<<

UNIVERSITY OF CALGARY

Middle Stone Age Lithic Technology at Mvumu, Niassa,

by

Tim Bennett

A THESIS

SUBMITTED TO THE FACULTY OF GRADUATE STUDIES

IN PARTIAL FULFILMENT OF THE REQUIREMENTS FOR THE

DEGREE OF MASTER OF ARTS

DEPARTMENT OF ARCHAEOLOGY

CALGARY, ALBERTA

June 2011

© Tim Bennett 2011 Library and Archives Bibliothèque et Archives Canada Published Heritage Direction du Branch Patrimoine de l'édition

395 Wellington Street 395, rue Wellington Ottawa ON K1A 0N4 Ottawa ON K1A 0N4 Canada Canada Your file Votre référence ISBN: 978-0-494-83608-8

Our file Notre référence ISBN: 978-0-494-83608-8

NOTICE: AVIS: The author has granted a non- L'auteur a accordé une licence non exclusive exclusive license allowing Library and permettant à la Bibliothèque et Archives Archives Canada to reproduce, Canada de reproduire, publier, archiver, publish, archive, preserve, conserve, sauvegarder, conserver, transmettre au public communicate to the public by par télécommunication ou par l'Internet, prêter, telecommunication or on the Internet, distribuer et vendre des thèses partout dans le loan, distrbute and sell theses monde, à des fins commerciales ou autres, sur worldwide, for commercial or non- support microforme, papier, électronique et/ou commercial purposes, in microform, autres formats. paper, electronic and/or any other formats.

The author retains copyright L'auteur conserve la propriété du droit d'auteur ownership and moral rights in this et des droits moraux qui protege cette thèse. Ni thesis. Neither the thesis nor la thèse ni des extraits substantiels de celle-ci substantial extracts from it may be ne doivent être imprimés ou autrement printed or otherwise reproduced reproduits sans son autorisation. without the author's permission.

In compliance with the Canadian Conformément à la loi canadienne sur la Privacy Act some supporting forms protection de la vie privée, quelques may have been removed from this formulaires secondaires ont été enlevés de thesis. cette thèse.

While these forms may be included Bien que ces formulaires aient inclus dans in the document page count, their la pagination, il n'y aura aucun contenu removal does not represent any loss manquant. of content from the thesis. Abstract

For over ten years the origin of modern behaviour has been the focus of the majority of (MSA) research in Africa. However, the analytical value of the modern behaviour concept has recently been questioned. Critics argue that studying less abstract MSA ecology and behavioural variability would better serve our understanding of the period, but such studies are hindered by the fragmentary and unevenly distributed nature of the fundamental aspects of MSA archaeology. This thesis makes a step forward in filling the gaps in this record by presenting lithic technology from a final MSA open air site called Mvumu in the Niassa basin of northern

Mozambique. The results suggest that Mvumu fits well with other Niassa basin sites and sites throughout the south-central African woodlands and forests suggests that Mvumu‟s lithic technology belongs to an overarching technological complex throughout the greater area.

ii Acknowledgements

I am greatly indebted to Julio Mercader for his support and guidance throughout the writing of this thesis. Your vision for the Projecto Património Arqueológico e

Cultural (Archaeological and Cultural Patrimony Project or PAC) that led to this thesis and devotion to Mozambican archaeology is inspirational. I would also like to thank the remaining members of my committee, Brian Kooyman and Len Hills, and the entirety of the University of Calgary Department of Archaeology for their support over the years and into the future as I embark on my Ph.D. research. Dale Walde also deserves special mention for his support of my Ph.D. application and as a co-author on several papers published during my MA program. John Gosse of the Dalhousie Geochronology Center made an important contribution to this thesis by providing dates for Mvumu and my fellow PAC members Mussa Raja, Steven Simpson, and Arianna Fogelman also deserve recognition for their role in bringing my research to this point. Much appreciated financial support came from both the University of Calgary and the Social Sciences and

Humanities Research Council of Canada.

