (Eds.), Guide to Archaeological Sites in the Northern Cape. McGregor Museum, Kimberley, pp. 14-17.

Beaumont, P.B., 1990b. Kathu Pan, in: Beaumont, P.B., Morris, D. (Eds.), Guide to Archaeological Sites in the Northern Cape. McGregor Museum, Kimberley, pp. 75-100.

Beaumont, P.B., 1990c. Nooitgedacht 2 and Roseberry Plain 1, in: Beaumont, P.B., Morris, D. (Eds.), Guide to Archaeological Sites in the Northern Cape. McGregor Museum, Kimberley.

Beaumont, P.B., 1990d. Pniel 1, in: Beaumont, P.B., Morris, D. (Eds.), Guide to Archaeological Sites in the Northern Cape. McGregor Museum, Kimberley, pp. 10-12.

Beaumont, P.B., 1990e. Pniel 6, in: Beaumont, P.B., Morris, D. (Eds.), Guide to Archaeological Sites in the Northern Cape. McGregor Museum, Kimberley, pp. 7-9.

Beaumont, P.B., 1990f. Rooidam 1 and 2 and Biesiesput, in: Beaumont, P.B., Morris, D. (Eds.), Guide to Archaeological Sites in the Northern Cape. McGregor Museum, Kimberley, pp. 19-22.

Beaumont, P.B., 1990g. Wonderwerk Cave, in: Beaumont, P., Morris, D. (Eds.), Guide to Archaeological Sites in the Northern Cape. McGregor Museum, Kimberley, pp. 101-134.

Beaumont, P.B., 2004a. Kathu Pan and Kathu Townlands/Uitkoms, in: Morris, D., Beaumont, P.B. (Eds.), Archaeology in the Northern Cape: Some Key Sites. McGregor Museum, Kimberley, pp. 50-52.

Beaumont, P.B., 2004b. Wonderwerk Cave, in: Morris, D., Beaumont, P.B. (Eds.), Archaeology in the Northern Cape: Some Key Sites. McGregor Museum, Kimberley, pp. 31-36.

Beaumont, P.B., McNabb, J., 2000. Canteen Koppie - The Recent Excavations. The Digging Stick 17, 3-7.

Beaumont, P.B., Morris, D., 1990. Guide to Archaeological Sites in the Northern Cape. McGregor Museum, Kimberley.

Beaumont, P.B., Vogel, J.C., 2006. On a timescale for the past million years of in central South Africa. South African Journal of Science 102, 6-6.

Bergman, C.A., Newcomer, H., 1983. Flint arrowhead breakage: examples from Kasr Akil, Lebanon. Journal of Field Archaeology 10, 238-243.

Berna, F., Goldberg, P., Horwitz, L.K., Brink, J., Holt, S., Bamford, M., Chazan, M., 2012. Microstratigraphic evidence of in situ fire in the Acheulean strata of Wonderwerk Cave, Northern Cape province, South Africa. Proceedings of the National Academy of Sciences.

258

Beyries, S., Plisson, H., 1998. Pointes ou outils triangulaires ? Données fonctionnelles dans le Moustérien levantin. Paleorient, 5-24.

Binford, L.R., 1979. Organization and formation processes: looking at curated technologies. Journal of Anthropological Research 35, 255-273.

Binford, L.R., 1980. Willow Smoke and Dogs' Tails: Hunter-Gatherer Settlement Systems and Archaeological Site Formation. American Antiquity 45, 4-20.

Binford, L.R., 1981. Bones: Ancient men and modern myths. Academic Press, New York.

Binford, L.R., 1985. Human ancestors: Changing views of their behavior. Journal Of Anthropological Archaeology 4, 292-327.

Binford, L.R., 2001. Constructing Frames of Reference: An Analytical Method for Archaeological Theory Building Using Ethnographic and Environmental Data Sets. University of Press, Berkeley.

Bird, C., Minichillo, T., Marean, C.W., 2007. Edge damage distribution at the assemblage level on Middle Stone Age lithics: an image-based GIS approach. Journal of Archaeological Science 34, 771-780.

Bleed, P., 1986. The optimal design of hunting weapons: maintainability or reliability. American Antiquity 51, 737-747.

Boëda, E., 1986. Approche technologique du conept Levallois et evaluation de son Champ d'Application: Etude de trois gisements Salliens et Weichseliens de la France Septentionale. Universite de Paris IX.

Boëda, E., 1991. Approches de la variabilite des sytems de production lithique des industries de paleolithique inferieure et moyen: chronique d'une varibilite attendue. Techniques culture 17-18, 31-79.

Boëda, E., 1995. Levallois: A Volumetric Construction, Methods, a Technique, The Definition and Interpretation of Levallois Technology. Prehistory Press, Madison, pp. 41-68.

Boëda, E., Geneste, J.M., Meignen, L., 1990. Identification des chaines operatoires lithiques du Paleolithique ancien et moyen. Paleo 2.

Boëda, E., Pelegrin, J., 1980. Approches technologiqe du nucleus Levallois a eclat. Etudes Prehist 15, 41-48.

Bond, G., 1948. Rhodesian Stone Age Man and His Raw Materials. The South African Archaeological Bulletin 3, 55-60.

Boyd, R., Richerson, P.J., 1985. Culture and the Evolutionary Process. University of Chicago Press, Chicago.

259

Boyd, R., Richerson, P.J., 1987. The Evolution of Ethnic Markers. Cultural Anthropology 2, 65- 79.

Boyd, R., Richerson, P.J., 2005. The Origin and Evolution of Cultures. Oxford University Press, New York.

Brauer, G., Yokoyama, Y., Falgueres, C., Mbua, E., 1997. Modern human origins backdated. Nature 386, 337-338.

Braun, D.R., Harris, J.W.K., Levin, N.E., McCoy, J.T., Herries, A.I.R., Bamford, M.K., Bishop, L.C., Richmond, B.G., Kibunjia, M., 2010. Early hominin diet included diverse terrestrial and aquatic animals 1.95 Ma in East Turkana, Kenya. Proceedings of the National Academy of Sciences 107, 10002-10007.

Braun, D.R., Plummer, T., Ditchfield, P., Ferraro, J.V., Maina, D., Bishop, L.C., Potts, R., 2008. Oldowan behavior and raw material transport: perspectives from the Kanjera Formation. Journal Of Archaeological Science 35, 2329-2345.

Braun, D.R., Plummer, T., Ferraro, J.V., Ditchfield, P., Bishop, L.C., 2009. Raw material quality and Oldowan hominin toolstone preferences: evidence from Kanjera South, Kenya. Journal Of Archaeological Science 36, 1605-1614.

Brink, J.S., Herries, A.I.R., Moggi-Cecchi, J., Gowlett, J.A.J., Bousman, C.B., Hancox, J.P., Grün, R., Eisenmann, V., Adams, J.W., Rossouw, L., 2012. First hominine remains from a ∼1.0 million year old bone bed at Cornelia-Uitzoek, Free State Province, South Africa. Journal of Human Evolution 63, 527-535.

Brooks, A., Behrensmeyer, A.K., Yellen, J., Deino, A., Sharp, W., Potts, R., 2007. After the Acheulean: Stratigraphy, Dating, and Archeology of Two New Formations in the Olorgesailie Basin, Southern Kenya Rift. Abstracts of the PaleoAnthropology Society 2007 Meetings, PaleoAnthropology, A5.

Brooks, A., Nevell, L., Yellen, J., Hartman, G., 2006. Projectile Technologies of the African MSA: Implications for Modern Human origins, in: Hovers, E., Kuhn, S.L. (Eds.), Transitions before the Transition: Evolution and Stability in the Middle Paleolithic and Middle Stone Age. Springer, New York, pp. 233-255.

Brown, K.S., 2011. The Sword in the Stone: lithic raw material exploitation in the Middle Stone Age at Pinnacle Point Site 5-6, Southern Cape, South Africa, Unpublished Dissertation. , Cape Town.

Brown, K.S., Marean, C.W., 2010. Wood Fuel Availability for Heat Treatment Drives the Rise and Fall of Silcrete as a Raw Material in the Middle Stone Age of South Africa. Abstracts of the PaleoAnthropology Society 2010 Meetings, PaleoAnthropology 2010, A6-A6.

Brown, K.S., Marean, C.W., Herries, A.I.R., Jacobs, Z., Tribolo, C., Braun, D., Roberts, D.L., Meyer, M.C., Bernatchez, J., 2009. Fire As an Engineering Tool of Early Modern Humans. Science 325, 859-862.

260

Bruggemann, J.H., Buffler, R.T., Guillaume, M.M.M., Walter, R.C., von Cosel, R., Ghebretensae, B.N., Berhe, S.M., 2004. Stratigraphy, palaeoenvironments and model for the deposition of the Abdur Reef Limestone: context for an important archaeological site from the last interglacial on the coast of Eritrea. Palaeogeography, Palaeoclimatology, Palaeoecology 203, 179-206.