My research benefited from the support of the Department of Archaeology and

Anthropology at Eduardo Mondlane in Maputo, especially Professors Hilário Madiquida and Solange Macamo. Their involvement with this research and insight into practicing archaeology in Mozambique has been instrumental in my professional development.

This research would have been impossible without archaeological permits from the

Mozambican Ministry of Education and Culture (03-2003 and 01-2007) and temporary artifact export permits from the Mozambican Chamber of Commerce (Certificate of

Origin no. 0134) and the Mozambican Customs Service (Export License no. 24399). I

iii also need to thank Lourenço Thawe and his family, Justin and Sofia Sondergaard, and the numerous workers, friends, and authorities in Niassa for their significant contributions to this work.

And lastly I need to acknowledge my family and friends for their support.

Especially my wife Ingrid Dinsmore for putting up with me disappearing to Africa for months at a time and “barely” batting an eye when I told her I was applying to continue in the Ph.D. program. Also needing special recognition are my parents Bob and Brenda

Bennett, band mates from The Drive, Inquisition, and Rx, and Chris Esselmont for his involvement as a co-author and his excellent suggestion, which sadly did not make it into the final version of this text, that my introduction should read “I found some stones, here‟s more than you ever wanted to know about them”. Maybe I‟ll find a place for those words in my dissertation.

iv Dedication

For Ingrid Dinsmore

Throughout it all you were the light that led me home

v Table of Contents

Approval Page ...... ii Abstract ...... ii Acknowledgements ...... iii Dedication ...... v Table of Contents ...... vi List of Tables ...... viii List of Figures and Illustrations ...... ix

CHAPTER ONE: INTRODUCTION ...... 1

CHAPTER TWO: THE AFRICAN MIDDLE STONE AGE AND THE ORIGIN OF MODERN ...... 5 2.1 – Theories regarding the origin of Homo sapiens ...... 6 2.1.1 – Multiregional theories ...... 8 2.1.2 – Out of Africa theories ...... 9 2.1.3 – Fossil evidence for the origins of H. sapiens ...... 10 2.1.4 – Genetic evidence for the origins of H. sapiens ...... 12 2.2 Archaeological evidence for complexity in Middle Stone Age behaviour ...... 14 2.3 – Middle Stone Age technological behaviour ...... 22

CHAPTER THREE: ARCHAEOLOGY IN MOZAMBIQUE ...... 37 3.1 – History of archaeology in Mozambique ...... 39 3.2 – Archaeology in the Niassa Rift ...... 44 3.2.1 – Physiography of the study area ...... 46 3.2.2 – Archaeology in Sanga District ...... 50 3.2.3 Archaeology in Lago District ...... 53 3.3 – Mvumu ...... 55

CHAPTER FOUR: ANALYTICAL METHODS AND LITHIC CLASSIFICATION SCHEME ...... 63 4.1 – Study assemblage and methodological approach ...... 63 4.2 – Raw material analytical methods ...... 63 4.3 Lithic Artifact Analytical Methodology ...... 64 4.3.1 – Core classification scheme ...... 65 4.3.1.1 – Simple Cores ...... 66 4.3.1.2 – Prepared cores ...... 68 4.3.2 – Debitage classification scheme ...... 71 4.3.2.1 – Core preparation products ...... 75 4.3.2.2 – Angular waste ...... 76 4.3.3 – Tool classification scheme ...... 77 4.3.3.1 – Scrapers ...... 77 4.3.3.2 – Awls ...... 81 4.3.3.3 – Points ...... 83 4.3.3.4 – Other tools ...... 85 4.4 – Metric and qualitative data ...... 88

vi 4.4.1 Core metrics and qualitative traits ...... 89 4.4.2 – Debitage metrics and qualitative traits ...... 91 4.4.3 – Tool metrics and qualitative traits ...... 93 4.5 – Experimental reduction ...... 94

CHAPTER FIVE: MIDDLE STONE AGE LITHICS FROM MVUMU ...... 97 5.1 – The Mvumu Lithic Study Assemblage ...... 97 5.2 – Raw Materials ...... 98 5.3 – Artifact analysis results ...... 100 5.3.1 – Cores and core reduction ...... 100 5.3.2 – Debitage ...... 108 5.3.3 – Tools ...... 112 5.3.3.1 – Scrapers ...... 113 5.3.3.2 – Points ...... 122 5.4 – Experimental reduction results ...... 126

CHAPTER SIX: DISCUSSION ...... 129 6.1 – Mvumu depositional context and site fidelity ...... 129 6.2 – Technological behaviour at Mvumu ...... 132 6.2.1 – Raw Materials ...... 132 6.2.2 – Lithic Reduction Strategies at Mvumu ...... 138 6.2.3 – Debitage ...... 144 6.2.4 – Tools ...... 151 6.3 – Behavioural implications of lithic technology at Mvumu ...... 158 6.4 – Comparisons with temporally and geographically related sites ...... 163

CHAPTER SEVEN: CONCLUSIONS ...... 179

REFERENCES ...... 182

vii List of Tables

Table 2.1 – Behavioural traits indicative of Modern Human Behaviour (adapted from McBrearty and Brooks 2000) ...... 15

Table 5.1 – Assemblage Composition ...... 97

Table 5.2 – Raw Material Types ...... 99

Table 5.3 – Core Types ...... 101

Table 5.4 – Debitage Types ...... 109

Table 5.5 – Tool Types ...... 114

Table 6.1 - Size and mass distributions for the experimental reduction of cobbles ...... 131

Table 6.2 – Cortex by Size Fraction for Flakes, Flake Fragments, and Truncations ...... 150

viii List of Figures and Illustrations

Figure 2.1 – Step-Wise Emergence of Modern Behaviours (source McBrearty and Brooks 2000:530)...... 19

Figure 3.1 – Study Area Location (Mercader et al, in prep) ...... 38

Figure 3.2 – PAC Survey Area and Study Area Geology (Mercader et al, in prep) ...... 45

Figure 3.3 – Niassa Sediment Column Profiles (Mercader et al, in prep) ...... 48

Figure 3.4 – Niassa Basin Profile through Study Area (adapted from Mercader et al 2010) ...... 49

Figure 3.5 – Main Excavation Trench, Study Assemblage Source Units, and In-Situ Artifact locations (Mercader et al, in prep) ...... 56

Figure 3.6 – Mvumu Stratigraphy (Mercader et al, in prep) ...... 57

Figure 3.7 – Mvumu Stone line (photograph by J. Mercader, modified by T. Bennett) .. 59

Figure 4.1 – The Truncation Technique (adapted from Moore et al 2009) ...... 74

Figure 4.2 – Demi-Cone on Truncated Piece...... 75

Figure 4.3 – Experimental Knapping. A) Lake Cobble, B) Hard Hammer Percussion, C) Truncation ...... 96

Figure 5.1 – Simple Cores: A) Simple Core, B) Elongated High Backed Core, C) Simple Core, D) Simple Core ...... 103

Figure 5.2 – Prepared Cores: A) Levallois, B) Preferential Discoidal Recurrent, C) Preferential Discoidal Single, D) Discoidal, E) Unifacial Radial ...... 104

Figure 5.3 – Scrapers: A) Convex Side on Flake, B) Convex Side on Flake, C) Semi- Circular, D) Denticulate, E) Bevel-Based on Flake, F) Bevel-Based on Core Fragment, G) Convex Side on Core, H) Denticulate Core, I) Convex End on Flake, J) Convex End on Core Fragment, K) Core ...... 116

Figure 5.4 – Awls: A) - D) Single Truncation, E) - G) Double Truncation ...... 120

Figure 5.5 – Points: A) - C) Corner-Struck, D) Base-Struck, E) - F) Levallois ...... 122