Butzer, K.W., 1974. Geo-archeological interpretation of Acheulian calc-pan sites at Doornlaagte and Rooidam (Kimberley, South Africa). Journal of Archaeological Science 1, 1-25.

Butzer, K.W., 1984. Archeology and Quarternary environment in the interior of southern Africa, in: Klein, R.G. (Ed.), Southern African Prehistory and Paleoenvironments. A. A. Balkema, Rotterdam, pp. 1-64.

Chazan, M., 1995. The Language Hypothesis for the Middle-to-Upper Paleolithic Transistion. An Examination based on a Multiregional Lithic Analysis. Current Anthropology 36, 749-768.

Chazan, M., 1997. Redefining Levallois. Journal of Human Evolution 33, 719-735.

Chazan, M., in press. The Earlier Stone Age Lithic Assemblage from Wonderwerk Cave, Northern Cape Province, South Africa, Proceedings of the Society of Africanist Archaeologists, Dakar, Sengal.

Chazan, M., Avery, D.M., Bamford, M.K., Berna, F., Brink, J., Fernandez-Jalvo, Y., Goldberg, P., Holt, S., Matmon, A., Porat, N., Ron, H., Rossouw, L., Scott, L., Horwitz, L.K., 2012a. The Oldowan horizon in Wonderwerk Cave (South Africa): Archaeological, geological, paleontological and paleoclimatic evidence. Journal of Human Evolution 63, 859-866.

Chazan, M., Avery, M., Goldberg, P., Matmon, A., Porat, N., Ron, H., Ruther, H., Sumner, A., Yates, R., Horwitz, L.K., in press. The Earlier Stone Age Sequence in the Northern Cape Province, South Africa: New Research at Wonderwerk Cave, in: De Lumley, H. (Ed.), Les Culture A Biface du Pleistocene Inferieur et Moyen dans le Monde, Tautavel.

Chazan, M., Horwitz, L.K., 2009. Milestones in the development of symbolic behaviour: a case study from Wonderwerk Cave, South Africa. World Archaeology 41, 521-539.

Chazan, M., Horwitz, L.K., in press. An Overview of Recent Research at Wonderwerk Cave, South Africa, Proceedings of the Society of Africanist Archaeologists, Dakar, .

Chazan, M., Ron, H., Matmon, A., Porat, N., Goldberg, P., Yates, R., Avery, M., Sumner, A., Horwitz, L.K., 2008. Radiometric dating of the Earlier Stone Age sequence in Excavation I at Wonderwerk Cave, South Africa: preliminary results. Journal of Human Evolution 55, 1-11.

Chazan, M., Wilkins, J., Morris, D., Berna, F., 2012b. Bestwood 1: a newly discovered Earlier Stone Age living surface near Kathu, Northern Cape Province, South Africa. Antiquity Bulletin 86.

Churchill, S.E., 1993. Weapon technology, prey size selection, and hunting methods in modern hunter-gatherers: implications for hunting in the Palaeolithic and . Archeological Papers of the American Anthropological Association 4, 11-24.

261

Clark, D.E., Purdy, B.A., 1979. Electron microprobe analysis of weathered Florida chert. American Antiquity 44, 517-524.

Clark, G., 1969. World prehistory: A new synthesis. Cambridge University Press, Cambridge.

Clark, G., 1977. World Prehistory in New Perspective. Cambridge University Press, Cambridge.

Clark, G.A., Riel-Salvatore, J., 2005. Observations on Systematics in Paleolithic Archaeology, in: Hovers, E., Kuhn, S.L. (Eds.), Transitions before the Transition. Springer, New York, pp. 29- 56.

Clark, J., de Heinzelin, J., Schick, K., Hart, W., White, T., WoldeGabriel, G., Walter, R., Suwa, G., Asfaw, B., Vrba, E., et, a., 1994. African Homo erectus: old radiometric ages and young Oldowan assemblages in the Middle Awash Valley, Ethiopia. Science 264, 1907-1910.

Clark, J.D., 1954. An Early Upper Pleistocene Site at the Kalambo Falls on the Northern Rhodesia/Tanganyika Border. The South African Archaeological Bulletin 9, 51-56.

Clark, J.D., 1959a. Further Excavations at Broken Hill, Northern Rhodesia. The Journal of the Royal Anthropological Institute of Great Britain and Ireland 89, 201-232.

Clark, J.D., 1959b. The Prehistory of Southern Africa. Penguin Books, London.

Clark, J.D., 1967. Atlas of African Prehistory. Univeristy of Chicago Press, Chicago.

Clark, J.D., 1970. The Prehistory of Africa. Thames and Hudson, Praegers, London.

Clark, J.D., 1980. Raw material and African lithic technology. Man and Environment 4, 44-55.

Clark, J.D., 1982. The Cultures of the Middle Paleolithic/Middle Stone Age, Cambridge , Vol. I: From the Earliest Times to ca. 500 B.C. Cambridge University Press, London, pp. 284-341.

Clark, J.D., 1992. African and Asian Perspectives on the origins of modern humans. Philosophical Transactions: Biological Sciences 337, 201-215.

Clark, J.D., 2001. Kalambo Falls Prehistoric Site III. The earlier cultures: Middle and Earlier Stone Age. Cambridge University Press, Cambridge.

Clark, J.D., Beyene, Y., WoldeGabriel, G., Hart, W.K., Renne, P.R., Gilbert, H., Defleur, A., Suwa, G., Katoh, S., Ludwig, K.R., Boisserie, J.-R., Asfaw, B., White, T.D., 2003. Stratigraphic, chronological and behavioural contexts of Pleistocene Homo sapiens from Middle Awash, Ethiopia. Nature 423, 747-752.

Clark, J.D., Kurashina, H., 1979. An Analysis of Earlier Stone Age Bifaces from Gadeb (Locality 8E), Northern Bale Highlands, Ethiopia. The South African Archaeological Bulletin 34, 93-109.

Clarke, R.J., 1990. The Ndutu cranium and the origin of Homo sapiens. Journal of Human Evolution 19, 699-736.

262

Conard, N., 1989. Laminar lithic assemblages from the last interglacial complex in northwestern Europe. Journal of Anthropological Research 46 243-262.

Conkey, M.W., Beltrán, A., Clark, G.A., Echegaray, J.G.l., Guenther, M.G., Hahn, J., Hayden, B., Paddayya, K., Straus, L.G., Valoch, K., 1980. The Identification of Prehistoric Hunter- Gatherer Aggregation Sites: The Case of Altamira [and Comments and Reply]. Current Anthropology 21, 609-630.

Conroy, G.C., Weber, G.W., Seidler, H., Recheis, W., Zur Nedden, D., Mariam, J.H., 2000. Endocranial capacity of the Bodo cranium determined from three‐dimensional computed tomography. American Journal of Physical Anthropology 113, 111-118.

Cruz-Uribe, K., Klein, R.G., Avery, G., Avery, M., Halkett, D., Hart, T., Milo, R.G., Garth Sampson, C., Volman, T.P., 2003. Excavation of buried Late Acheulean (Mid-) land surfaces at Duinefontein 2, Western Cape Province, South Africa. Journal of Archaeological Science 30, 559-575. d'Errico, F., Henshilwood, C., Nilssen, P., 2001. An engraved bone fragment from c. 70,000- year-old Middle Stone Age levels at Blombos Cave, South Africa: implications for the origin of symbolism and language. ANTIQUITY 75, 309-318. de Wit, M.C.J., 2008. Canteen Koppie at : South Africa’s first diamond mine. South African Journal of Geology 111, 53-66. de Wit, M.C.J., Marshall, T.R., Partridge, T.C., 2000. Fluvial deposits and drainage evolution, The of Southern Africa. Oxford University Press, New York, pp. 55-72.

Deacon, H.J., 1989. Late Pleistocene paleoecology and archaeology in the southern Cape, (Ed. P. Mellars and C. Stringer) The Human Revolution: Behavioral and Biological Perspectives on the origins of Modern Humans. Press, Princton, NJ, pp. 547-564.

Deacon, H.J., Deacon, J., 1999. Human Beginnings in South Africa: Uncovering the Secrets of the Stone Age. David Philip Publishing, Claremont.

DeHeinzelin, J., Clark, J.D., Schick, K.D., Gilbert, W.H., 2000. The Acheulean and the Plio- Pleistocene Deposits of the Middle Awash Valley, Ethiopia Royal Museum of Central Africa, Tervuren, Belgium; Ann. Sci. Geol. 104.

Deino, A., Potts, R., 1990. Single-Crystal 40Ar/39Ar Dating of the Olorgesailie Formation, Southern Kenya Rift. J. Geophys. Res. 95, 8453-8470.

Deino, A., Potts, R., 1992. Age-probability spectra for examination of single-crystal 40Ar39Ar dating results: Examples from Olorgesailie, southern Kenya Rift. Quaternary International 13– 14, 47-53.

Deino, A.L., McBrearty, S., 2002. 40Ar/39Ar dating of the Kapthurin Formation, Baringo, Kenya. Journal of Human Evolution 42, 185-210.

263

Delagnes, A., Meignen, L., 2006. Diversity of Lithic Production Systems During the Middle Paleolithic in France, in: Hovers, E., Kuhn, S.L. (Eds.), Transitions Before the Transition: Evolution and Stability in the Middle Paleolithic and Middle Stone Age.

Dobres, M.-A., 2000. Technology and Social Agency: Outlining a practice framework for archaeology. Blackwell Publishing, Oxford.

Dominguez- Rodrigo, M., Serrallonga, J., Juan-Tresserras, J., Alcala, L., Luque, L., 2001. Woodworking activities by early humans: a plant residue analysis on Acheulian stone tools from Peninj (Tanzania). Journal of Human Evolution 40, 289-299.

Domínguez-Rodrigo, M., Pickering, T.R., Semaw, S., Rogers, M.J., 2005. Cutmarked bones from Pliocene archaeological sites at Gona, Afar, Ethiopia: implications for the function of the world's oldest stone tools. Journal of Human Evolution 48, 109-121.

Donahue, R.E., Murphy, M.L., Robbins, L.H., 2004. Lithic Microwear Analysis of Middle Stone Age Artifacts from White Paintings Rock Shelter, Botswana. Journal of Field Archaeology 29, 155-163.

Dreyer, T.F., 1935. A human from Florisbad. Proceedings of the Academy of Science Amsterdam 38, 119-128.

Dunbar, R.I.M., 1998. The Social Brain Hypothesis. Evolutionary Anthropology: Issues, News, and Reviews 6, 178-190.

Dunbar, R.I.M., 2003. The Social Brain: Mind, Language, and Society in Evolutionary Perspective. Annual Review of Anthropology 32, 163-181.

Eren, M.I., Greenspan, A., Sampson, C.G., 2008. Are Upper Paleolithic blade cores more productive than Middle Paleolithic discoidal cores? A replication experiment. Journal of Human Evolution 55, 952-961.

Eriksson, P.G., Altermann, W., Hartzer, F.J., 2006. The The Transvaal Supergroup and its Precursors, in: Johnson, H.R., Anhaeusser, C.R., Thomas, R.J. (Eds.), The . /Pretoria, Geological Society of South Africa and the Council for Geoscience, pp. 237-260.

Evernden, J.F., Curtis, G.H., 1965. The Potassium-Argon Dating of Late Cenozoic Rocks in East Africa and Italy Current Anthropology 6, 342-385.

Faith, T., 2008. Eland, buffalo, and wild : were Middle Stone Age humans effective hunters? Journal of Human Evolution 55, 24-36.

Feathers, J.K., 2002. Luminescence Dating in Less Than Ideal Conditions: Case Studies from Klasies River Main Site and Duinefontein, South Africa. Journal of Archaeological Science 29, 177-194.

264

Feblot-Augustins, J., 1993. Mobility Strategies in the Late Middle Palaeolithic of and Western Europe: Elements of Stability and Variability. Journal Of Anthropological Archaeology 12, 211-265.

Fischer, A., Vemming Hansen, P., Rasmussen, P., 1984. Macro and micro wear traces on lithic projectile points: experimental results and prehistoric examples. Journal of Danish Archaeology 3, 19-46.

Fisher, E., Bar-Matthews, M., Jerardino, A., Marean, C., 2010. Middle and Late Pleistocene paleoscape modeling along the southern coast of South Africa. Reviews 29, 1382-1398.

Fock, G.J., 1968. Rooidam: A sealed site of The First Intermediate. South African Journal of Science 64, 153-158.

Foley, R., Gamble, C., 2009. The ecology of social transitions in human evolution. Philosophical Transactions of the Royal Society B: Biological Sciences 364, 3267-3279.

Foley, R., Lahr, M.M., 2003. On stony ground: Lithic technology, human evolution, and the emergence of culture. Evolutionary Anthropology: Issues, News, and Reviews 12, 109-122.

Foley, R.A., Lahr, M.M., 1997. Mode 3 technologies and the evolution of modern humans. Cambridge Archaeological Journal 7, 3-36.

Gamble, C., 1999. The Palaeolithic Societies of Europe. Cambridge University Press, Cambridge.

Gamble, C., 2010. Technologies of Separation and the Evolution of Social Extension. Proceedings of the British Academy 158, 17-42.

Gamble, C., Gowlett, J., Dunbar, R., 2011. The Social Brain and the Shape of the Palaeolithic. Cambridge Archaeological Journal 21, 115-136.

Geraads, D., Hublin, J.-J., Jaeger, J.-J., Tong, H., Sen, S., Toubeau, P., 1986. The Pleistocene hominid site of Ternifine, Algeria: New results on the environment, age, and human industries. Quaternary Research 25, 380-386.

Gibbon, R.J., Granger, D.E., Kuman, K., Partridge, T.C., 2009. Early Acheulean technology in the Rietputs Formation, South Africa, dated with cosmogenic nuclides. Journal of Human Evolution 56, 152-160.

Goldman-Neuman, T., Hovers, E., 2012. Raw material selectivity in Late Pliocene Oldowan sites in the Makaamitalu Basin, Hadar, Ethiopia. Journal of Human Evolution 62, 353-366.

Goodwin, A.J.H., 1929. The Montagu Cave: A Full Report of the Investigation of the Montagu Rock-shelter. Annals of the South African Museum XXIV.

Goodwin, A.J.H., Van Riet Lowe, C., 1929. The Stone Age Cultures of South Africa. Annals of the South African Museum 27.

265

Gopher, A., Ayalon, A., Bar-Matthews, M., Barkai, R., Frumkin, A., Karkanas, P., Shahack- Gross, R., 2010. The chronology of the late Lower Paleolithic in the Levant based on U-Th ages of speleothems from Qesem Cave, Israel. Quaternary Geochronology 5, 644-656.

Grace, R., 1989. Interpreting the Function of Stone Tools: The Quantification and Computerization of Microwear Analysis. BAR, Oxford.

Grayson, D.K., Cole, S.C., 1998. Stone Tool Assemblage Richness during the Middle and Early Upper Palaeolithic in France. Journal of Archaeological Science 25, 927-938.

Greenfield, P.M., 1991. Language, tools and brain: The ontogeny and phylogeny of hierarchically organized sequential behavior. Behavioral and Brain Sciences 14, 531-551.

Grine, F.E., Jungers, W.L., Tobias, P.V., Pearson, O.M., 1995. Fossil Homo femur from Berg Aukas, northern Namibia. American Journal of Physical Anthropology 97, 151-185.

Grove, M., 2010. The Archaeology of Group Size. Proceedings of the British Academy 158, 391-411.

Grün, R., Beaumont, P., 2001. Border Cave revisited: a revised ESR chronology. Journal of Human Evolution 40, 467-482.

Grün, R., Brink, J.S., Spooner, N.S., Taylor, L., Stringer, C., Franciscus, R.G., Murray, A., 1996. Direct Dating of Florisbad Hominid. Nature.

Hammer, Ø., Harper, D.A.T., Ryan, P.D., 2001. PAST: Paleontological statistics software package for education and data analysis. Palaeontologia Electronica 4, 9 pp.

Hayden, B., 1979. The Ho Ho Classification and Nomenclature Committee Report, in: Hayden, B. (Ed.), LIthic Use-Wear Analysis. Academic Press, New York, pp. 133-135.

Helgren, D.M., 1978. Acheulian settlement along the Lower Vaal river, South Africa. Journal of Archaeological Science 5, 39-60.

Henrich, J., 2004. Demography and Cultural Evolution: How Adaptive Cultural Processes can Produce Maladaptive Losses: The Tasmanian Case. American Antiquity 69, 197-214.

Henshilwood, C., d'Errico, F., Vanhaeren, M., Van Niekerk, K., Jacobs, Z., 2004. Middle Stone Age Shell Beads from South Africa. Science 304, 404.

Henshilwood, C.S., d'Errico, F., Watts, I., 2009. Engraved ochres from the Middle Stone Age levels at Blombos Cave, South Africa. Journal of Human Evolution 57, 27-47.

Henshilwood, C.S., Marean, C.W., 2003. The Origin of Modern Human Behavior: Critique of the Models and Their Test Implications. Current Anthropology 44, 627-651.

Henshilwood, C.S., Sealy, J.C., Yates, R., Cruz-Uribe, K., Goldberg, P., Grine, F.E., Klein, R.G., Poggenpoel, C., van Niekerk, K., Watts, I., 2001. Blombos Cave, Southern Cape, South Africa:

266

Preliminary Report on the 1992-1999 Excavations of the Middle Stone Age Levels. Journal of Archaeological Science 28, 421-448.