Figure 5.6 – Other Tools: A) and B) Crescents, C) Notch, D) - H) Snapped Retouched Pieces ...... 125

Figure 5.7 – Experimental Replica Tools: A) and B) Discoidal Cores, C) Corner-Struck Point, and D) Quartz Corner-Struck Point ...... 127

ix Figure 6.1 – Quartz Coverage on the Slopes near Lake Niassa (Photograph J. Mercader) ...... 133

x 1

Chapter One: Introduction

The Middle Stone Age (MSA) represents a milestone in the development of the human condition as we know it today. It stands out as particularly relevant to humanity in our current day and age because of two key developments that form the foundation of the human experience as we know it: 1) the origin of our species, Homo sapiens, and 2) the emergence of modern human behaviour patterns (Shea 2011). Over the past decade,

MSA research has been dominated by aspects of these broad topics. When did H. sapiens first appear in the fossil record (Clark et al 2003; White et al 2003; McDougall et al 2005; McDougall et al 2008; Pearson et al 2008)? Where are the earliest H. sapiens found (Clark 2003; White et al 2003; McDougall et al 2005; McDougall et al 2008;

Pearson et al 2008) and by what evolutionary mechanisms did they arise (Aiello 1993;

Stringer 2001)? Likewise, when and where did modern behaviour patterns first come to define humanity (McBrearty and Brooks 2000; Henshilwood and Marean 2003; Klein

2009)? Did they emerge in a behavioural revolution as a complete package (Klein 2009) or slowly accrete over time (McBrearty and Brooks 2000)? If it was a slow coalescence of traits that led to behavioural modernity did they emerge in an isolated corner of Africa with unique conditions that selected for these new behaviours (McBrearty and Brooks

2000) or was it an amalgamation of novel behaviours from throughout the continent that became the signature of modernity (McBrearty and Brooks 2000)? And what does it mean to be modern anyway (Basell 2008; Shea 2011)?

The abstract nature of such questions has led to the development of a wide range of theories to explain these phenomena (see chapter 2), but it has also led some

2

researchers to question whether we are approaching the study of MSA behaviour in an appropriate way (Shea 2011) or if we need to take a step back and evaluate whether we have overextended ourselves by pursuing such topics without first having adequate grounding in the conditions under which they developed (Basell 2008). This notion brings to light a valid point that is often disregarded in MSA research: despite the amount of research attention the period has recently received our conclusions are still based on limited, fragmentary data from specific ecoregions that are used to formulate theories that are applied pancontinentally. This is true even when it comes to some of the most fundamental aspects of MSA archaeology that have long research histories. One such aspect of MSA research is lithic technology. This thesis aims to fill one of the gaps that exists in our knowledge of MSA lithic technology by presenting the first detailed quantitative technological description of a final MSA lithic industry from an open air site called Mvumu in Niassa, northern Mozambique.

The Mozambican archaeological record on the whole is relatively poorly known compared to adjacent countries; , which lays immediately south of

Mozambique, is arguably the most thoroughly investigated part of Africa in terms of

Stone Age archaeology. This is, however, by no means a reflection of its importance to our understanding of the MSA. Mozambique is located at the southern extent of the

Great African Rift System. This places it in a key location between Southern, East, and

Central Africa. The Niassa region in particular has been argued to represent a conduit for the movement of people between these regions (Dixey 1927; Clark 1966; Juwayeyi and

Betzler 1995; Betzler and Ring 1995; Bromage et al 1995; Mercader et al 2009a;

3

Thompson 2010). It is also close to hypothesized refugia (Lahr and Foley 1998) that are argued to have remained habitable by MSA humans during megadroughts

(Cohen et al 2007; Scholz et al 2007; Basell 2008).