Herries, A.I.R., 2011. A Chronological Perspective on the Acheulian and Its Transition to the Middle Stone Age in Southern Africa: The Question of the Fauresmith. International Journal of Evolutionary Biology 2011, 1-25.

Herries, A.I.R., Curnoe, D., Adams, J.W., 2009. A multi-disciplinary seriation of early Homo and bearing palaeocaves in southern Africa. Quaternary International 202, 14-28.

Herries, A.I.R., Latham, A.G., 2009. Archaeomagnetic studies at the Cave of Hearths, in: McNabb, J., Sinclair, A.G.M. (Eds.), The Cave of Hearths: Makapan Middle Pleistocene Research Project. Achaeopress, Oxford, pp. 59-64.

Hershkovitz, I., Smith, P., Sarig, R., Quam, R., Rodríguez, L., García, R., Arsuaga, J.L., Barkai, R., Gopher, A., 2011. Middle pleistocene dental remains from Qesem Cave (Israel) American Journal of Physical Anthropology 144, 575-592.

Hill, K., Barton, M., Hurtado, A.M., 2009. The Emergence of Human Uniqueness: Characters Underlying Behavioral Modernity. Evolutionary Anthropology 18.

Hill, K., Hurtado, A.M., 2009. Cooperative breeding in South American hunter–gatherers. Proceedings of the Royal Society B: Biological Sciences 276, 3863-3870.

Hill, K.R., Walker, R.S., Božičević, M., Eder, J., Headland, T., Hewlett, B., Hurtado, A.M., Marlowe, F., Wiessner, P., Wood, B., 2011. Co-Residence Patterns in Hunter-Gatherer Societies Show Unique Human Social Structure. Science 331, 1286-1289.

Hodgson, D., 2009. Evolution of the visual cortex and the emergence of symmetry in the Acheulean techno-complex. Comptes rendus - Palevol 8, 93-97.

Howard, C.D., 2002. The Gloss Patination of Flint Artifacts. The Plains Anthropologist 47, 283- 287.

Howell, F.C., Cole, G.H., Kleindienst, M.R., 1962. Isimila an Acheulian occupation site in the Iringa Highlands, Southern Highlands Province, Tanganyika, in: Mortelmans, G., Nenquin, J. (Eds.), Actes du IVe Congres Panafricaine de Prehistoire et de l'Etude du Quaternaire Musee Royal de l'Afrique Centrals, pp. 43-80.

Howell, F.C., Cole, G.H., Kleindienst, M.R., Szabo, B.J., Oakley, K.P., 1972. Uranium-series Dating of Bone from the Isimila Prehistoric Site, Tanzania. Nature 237, 51-52.

Hublin, J.-J., 1991. L’Emergence des Homo sapiens archaiques: Afrique du Nord-Ouest et Euorpe occidentale. The`se d’Etat, Universite´ de Bordeaux 1.

Hughes, S.S., 1998. Getting to the Point: Evolutionary Change in Prehistoric Weaponry. Journal of Archaeological Method and Theory 5, 345-408.

267

Humphreys, A.J.B., 1970. The Role of Raw Material and the Concept of the Fauresmith. The South African Archaeological Bulletin 25, 139-144.

Humphreys, A.J.B., Thackeray, A.I., 1983. Chapter 2: The Environmental Setting, in: Humphreys, A.J.B., Thackeray, A.I. (Eds.), Ghaap and Gariep: Later Stone Age studies in the Northern Cape. The South African Archaeological Society, Cape Town, pp. 15-32.

Hutchings, W.K., 2011. Measuring Use-Related Fracture Velocity in Lithic Armatures to Identify Spears, Javelins, Darts, and Arrows. Journal of Archaeological Science 38, 1737-1746.

Inizan, M.L., Reduron-Ballinger, M., Roche, H., Tixier, J., 1999. Technology and Terminology of Knapped Stone. Nanterre:CREP.

Iovita, R., 2011. Shape Variation in Aterian Tanged Tools and the Origins of Projectile Technology: A Morphometric Perspective on Stone Tool Function. PLoS ONE 6, e29029.

Isaac, G.L., 1977. Olorgesailie: Archaeological Studies of a Middle Pleistocene Lake Basin in Kenya. The University of Chicago Press, Chicago.

Jacobs, Z., 2010. An OSL chronology for the sedimentary deposits from Pinnacle Point Cave 13B--A punctuated presence. Journal of Human Evolution 59, 289-305.

Jacobs, Z., Duller, G.A.T., Wintle, A.G., Henshilwood, C.S., 2006. Extending the chronology of deposits at Blombos Cave, South Africa, back to 140 ka using optical dating of single and multiple grains of quartz. Journal of Human Evolution 51, 255-273.

Jacobs, Z., Roberts, R.G., 2008. Testing Times: Old and New Chronologies for the Howieson's Poort and Still Bay Industries in Environmental Context. South African Archaeological Society Goodwin Series, 9-34.

Jacobs, Z., Roberts, R.G., Galbraith, R.F., Deacon, H.J., Grun, R., Mackay, A., Mitchell, P., Vogelsang, R., Wadley, L., 2008. Ages for the Middle Stone Age of Southern Africa: Implications for Human Behavior and Dispersal. Science 322, 733-735.

Jacobson, L., de Beer, F.C., Nshimirimana, R., 2011. Tomography imaging of South African archaeological and heritage stone and objects. Nuclear Inst. and Methods in Physics Research, A 651, 240-243.

Jacobson, L., de Beer, F.C., Nshimirimana, R., Horwitz, L.K., Chazan, M., in press. Neutron tomographic assessment of incisions on prehistoric stone slabs: a case study from Wonderwerk Cave, South Africa. Archaeometry.

Johnson, C.R., McBrearty, S., 2010. 500,000 year old blades from the Kapthurin Formation, Kenya. Journal of Human Evolution 58, 193-200.

Jones, P.R., 1979. Effects of Raw Materials on Biface Manufacture. Science 204, 835-836.

Jones, P.R., 1994. Results of experimental work in relation to the stone industries of Olduvai Gorge, Olduvai Gorge. Cambridge University Press, Cambridge, pp. 254-298.

268

Kandler, A., Laland, K.N., 2009. An investigation of the relationship between innovation and cultural diversity. Theoretical Population Biology 76, 59-67.

Kaplan, H., Hill, K., Lancaster, J., Hurtado, A.M., 2000. A Theory of Human Life History Evolution: Diet, Intelligence, and Longevity. Evolutionary Anthropology: Issues, News, and Reviews 9, 156-185.

Keeley, L.H., 1980. Experimental Determination of Stone Tool Uses: A Microwear Analysis. University of Chicago Press, Chicago.

Keller, C.M., 1973. Montagu Cave in prehistory: a descriptive analysis. University of California Press, Berkeley.

Kelly, R.L., 1995. The Foraging Spectrum: Diversity in Hunter-Gatherer Lifeways. Press, Washigton D.C.

Kempson, H., 2007. Late Earlier Stone Age Sites in the Mapungubwe National Park, South Africa: A Technological Study. University of the Witwatersrand, Johannesburg.

Kiberd, P., 2006. Bundu Farm: A Report on Archaeological and Palaeoenvironmental Assemblages from a Pan Site in Bushmanland, Northern Cape, South Africa. The South African Archaeological Bulletin 61, 189-201.

Klein, R.G., 1978. The fauna and overall interpretation of the “Cutting 10” Acheulean site at Elandsfontein (Hopefield), southwestern Cape Province, South Africa. Quaternary Research 10, 69-83.

Klein, R.G., 1988. The archaeological significance of animal bones from Acheulean sites in southern Africa. African Archaeological Review 6, 3-25.

Klein, R.G., 1989. The human career: human biological and cultural origins. University of Chicago Press., Chicago.

Klein, R.G., 2001. Southern Africa and Modern Human Origins. Journal of Anthropological Research 57, 1-16.

Klein, R.G., Avery, G., Cruz-Uribe, K., Halkett, D., Hart, T., Milo, R.G., Volman, T.P., 1999. Duinefontein 2: an Acheulean Site in the Western Cape Province of South Africa. Journal of Human Evolution 37, 153-190.

Koechlin, E., Basso, G., Pietrini, P., Panzer, S., Grafman, J., 1999. The role of the anterior prefrontal cortex in human cognition. Nature 399, 148-151.

Kuman, K., in press. Book Review of 'A report on the archaeological assemblages from excavations by Peter Beaumont at Canteen Koppie, Northern Cape, South Africa, by John McNabb and Peter Beaumont. Oxford: Archaeopress, British Archaeological Reports International Series, No. 2275, 2011'. Azania: Archaeological Research in Africa.

269

Kuman, K., Clarke, R.J., 2000. Stratigraphy, artefact industries and hominid associations for , Member 5. Journal of Human Evolution 38, 827-847.

Kuman, K., Gibbon, R., Kempson, H., Langejans, G., Baron, J.L., Pollarolo, L., Sutton, M., 2005a. Stone Age Signatures in Northernmost South Africa:early archaeology in the Mapungubwe National Park and vicinity, in: d'Errico, F., Backwell, L. (Eds.), From Tools to Symbols: From Early Hominids to Modern Humans. Witwatersrand University Press, Johannesburg pp. 163-183.

Kuman, K., Inbar, M., Clarke, R.J., 1999. Palaeoenvironments and Cultural Sequence of the Florisbad Middle Stone Age Hominid Site, South Africa. Journal of Archaeological Science 26, 1409-1425.

Kuman, K., Le Baron, J.C., Gibbon, R.J., 2005b. Earlier Stone Age Archaeology and the Vhembe-Dongola National Park (South Africa) and Vicinity. Quaternary International 120, 23- 32.

Kuman, K.A., 1989. Florisbad and ≠Gi: The Contribution of Open-air Sites to Study of the Middle Stone Age in Southern Africa, Department of Anthropology. University of Pennsylvania, Philadelphia.

Lepre, C.J., Roche, H., Kent, D.V., Harmand, S., Quinn, R.L., Brugal, J.-P., Texier, P.-J., Lenoble, A., Feibel, C.S., 2011. An earlier origin for the Acheulian. Nature 477, 82-85.

Levi Sala, I., 1986. Use wear and post-depositional surface modification: A word of caution. Journal of Archaeological Science 13, 229-244.

Lombard, M., 2004a. Distribution Patterns of Organic Residues on Middle Stone Age Points from Sibudu Cave, Kwazulu-Natal, South Africa. The South African Archaeological Bulletin 59, 37-44.

Lombard, M., 2004b. Distribution Patterns of Organic Residues on Middle Stone Age Points from Sibudu Cave, Kwazulu-Natal, South Africa. S. Afr. Archaeol. Bull. 59, 37-44.

Lombard, M., 2005a. Evidence of hunting and hafting during the Middle Stone Age at Sibudu Cave, KwaZulu-Natal, South Africa: a multianalytical approach. Journal of Human Evolution 48, 279-300.

Lombard, M., 2005b. The Howiesons Poort of South Africa: what we know, what we think we know, what we need to know. Southern African Humanities 17, 35-55.

Lombard, M., 2005c. A Method for Identifying Stone Age Hunting Tools. South African Archaeological Bulletin 60, 115.

Lombard, M., 2006. Direct Evidence for the use of ochre in the hafting technology of Middle Stone age tools from Sibudu Cave. Southern African Humanities 18, 57-67.

Lombard, M., 2007. Evidence for change in Middle Stone Age hunting behaviors at Blombos Cave: results of a macrofracture analysis. South African Archaeological Bulletin 62, 62-67.

270

Lombard, M., 2008. Finding resolution for the Howiesons Poort through the microscope: micro- residue analysis of segments from Sibudu Cave, South Africa. Journal of Archaeological Science 35, 26-41.

Lombard, M., 2011. Quartz-tipped arrows older than 60 ka: further use-trace evidence from Sibudu, KwaZulu-Natal, South Africa. Journal of Archaeological Science 38, 1918-1930.

Lombard, M., Pargeter, J., 2008. Hunting with Howiesons Poort segments: pilot experimental study and the functional interpretation of archaeological tools. Journal of Archaeological Science 35, 2523-2531.

Lombard, M., Parsons, I., van der Ryst, M.M., 2004. Middle Stone Age lithic point experimentation for macro-fracture and residue analyses: the process and preliminary results with reference to Sibudu Cave points. South African Journal of Science 100, 159-166.

Lombard, M., Phillipson, L., 2010. Indications of bow and stone-tipped arrow use 64 000 years ago in KwaZulu-Natal, South Africa. Antiquity 84, 635-648.

Louw, A.W., 1969. Bushman Rock Shelter, Ohrigstad, Eastern Transvaal: A Preliminary Investigation, 1965. The South African Archaeological Bulletin 24, 39-51.

Luedtke, B.E., 1992. An Archaeologist's Guide to Chert and Flint. Institute of Archaeology, University of California, Los Angeles.

Malan, B.D., 1947. Flake Tools and Artefacts in the -Fauresmith Transition in the Vaal River Valley. South African journal of Science XLIII, 350-362.

Marean, C.W., 2010. Pinnacle Point Cave 13B (Western Cape Province, South Africa) in context: The Cape Floral kingdom, shellfish, and modern human origins. Journal of Human Evolution 59, 425-443.

Marean, C.W., Assefa, Z., 1999. Zooarchaeological evidence for the faunal exploitation behavior of Neandertals and Early Modern Humans. Evolutionary Anthropology 8, 22-37.

Marean, C.W., Assefa, Z., 2005. The Middle and Upper Pleistocene African Record for the Biological and Behavioral Origins of Modern Humans., in: Stahl, A.B. (Ed.), : A Critical Introduction. Blackwell Publishing, Malden, MA, pp. 93-129.

Marean, C.W., Bar-Matthews, M., Bernatchez, J., Fisher, E., Goldberg, P., Herries, A.I.R., Jacobs, Z., Jerardino, A., Karkanas, P., Minichillo, T., Nilssen, P.J., Thompson, E., Watts, I., Williams, H.M., 2007. Early human use of marine resources and pigment in South Africa during the Middle Pleistocene. Nature 449, 905-908.

Marks, A.E., Monigal, K., 1995. Modeling the Production of Elongated Blanks from the Early Levantine Mousterian at Rosh Ein Mor, The Definition and Interpretation of Levallois Technology. Prehistory Press, Madison, pp. 267-278.

271

Martinón-Torres, M., Bermúdez de Castro, J.M., Gómez-Robles, A., Prado-Simón, L., Arsuaga, J.L., 2012. Morphological description and comparison of the dental remains from Atapuerca- Sima de los Huesos site (Spain). Journal of Human Evolution 62, 7-58.

Mason, R., 1962. Prehistory of the Transvaal: A Record of Human Activity. Witwatersrand University Press, Johannesburg.

Mason, R.J., 1959. Some South African Stone Age Cultures. Nature 183, 377-379.

Matmon, A., Ron, H., Chazan, M., Porat, N., Horwitz, L.K., 2012. Reconstructing the history of sediment deposition in caves: A case study from Wonderwerk Cave, South Africa. Geological Society of America Bulletin 124, 611-625.

McBrearty, S., 1987. Une Evaluation du Sangoen: Son Age, Sone Environment et son Rapport avec L'origine de L'Homo Sapiens. L'Anthropologie 91, 497-510.

McBrearty, S., 1988. The Sangoan-Lupemban and Middle Stone Age Sequence at the Muguruk Site, Western Kenya. World Archaeology 19, 388-420.

McBrearty, S., 1991. Recent Research in Western Kenya and its Implications for the Status of the Sangoan Industry, in: Clark, J.D. (Ed.), Cultural Beginnings: Approaches to Understanding Early Hominid Life-ways in the African Savanna. Bonn: Dr. Rudolf Habelt GMBH, pp. 159-176.

McBrearty, S., 1992. Sangoan Technology and Habitat at Simbi. Nyame Akuma 38, 34-39.

McBrearty, S., Bishop, L., Kingston, J., 1996. Variability in traces of Middle Pleistocene hominid behavior in the Kapthurin Formation, Baringo, Kenya. Journal of Human Evolution 30, 563-580.

McBrearty, S., Brooks, A.S., 2000. The revolution that wasn't: a new interpretation of the origin of modern human behavior. Journal of Human Evolution 39, 453-563.

McBrearty, S., Tryon, C.A., 2006. From Acheulean to Middle Stone Age in the Kapthurin Formation, Kenya, in: Hovers, E., Kuhn, S.L. (Eds.), Transitions Before the Transition: Evolution and Stability in the Middle Paleolithic and Middle Stone Age. Springer New York, pp. 257-276.

McCarthy, T., Rubridge, B., 2005. The Story of LIfe on Earth: A southern African perspective on a 4.6-billion-year journey. Stuik Publishers, Cape Town.

McDougall, I., Brown, F.H., 2006. Precise 40Ar/39Ar geochronology for the upper Koobi Fora Formation, Turkana Basin, northern Kenya. Journal of the Geological Society 163, 205-220.

McDougall, I., Brown, F.H., Fleagle, J.G., 2005. Stratigraphic placement and age of modern humans from Kibish, Ethiopia. Nature 433, 733-736.

McNabb, J., 2001. The shape of things to come: A speculative essay on the role of the Victoria West phenomenon at Canteen Koppie, during the South African Earlier Stone Age, A Very

272

Remote Period Indeed: Papers on the Palaeolithic presented to Derek Roe. Oxbow Books, Oxford, pp. 37-46.