While work may have progressed more slowly in Mozambique than neighbouring countries, it has not been completely ignored by the archaeological community. Chapter

3 of this thesis provides a brief summary of previous archaeological work that has been conducted in Mozambique and also outlines the work of a current research project based out of the University of Calgary called Projecto Património Arqueológico e Cultural

(PAC; the Cultural and Archaeological Patrimony Project) that began working in

Mozambique‟s northern Niassa province in 2003. My involvement with PAC began as an undergraduate in 2005. Within Niassa, PAC has conducted archaeological survey work and excavated several sites (see Mercader et al 2008; Mercader et al 2009a; Bennett et al 2010; Mercader et al, in prep) along the coast of the province‟s namesake Lake

Niassa (also referred to as Lake Malawi in Malawi and Lake Nyasa in ). In addition to archaeology, PAC‟s research includes extensive palaeo- and contemporary botanical studies aimed at characterizing modern plant communities and using that data to reconstruct the MSA environments in which early Mozambican H. sapiens lived

(Mercader 2009; Mercader et al 2008; Mercader et al 2009b; Mercader et al 2010;

Mercader et al 2011; Mercader et al, in prep). We have also made a conscious effort to promote capacity building with Mozambican archaeology students and have completed a number of development related projects including the construction of a school, water pump, and museum in the communities in which we have worked (Mercader 2010).

4

This research, including the contribution made by this lithic study, provides a framework upon which theoretical discussions on cultural development and its relationship to biological change can be built and evaluated (Foley and Lahr 1997). With this ultimate goal in mind, this thesis aims to develop a comprehensive understanding of the lithic technology employed by the MSA toolmakers at Mvumu. It creates both a techno-typological classification scheme that accurately accounts for and describes the variation seen in the study assemblage and a comprehensive quantitative dataset that characterizes the assemblage from a metric point of view. To help facilitate the task of intersite comparison, the classification scheme and methods used in this study closely follow existing techno-typologies whenever possible. Some idiosyncrasies of the

Mvumu study assemblage did, however, require new techno-typological categories to be created. The methods used to classify and record the metric attributes of these artifacts are outlined in chapter 4 and followed by the results of the analysis in chapter 5.

These results expand on previously published data on MSA lithic technology from Niassa (see Mercader et al 2008; Mercader et al 2009a) and provide a base for comparisons with lithics from neighbouring sites. Such comparisons suggest that

Mvumu‟s lithic technology fits well with many other assemblages in the Lake Niassa basin and throughout the south-central African woodlands and forests. New chronological data from Mvumu also show that the MSA in Niassa continues later than previously known. Only once variability in MSA behaviour is known on the regional and even basin scale sought by this thesis will it be possible to build strong theories on the nature of early modern human culture (Shea 2011).

5

Chapter Two: The African Middle Stone Age and the Origin of Modern Humans

In the early days of MSA research it was believed that it was coeval with the

European Upper Palaeolithic (UP) (McBrearty and Brooks 2000), but advancements in dating techniques have since improved MSA chronologies and shown that it predates the

UP. The beginning and end dates for the MSA have been amended many times as new discoveries are made, but clear evidence for MSA artifacts is present by ~300kya

(McBrearty and Brooks 2000) and it continues in some areas as late as ~25-26kya

(Opperman1996; Barham 2000). This places its chronology roughly in line with the

Middle Palaeolithic era of European prehistory and means it overlaps with African Late

Stone Age (LSA) sites (e.g. Enkapune Ya Muto, at ~46kya – Ambrose 1998).

Despite this temporal overlap it remains a distinct archaeological entity from both these periods. The original definition of the MSA comes from Goodwin (1929) and was based upon its lithic technology. Principle in Goodwin‟s (1929) definition of the MSA was a shift away from the core tool technologies of the Acheulean (e.g. handaxes and cleavers) toward flake base industries characterized by triangular flakes with faceted platforms.

Prepared core technologies were also included in Goodwin‟s outline of the MSA, but he saw no blades or microliths in the period; rather they were seen as markers of the onset of the (LSA).

Since Goodwin‟s time, our understanding of the MSA has expanded greatly and it has come to represent a much more complete picture of human prehistory in Africa.