McNabb, J., 2009. The ESA Stone Tool Assemblage from the Cave of Hearths, Beds 1-3, in: McNabb, J., Sinclair, A.G.M. (Eds.), The Cave of Hearths: Makapan Middle Pleistocene Research Project. Archaeopress, Oxford, UK, pp. 75-104.

McNabb, J., Beaumont, P.B., 2011. Report on the Archaeological Assemblages from Excavations by Peter Beaumont at Canteen Koppie, Northern Cape, South Africa. BAR International Series 2275, Archaeopress Oxford.

McNabb, J., Binyon, F., Hazelwood, L., 2004. The Large Cutting Tools from the South African Acheulean and the Question of Social Traditions. Current Anthropology 45, 653-677.

McPherron, S.P., Alemseged, Z., Marean, C.W., Wynn, J.G., Reed, D., Geraads, D., Bobe, R., Bearat, H.A., 2010. Evidence for stone-tool-assisted consumption of animal tissues before 3.39 million years ago at Dikika, Ethiopia. Nature 466, 857-860.

Mehlman, M.J., 1987. Provenience, age and associations of archaic Homo sapiens crania from Lake Eyasi, Tanzania. Journal of Archaeological Science 14, 133-162.

Merrick, H.V., Brown, F.H., Nash, W.P., 1994. Use and movement of obsidian in the Early and Middle Stone Ages of Kenya and northern Tanzania, Society, Culture, and Technology in Africa. MASCA 11, pp. 29-44.

Merrick, H.V., Brown, F.H.B., 1984. Obsidian sources and patterns of source utilization in Kenya and northern Tanzania: some initial findings. The African Archaeological Review 2, 129- 152.

Milo, R.G., 1998. Evidence for Hominid at Klasies River Mouth, South Africa, and its Implications for the Behaviour of Early Modern Humans. Journal of Archaeological Science 25, 99-133.

Minichillo, T., 2006. Raw material use and behavioral modernity: Howiesons Poort lithic foraging strategies. Journal of Human Evolution 50, 359-364.

Moen, H.F.G., 2006. The Olifantshoek Supergroup, in: Johnson, H.R., Anhaeusser, C.R., Thomas, R.J. (Eds.), The Geology of South Africa. Johannesburg/Pretoria, Geological Society of South Africa and the Council for Geoscience, pp. 319-324.

Mohapi, M., 2007. Rose Cottage Cave MSA Lithic Points: Does Technological Change Imply Change In Hunting Techniques? South African Archaeological Bulletin 62, 9-18.

Morgan, L.E., Renne, P.R., 2008. Diachronous dawn of Africa's Middle Stone Age: New 40Ar/39Ar ages from the Ethiopian Rift. Geology 36, 967-970.

Morgan, L.E., Renne, P.R., Kieffer, G., Piperno, M., Gallotti, R., Raynal, J.-P., 2012. A chronological framework for a long and persistent archaeological record: , Ethiopia. Journal of Human Evolution 62, 104-115.

273

Morris, D., Beaumont, P.B., 2004. Archaeology in the Northern Cape: Some Key Sites. McGregor Museum, Kimberley.

Mounier, A., Marchal, F., Condemi, S., 2009. Is Homo heidelbergensis a distinct species? New insight on the Mauer mandible. Journal of Human Evolution 56, 219-246.

Mourre, V., Villa, P., Henshilwood, C.S., 2010. Early Use of Pressure Flaking on Lithic Artifacts at Blombos Cave, South Africa. Science 330, 659-662.

Mturi, A.A., 1976. New hominid from Lake Ndutu, Tanzania. Nature 262, 484-485.

Naroll, R., 1962. Floor area and settlement population American Antiquity 27.

Negash, A., Shackley, M.S., 2006. Geochemical provenance of obsidian artefacts from the MSA site of Porc Epic, Ethiopia. Archaeometry 48, 1-12.

Odell, G.H., Odell-Vereecken, F., 1980. Verifying the reliability of lithic use-wear assessments by blind tests: the low-power approach. Journal of Field Archaeology 7, 87-120.

Oswalt, W.H., 1976. An Anthropological Analysis of Food-getting technology. John Wiley and Sons, Inc., New York.

Pargeter, J., 2007. Howiesons Poort Segments as Hunting Weapons: Experiments with Replicated Projectiles. The South African Archaeological Bulletin 62, 147-153.

Pargeter, J., 2011a. Assessing the macrofracture method for identifying Stone Age hunting weaponry. Journal of Archaeological Science 38, 2882-2888.

Pargeter, J., 2011b. Human and cattle trampling experiments in Malawi to understand macrofracture formation on Stone Age hunting weaponry. ANTIQUITY 85, 327.

Parkington, J., Poggenpoel, C., Rigaud, J.-P., Tixier, P.-J., 2005. From tool to symbol: the behavioral context of intentionally marked ostrich eggshell from Diepkloof, Western Cape, From tools to symbols: from early hominids to modern humans. Wits University Press, Johannesburg, pp. 475-492.

Pleurdeau, D., 2006. Human Technical Behavior in the African Middle Stone Age: The Lithic Assemblage of Porc-Epic Cave (Dire Dawa, Ethiopia). African Archaeological Review 22, 177- 197.

Porat, N., Chazan, M., Grün, R., Aubert, M., Eisenmann, V., Horwitz, L.K., 2010. New radiometric ages for the Fauresmith industry from Kathu Pan, southern Africa: Implications for the Earlier to Middle Stone Age transition. Journal of Archaeological Science 37, 269-283.

Portenga, E.W., Bierman, P.R., 2011. Understanding Earth's eroding surface with 10Be. GSA Today 21, 4-10.

Potts, R., 1998. Variability Selection in Hominid Evolution. Evolutionary Anthropology 7, 81- 96.

274

Powell, A., Shennan, S., Thomas, M.G., 2009. Late Pleistocene Demography and the Appearance of Modern Human Behavior. Science 324, 1298-1301.

Purdy, B.A., Clark, D.E., 1987. Weathering of Inorganic materials: Dating and other applications, in: Schiffer, M.B. (Ed.), Advances in archaeological Method and Theory. Academic, New York.

Raynal, J.P., Sbihi Alaoui, F.Z., Geraads, D., Magoga, L., Mohi, A., 2001. The earliest occupation of North-Africa: the Moroccan perspective. Quaternary International 75, 65-75.

Revillion, S., Tuffreau, A., 1994. Valeur et signification du débitage laminaire du gisement paleolithic moyen de Seclin (Nord), in: Revillion, S., Tuffreau, A. (Eds.), Les industries laminaires au Paléolithique moyen. Dossier de Documentation Archéologique 18. CNRA Éditions, Paris, pp. 19 -43.

Reynolds, S.C., Clarke, R.J., Kuman, K.A., 2007. The view from the Lincoln Cave: mid- to late Pleistocene fossil deposits from Sterkfontein hominid site, South Africa. Journal of Human Evolution 53, 260-271.

Richerson, P.J., Boyd, R., 2005. Not by Genes Alone: How Culture Transformed Human Evolution. University of Chicago Press, Chicago.

Riel-Salvatore, J., 2009. What Is a ‘Transitional’ Industry? The Uluzzian of Southern Italy as a Case Study, in: Camps, M., Chauhan, P. (Eds.), Sourcebook of Paleolithic Transitions. Springer, New York, pp. 377-396.

Rightmire, G.P., 1983. The Lake Ndutu cranium and early Homo Sapiens in Africa. American Journal of Physical Anthropology 61, 245-254.

Rightmire, G.P., 2001. Patterns of hominid evolution and dispersal in the Middle Pleistocene. Quaternary International 75, 77-84.

Rightmire, G.P., 2004. Brain size and encephalization in early to Mid-Pleistocene Homo. American Journal of Physical Anthropology 124, 109-123.

Rightmire, G.P., 2008. Homo in the middle pleistocene: Hypodigms, variation, and species recognition. Evolutionary Anthropology: Issues, News, and Reviews 17, 8-21.

Rightmire, G.P., 2009. Middle and Later Pleistocene Hominins in Africa and Southwest Asia. Proc. Natl. Acad. Sci. U. S. A. 106, 16046-16050.

Rightmire, G.P., 2012. The evolution of cranial form in mid-Pleistocene Homo. South African Journal of Science 108, 1-10.

Roberts, M.B., Parfitt, S.A., 1999. Boxgrove: A Middle Pleistocene Hominid Site at Earthen Quarry, Boxgrove, West Sussex. English Heritage, London.

Rosas, A., 2001. Occurrence of Neanderthal features in mandibles from the Atapuerca-SH site. American Journal of Physical Anthropology 114, 74-91.

275

Rots, V., 2003. Towards an understanding of hafting: the macro- and microscopic evidence. Antiquity 77, 805-815.

Rots, V., in press. Insights into early Middle Palaeolithic tool use and hafting in Western Europe. The functional analysis of level IIa of the early Middle Palaeolithic site of Biache-Saint-Vaast (France). Journal of Archaeological Science, DOI: 10.1016/j.jas.2012.06.042.

Rots, V., Van Peer, P., Vermeersch, P.M., 2011. Aspects of tool production, use, and hafting in Palaeolithic assemblages from Northeast Africa. Journal of Human Evolution 60, 637-664.

Ruff, C.B., Trinkaus, E., Holliday, T.W., 1997. Body mass and encephalization in Pleistocene Homo. Nature 387, 173-176.

Rüther, H., Chazan, M., Schroeder, R., Neeser, R., Held, C., Walker, S.J., Matmon, A., Horwitz, L.K., 2009. Laser scanning for conservation and research of African cultural heritage sites: the case study of Wonderwerk Cave, South Africa. Journal of Archaeological Science 36, 1847- 1856.

Sackett, J.R., 1982. Approaches to Style in Lithic Archaeology. Journal of Anthropological Archaeology 1, 59-112.

Sackett, J.R., 1985. Style and ethnicity in the Kalahari: a reply to Wiessner. American Antiquity 50, 154-159.

Sampson, C.G., 1972. The Stone Age Industries of the Orange River Scheme and South Africa. National Museum, Bloemfontein Memoir No. 6.

Sampson, C.G., 1974a. The Stone Age archaeology of Southern Africa. Academic Press Inc., New York.

Sampson, G.C., 1974b. The Stone Age archaeology of Southern Africa. Academic Press Inc., New York.

Schick, K., 1984. Processes of Paleolithic Site Formation: An Experimental Study. PhD Dissertation. University of California, Berkeley.

Schoville, B.J., 2010. Frequency and Distribution of Edge Damage on Middle Stone Age Lithic Points, Pinnacle Point 13B, South Africa. Journal of Human Evolution 59, 378-391.

Schoville, B.J., Brown, K.S., 2010. Comparing Lithic Assemblage Edge Damage Distributions: Examples from the Late Pleistocene and Preliminary Experimental Results. vis-a-vis: Explorations in Anthropology 10, 34-49.

Sharon, G., 2008. The impact of raw material on Acheulian large flake production. Journal of Archaeological Science 35, 1329-1344.

Sharon, G., 2009. Acheulian Giant‐Core Technology: A Worldwide Perspective. Current Anthropology 50, 335-367.

276

Shea, J., 1991. The behavioral significance of Levantine industrial variability. Harvard University.

Shea, J., 2011a. Refuting a Myth About Human Origins. American Scientist 99, 128-135.

Shea, J., in press. Lithic Modes A–I: A New Framework for Describing Global-Scale Variation in Stone Tool Technology Illustrated with Evidence from the East Mediterranean Levant. Journal of Archaeological Method and Theory DOI 10.1007/s10816-012-9128-5.

Shea, J., Davis, Z., Brown, K., 2001. Experimental Tests of Middle Palaeolithic Spear Points Using a Calibrated Crossbow. Journal of Archaeological Science 28, 807-816.

Shea, J.J., 2006. The origins of lithic technology: evidence from Africa, the Levant, and Europe. Journal of Archaeological Science 33, 823-846.

Shea, J.J., 2008. The Middle Stone Age archaeology of the Lower Omo Valley Kibish Formation: Excavations, lithic assemblages, and inferred patterns of early Homo sapiens behavior. Journal of Human Evolution 55, 448-485.

Shea, J.J., 2011b. Homo sapiens Is as Homo sapiens Was: Behavioral Variability versus “Behavioral Modernity” in Paleolithic Archaeology. Current Anthropology 52, 1-35.

Shea, J.J., Klenck, J.D., 1993. An Experimental Investigation of the Effects of Trampling on the Results of Lithic Microwear Analysis. Journal of Archaeological Science 20, 175-194.

Shennan, S.J., 1997. Quantifying Archaeology, 2nd ed. University of Press, Iowa City.

Sheppard, P.J., Kleindienst, M.R., 1996. Technological Change in the Earlier and Middle Stone Age of Kalambo Falls (Zambia). The African Archaeological Review 13, 171-196.

Shimelmitz, R., Barkai, R., Gopher, A., 2011. Systematic blade production at late Lower Paleolithic (400–200 kyr) Qesem Cave, Israel. Journal of Human Evolution 61, 458-479.

Shipton, C., 2010. Imitation and Shared Intentionality in the Acheulean. Cambridge Archaeological Journal 20, 197-210.

Shott, M., 1986. Technological Organization and Settlement Mobility: An Ethnographic Examination. Journal Of Anthropological Research 42, 15-51.

Sinclair, A., 2009. The MSA Stone Tool Assemblage from the Cave of Hearths, Beds 4-9, in: McNabb, J., Sinclair, A. (Eds.), The Cave of Hearths: Makapan Middle Pleistocene Research Project. Archaeopress, Oxford, UK, pp. 75-104.

Singer, R., Wymer, J., 1968. Archaeological Investigations at the Saldanha Skull Site in South Africa. The South African Archaeological Bulletin 23, 63-74.

Singer, R., Wymer, J., 1982. The Middle Stone Age at Klasies River Mouth in South Africa. University of Chicago Press, Chicago.

277

Sisk, M.L., Shea, J.J., 2009. Experimental use and quantitative performance analysis of triangular flakes (Levallois points) used as arrowheads. Journal of Archaeological Science 36, 2039-2047.

Sisk, M.L., Shea, J.J., 2011. The African Origin of Complex Projectile Technology: An Analysis Using Tip Cross-Sectional Area and Perimeter. International Journal of Evolutionary Biology 2011.

Soriano, S., Villa, P., Wadley, L., 2007. Blade technology and tool forms in the Middle Stone Age of South Africa: the Howiesons Poort and post-Howiesons Poort at Rose Cottage Cave. Journal of Archaeological Science 34, 681-703.

Steward, J.H., 1955. Theory of culture change:The methodology of multilinear evolution. University of Press, Urbana.

Stout, D., Quade, J., Semaw, S., Rogers, M.J., Levin, N.E., 2005. Raw material selectivity of the earliest stone toolmakers at Gona, Afar, Ethiopia. Journal of Human Evolution 48, 365-380.

Stout, D., Semaw, S., Rogers, M.J., Cauche, D., 2010. Technological variation in the earliest Oldowan from Gona, Afar, Ethiopia. Journal of Human Evolution 58, 474-491.

Stringer, C., 2012. The status of Homo heidelbergensis (Schoetensack 1908). Evolutionary Anthropology: Issues, News, and Reviews 21, 101-107.

Szabo, B.J., Butzer, K.W., 1979. Uranium-series dating of lacustrine limestones from pan deposits with final Acheulian assemblage at Rooidam, Kimberley district, South Africa. Quaternary Research 11, 257-260.

Taylor, N., 2011. The origins of hunting and gathering in the Congo basin: a perspective on the Middle Stone Age Lupemban industry. Before Farming 2011/1, 1-20.

Tennie, C., Call, J., Tomasello, M., 2009. Ratcheting up the ratchet: on the evolution of cumulative culture. Philosophical Transactions of the Royal Society B: Biological Sciences 364, 2405-2415.

Thieme, H., 1997. Lower Palaeolithic hunting spears from Germany. Nature 385, 807-810.

Thompson, E., Williams, H.M., Minichillo, T., 2010. Middle and late Pleistocene Middle Stone Age lithic technology from Pinnacle Point 13B (Mossel Bay, Western Cape Province, South Africa). Journal of Human Evolution 59, 358-377.

Tomasello, M., 1999. The Cultural Origins of Human Cognition. Harvard University Press, Cambridge, MA.

Tomasello, M., Carpenter, M., Call, J., Behne, T., Moll, H., 2005. Understanding and Sharing Intentions: The Origins of Cultural Cognition. Behavioral and Brain Sciences 28, 675-735.

Tomasello, M., Kruger, A.C., Ratner, H.H., 1993. Cultural learning. Behavioral and Brain Sciences 16, 495-511.

278

Tringham, R., Cooper, G., Odell, G., Voytek, B., Whitman, A., 1974. Experimentation in the Formation of Edge Damage: A New Approach to Lithic Analysis. Journal of Field Archaeology 1, 171-196.

Tryon, C.A., 2006. "Early" Middle Stone Age Lithic Technology of the Kapthurin Formation (Kenya). Current Anthropology 47, 367-375.

Tryon, C.A., McBrearty, S., 2006. Tephrostratigraphy of the Bedded Tuff Member (Kapthurin Formation, Kenya) and the nature of archaeological change in the later middle Pleistocene. Quaternary Research 65, 492-492.

Tryon, C.A., McBrearty, S., Texier, P.-J., 2005. Levallois Lithic Technology from the Kapthurin Formation, Kenya: Acheulian Origin and Middle Stone Age Diversity. African Archaeological Review 22, 199-199.