Among the key features that form our current definition of this period are the origin of

Homo sapiens and the emergence of a suite of behaviours that are commonly referred to

6

as modern human behaviour (McBrearty and Brooks 2000). As new discoveries continue to be made, and better chronologies are produced, archaeologists have been forced to reconsider the nature and timing of these significant developments. The technological criteria that define the MSA have also been refined over the years since Goodwin coined the term. This chapter briefly reviews the current palaeoanthropological and archaeological evidence for the MSA and uses these data to characterize the period as the majority of archaeologists understand it today. The discussion will begin with the most prominent theories regarding the origin of our species followed by a review of the major models for the increased behavioural complexity, often referred to as modern human behaviour, seen in the MSA. And lastly, it will cover our current understanding of MSA technology and some of its behavioural implications to provide context for the analysis presented in this thesis.

2.1 – Theories regarding the origin of Homo sapiens

The nature of the evolutionary origin of our species, Homo sapiens, has been a major research question in palaeoanthropology since human ancestor fossils began to be reported in the early 20th century. Four main models have been proposed to explain the origin of H. sapiens (Aiello 1993:73-74 as summarized by Stringer 2001:68):

1. The African Replacement Model – argues that modern humans first

arose in Africa ... and spread from there throughout the

world...Indigenous pre-modern populations in other areas of the

7

world were replaced by the migrating populations with little, if any,

hybridization between the groups.

2. The African Replacement with Hybridization Model – similar to

the one mentioned earlier, but allows for a greater or lesser extent

of hybridization between the migrating population and the

indigenous pre-modern populations.

3. The Assimilation Model – also accepts an African origin for

modern humans. However, it differs from the previous models in

denying replacement, or population migration, as a major factor in

the appearance of modern humans...Rather, this model emphasizes

the importance of gene flow, admixture, changing selection

pressures, and resulting directional morphological change.

4. The Multiregional Evolution Model – differs from the previous

three in denying a recent African origin for modern humans...It

emphasizes the role of both genetic continuity over time and gene

flow between contemporaneous populations in arguing that modern

humans arose not only in Africa but also in Europe and Asia from

their Middle Pleistocene forebears.

Models three and four can be subsumed under the more general label of “Multiregional” theories and one and two are considered “Out of Africa” theories (Stringer 2001). The

8

theories presented by these models are distinguished from one another based on their view of the geography, chronology, and the process by which H. sapiens appear (Aiello

1993).

2.1.1 – Multiregional theories

Multiregional continuity models grew out of comparisons between African and

Asian hominins conducted by Franz Weidenreich (1937) in the 1930s. In general, they posit that H. sapiens evolved simultaneously throughout the Old World from pre-modern hominin populations that occupied it during the Out of Africa I diaspora. Supporters of multiregional theories (e.g. Wolpoff 1988, 1989, 1996; Thorne and Wolpoff 1992;

Wolpoff and Relenthford 1997; Relenthford and Jorde 1999; Relenthford 1999; Wolpoff et al 2000) maintain that sufficient gene flow occurred between pre-sapiens hominin groups in all occupied regions that they remained a single species throughout this gradual process; non-African groups were not replaced by H. sapiens from Africa in an Out of

Africa II migration. The primary evidence cited in support of this theory comes from metric and non-metric anatomical traits that multiregional advocates claim show diachronic continuity between Homo ergaster/erectus populations dating to the Out of

Africa I migrations and modern H. sapiens in East Asia, H. neanderthalensis and modern

H. sapiens in Europe, and H. heidelbergensis and H. sapiens in Africa (Wolpoff 2000).

Prior to 1995 multiregional continuity theory often took an approach where no geographic region made a greater contribution to the evolution of H. sapiens (e.g. Thorne and Wolpoff 1992), but recently some advocates of multiregional continuity (e.g.

9

Wolpoff and Relenthford 1997; Relenthford and Jorde 1999; Relenthford 1999) have begun to argue that Africa may have played a larger role in the evolutionary process by virtue of larger population size (Stringer 2001). These more recent variations of multiregional continuity theory fit closely with the position put forth under assimilation models (Stringer 2001).

2.1.2 – Out of Africa theories

Out of Africa theories take the perspective that H. sapiens originated in Africa then moved outward in one or more migrations during the early stages of the late

Pleistocene effectively replacing the existing archaic populations that occupied the Old

World during the Out of Africa I migrations (e.g. Howells 1976; Stringer and Andrews

1988a; Stringer and Andrews 1988b; Lahr and Foley 1994, 1998; Watson et al 1997;

Stringer 2001, 2003; Macaulay et al 2005; Reed and Tishkoff 2006; Campbell and

Tishkoff 2010). This migration is commonly referred to as Out of Africa II. As with multiregional theorists, Out of Africa advocates cite a range of morphological and metric criteria in their analyses, but instead of showing continuity, they see differential markers in the data that show a clear separation between early African H. sapiens and the archaic species found throughout the Old World. Traits considered in these analyses tend to be focused on cranial features including greater rounding of features and general gracilization (see Pearson 2008 for a review of features and methods). Chronologies associated with the current fossil evidence suggest that H. sapiens have greater antiquity in Africa than elsewhere in the Old World. Therefore, the fossil record provides strong

10

support for Out of Africa theories. Increasingly, genetic data are also being cited to support Out of Africa theories. The following two sections of this chapter will briefly review the fossil and genetic data regarding the origin of our species.

2.1.3 – Fossil evidence for the origins of H. sapiens

Currently, studies performed on dated H.sapiens fossils support the Out of Africa theory‟s postulation that the earliest H. sapiens are found in Africa. The earliest fossil evidence comes from northern East Africa. Specimens from Omo, Ethiopia are the oldest yet found and date to 195±5kya (Day 1969; Day et al 1991; McDougall et al 2005;

McDougall et al 2008; Pearson et al 2008). Other early examples of our species have been found at other northern East African sites including Herto, Ethiopia which contained fossils dating to 160-154kya (Clark et al 2003; White et al 2003), and Soleb, Sudan at

160-90kya (McBrearty and Brooks 2000). The earliest modern human fossils in

Southern, southern East, and North Africa tend to be younger than those in northern East

Africa. Southern African fossil remains from Border have been dated at ~90kya

(Grün and Stringer 1991), 80-60kya at Die Kelders (Grine 2000), 127->40kya at Sea

Harvest (Klein 1999), and as early as 118kya at Klasies River (Rightmire and Deacon

1991; Bräuer 2001b). Early examples of H. sapiens from southern East Africa and North

Africa are rarer than those from the northeast and south, but they include specimens found at Mumba , Tanzania dating to 132-88kya (Mehlman 1987; Bräuer and Mehlman 1988; Domínguez-Rodrigo et al 2008), Taramsa, at 80-50kya

11

(Vermeersch et al 1998), and , dating to 190-90kya (Grün and

Stringer 1991). Pleistocene H. sapiens fossils from Central Africa are unknown.

The earliest H. sapiens fossils known from outside of Africa come from the

Levant. Excavations at Jebel Qafzeh, have produced the earliest non-African H. sapiens fossils dating to ~92kya (Valladas et al 1988; Bar-Yosef 1992). Early East Asian

H. sapiens fossils have been found at several sites including Niah Cave in Sarawak,

Malaysian Borneo dating to 42kya (Harrison 1975; Barker et al 2002; Barker et al 2007),

Tabon Cave, Palawan dating to 47kya (Fox 1970; Dizon et al 2002; Détroit et al 2004) and , Phillipines where a metatarsal dated to 68kya has been tentatively classified as H. sapiens and become the oldest Asian evidence for modern human occupation (Mijares et al 2010). Petraglia et al (2007) have also speculated that H. sapiens may have been responsible for the lithic assemblage at Jwalapuram in the Jurreru

River valley of southern , dated to ~75kya, but here the species identification is only inferred based upon the technological nature of the industry and no fossils have been reported. H. sapiens do not appear in Europe until ~40kya at Peştera cu Oase,

(Rougier et al 2007). As such, current fossil evidence shows that H. sapiens emerged and evolved in Africa ~100,000 years before they entered other regions. This is consistent with the view put forth by Out of Africa theorists that modern human origins were in

Africa.