Underhill, D., 2011. The study of the Fauresmith: a review. South African Archaeological Bulletin 66, 15-26. van der Elst, W., 1950. Some Implements from the Van der Elst Donga. The South African Archaeological Bulletin 5, 43-50.

Van Peer, P., 1992. The Levallois Reduction Strategy. Prehistory Press, Madison.

Van Peer, P., Fullagar, R., Stokes, S., Bailey, R.M., Moeyersons, J., Steenhoudt, F., Geerts, A., Vanderbeken, T., De Dapper, M., Geus, F., 2003. The Early to Middle Stone Age Transition and the Emergence of Modern Human Behaviour at site 8-B-11, Sai Island, Sudan. Journal of Human Evolution 45, 187-193.

Van Peer, P., Rots, V., Vermeersch, P.M., 2008. A Wasted Effort at the Quarry: Wear Analysis and Interpretation of an MSA Lanceolate Point from Taramsa-8, Egypt. PaleoAnthropology, 234−250.

Van Riet Lowe, C., 1927. The Fauresmith Coup de Poing. South African Journal of Science XXIV, 502-505.

Van Riet Lowe, C., 1929. A Few Notes on the Archaeology of Sheppard Island, in: Goodwin, A.J.H., Van Riet Lowe, C. (Eds.), The Stone Age Cultures of South Africa. AMS Press, New York, pp. 235-243.

Van Riet Lowe, C., 1933. Further Notes on the Fauresmith Culture. South African Journal of Science XXX, 527-529.

Van Riet Lowe, C., 1935. Implementiferous Gravels of the Vall River at Riverveiw Estates. Nature 136, 53-56.

Van Riet Lowe, C., 1937. The Archaeology of the Vaal River Basin in: Söhnge, P.G., Visser, D.J.L., Van Riet Lowe, C. (Eds.), The Geology and Archaeology of the Vaal River Basin, Geological Survey Memoirs 35. Government Printer, Pretoria, pp. 61-164.

279

Van Riet Lowe, C., 1952. The Vaal River Chronology: An Up-to-Date Summary. The South African Archaeological Bulletin 7, 135-149.

Vermeersch, P.M., Paulisson, E., Philip, V.P., 1990. Palaeolithic Chert Exploitation in the Limestone stretch of the Egyptian Valley. The African Archaeological Review 8, 77-102.

Villa, P., Boscato, P., Ranaldo, F., Ronchitelli, A., 2009a. Stone tools for the hunt: points with impact scars from a Middle Paleolithic site in southern Italy. Journal of Archaeological Science 36, 850-859.

Villa, P., Delagnes, A., Wadley, L., 2005. A late Middle Stone Age artifact assemblage from Sibudu (KwaZulu-Natal): comparisons with the European Middle Paleolithic. Journal of Archaeological Science 32, 399-422.

Villa, P., Lenoir, M., 2006. Hunting weapons of the Middle Stone Age and the Middle Palaeolithic; spear points from Sibudu, Rose Cottage and Bouheben. Southern African Humanities 18, 89-122.

Villa, P., Soressi, M., Henshilwood, C.S., Mourre, V., 2009b. The Still Bay points of Blombos Cave (South Africa). Journal of Archaeological Science 36, 441-460.

Villa, P., Soriano, S., Teyssandier, N., Wurz, S., 2010. The Howiesons Poort and MSA III at Klasies River main site, Cave 1A. Journal of Archaeological Science 37, 630-655.

Villa, P., Soriano, S., Tsanova, T., Degano, I., Higham, T.F.G., d’Errico, F., Backwell, L., Lucejko, J.J., Colombini, M.P., Beaumont, P.B., 2012. Border Cave and the beginning of the Later Stone Age in South Africa. Proceedings of the National Academy of Sciences 109, 13208- 13213.

Volman, T.P., 1984. Early prehistory of southern Africa, Southern African Prehistory and Paleoenvironments. A.A. Balkema. Earth, Rotterdam, pp. 169-220.

Wadley, L., 2001. What is cultural modernity? A general view and a South African perspective from Rose Cottage Cave. Cambridge Archaeological Journal 11, 201-221.

Wadley, L., 2003. How some archaeologists recognize culturally modern behaviour. South African Journal of Science 99, 247-250.

Wadley, L., 2010. Were snares and traps used in the Middle Stone Age and does it matter? A review and a case study from Sibudu, South Africa. Journal of Human Evolution 58, 179-192.

Wadley, L., Hodgskiss, T., Grant, M., 2009. Implications for complex cognition from the hafting of tools with compound adhesives in the Middle Stone Age, South Africa. Proceedings of the National Academy of Sciences 106, 9590-9594.

Wadley, L., McNabb, J., 2009. Chapter 1: Introduction, in: McNabb, J., Sinclair, A.G.M. (Eds.), The Cave of Hearths: Makapan Middle Pleistocene Research Project. Field Research by Anthony Sinclair and Patrick Quinney, 1996-2001. BAR International Series, pp. 1-11.

280

Wadley, L., Mohapi, M., 2008. A Segment is not a Monolith: evidence from the Howiesons Poort of Sibudu, South Africa. Journal of Archaeological Science 35, 2594-2605.

Walter, R.C., Buffler, R.T., Bruggemann, J.H., Guillaume, M.M.M., Berhe, S.M., Negassi, B., Libsekal, Y., Cheng, H., Edwards, R.L., von Cosel, R., Neraudeau, D., Gagnon, M., 2000. Early human occupation of the Red Sea coast of Eritrea during the last interglacial. Nature 405, 65-69.

Watts, I., 2002. Ochre in the Middle Stone Age of southern Africa: ritualised display or hide preservative? South African Archaeological Bulletin 57, 1-14.

Waweru, V., 2002. New Excavations at the Middle Stone Age Cartwright's site, Keny. Nyame Akuma 58, 26-33.

Wendorf, F., Schild, R., Close, A.E., 1993. Egypt during the Last Interglacial: The Middle Paleolithic of Bir Tarfawi and Bir Sahara East. Plenum Press, New York.

White, T.D., Asfaw, B., DeGusta, D., Gilbert, H., Richards, G.D., Suwa, G., Clark Howell, F., 2003. Pleistocene Homo sapiens from Middle Awash, Ethiopia. Nature 423, 742(746)-742(746).

Wiessner, P., 1974. A functional estimate of population from floor area. American Antiquity 39.

Wiessner, P., 1983. Style and social information in Kalahari San projectile points. American Antiquity 48, 253-276.

Wilkins, J., 2010. Style, Symboling, and Interaction in Middle Stone Age Society. vis-a-vis: Explorations in Anthropology 10, 102-125.

Wilkins, J., Chazan, M., 2012. Blade production ~500 thousand years ago at Kathu Pan 1, South Africa: Support for a multiple origins hypothesis for early Middle Pleistocene blade technologies. Journal of Archaeological Science 39, 1883-1900.

Wilkins, J., Pollarolo, L., Kuman, K., 2010. Prepared core reduction at the site of Kudu Koppie in northern South Africa: temporal patterns across the Earlier and Middle Stone Age boundary. Journal Of Archaeological Science 37, 1279-1292.

Wilkins, J., Schoville, B.J., Brown, K.S., Chazan, M., 2012. Evidence for Early Hafted Hunting Technology. Science 338, 942-946.

Willoughby, P., 2007. The Evolution Of Modern Humans In Africa: A Comprehensive Guide. AltaMira Press, Plymouth.

Winter, L.K., 2000. The Howiesons Poort: A case of conformity or modernity? Master of Sciences, University of the Witwatersrand.

Wobst, H.M., 1977. Stylistic Behavior and Information Exchange, For the Director: Research Essays in Honor of James B. Griffen. University of Michigan Museum of Anthropology, Anthropological Papers, Ann Arbour, pp. 317-342.

281

Wood, B., Leakey, M., 2011. The Omo‐Turkana Basin Fossil Hominins and Their Contribution to Our Understanding of Human Evolution in Africa. Evolutionary Anthropology: Issues, News, and Reviews 20, 264-292.

Wurz, S., 2000. The Middle Stone Age sequence at Klasies River, South Africa. University of Stellenbosch.

Wurz, S., 2002. Variability in the Middle Stone Age Lithic Sequence, 115,000-60,000 Years Ago at Klasies River, South Africa. Journal of Archaeological Science 29, 1001-1015.

Wynn, T., Coolidge, F.L., 2011. The Implications of the Working Memory Model for the Evolution of Modern Cognition. International Journal of Evolutionary Biology 2011.

Yellen, J., Brooks, A., Helgren, D., Tappen, M., Ambrose, S., Bonnefille, R., Feathers, J., Goodfriend, G., Ludwig, K., Renne, P., Stewart, K., 2005. The Archaeology of Aduma Middle Stone Age Sites in the Awash Valley, Ethiopia. PaleoAnthropology 3, 25-100.