12

2.1.4 – Genetic evidence for the origins of H. sapiens

Genetic data also provide support for an African origin for H. sapiens.

Mitochondrial DNA (mtDNA) research first published by Cann et al (1987) utilized known mtDNA mutation rates to suggest that all humans living today share a common ancestor, often referred to as Mitochondrial Eve, between 280-140kya. Vigilant et al

(1991) later narrowed the age of this common ancestor to 249-166kya, and in 2000

Ingman et al placed the most recent common ancestor of H. sapiens at 171.5±50kya.

Recently, this date has been further refined to 195±32.5kya (Gonder et al 2007; Campbell and Tishkoff 2008); a date that is in remarkably close agreement with the dates for the earliest known H. sapiens fossil material. Genetic data also support the fossil evidence in terms of the geographical origin of H. sapiens within Africa; palaeogenetic research published by Tishkoff et al (2009) suggests an East African origin for H. sapiens. Their argument is based on their observation that the highest degree of mtDNA diversity in a modern population, which extends across all mtDNA haplogroups, is found in Tanzania.

Dates derived from genetic data can also be used to estimate the timing of Out of

Africa migrations, but the dates cited by different research groups remain variable. Many researchers pursuing this line of inquiry see the genetic data closely following the fossil evidence; H. sapiens moved out of Africa and entered the Levant approximately 100kya then moved into Eurasia between 80-40kya (Watson et al 1997; Macaulay et al 2005;

Reed and Tishkoff 2006; Campbell and Tishkoff 2010). Liu et al (2006) prefer a younger date of ~56kya for the earliest migration of H. sapiens that are ancestral to modern populations. They acknowledge the presence of earlier H. sapiens fossils in the Near

13

East dating to 80-100kya but note that such finds are “isolated and may represent an early offshoot that died out” (Liu et al 2006:235) and did not belong to the same migratory population that first managed a sustained presence outside of Africa and expanded beyond the Near East. Recent fossil evidence from East Asia that predates Liu et al‟s

(2006) estimate for the successful colonization of the Old World outside of Africa by H. sapiens, such as that from Callao Cave, (Mijares et al 2010; see above), may serve as a starting point to challenge the younger date, but this evidence is currently based on a single specimen that has been given a tentative assignment to H. sapiens. A larger sample of fossil materials and/or genetic testing of the Callao Cave remains are necessary before any strong conclusions can be drawn. Despite differences regarding the chronology of the earliest successful migrants out of Africa, proponents of both theories agree that this founder population for non-African H. sapiens was small, consisting of

~1000-1500 effective individuals (Liu et al 2006; Campbell and Tishkoff 2010).

In addition to determining the geography and chronology of the origin of H. sapiens, genetic data can be useful in unravelling the process by which our species evolved. Out of Africa theories place less emphasis on gene flow between archaic and modern humans than multiregional theories, but it does play a role in some variations. In the African replacement model gene flow is essentially a non-factor. However, the

African replacement with hybridization model does allow for some level of gene flow between archaic and modern humans, but even if it did occur some palaeoanthropologists question the true contribution it would have made to the modern human genome by suggesting that “while Neanderthal and archaic Asians were…probably not completely

14

uninvolved in the composition of early modern human gene pools, most likely their contribution was generally so small that replacement may be seen as the decisive process” (Bräuer 1992). Since Bräuer made this comment new work on the Neanderthal genome has suggested the Neanderthals did make a lasting, albeit small (1-4%), contribution to the genomes of some modern non-African H. sapiens populations (Green et al 2010). African populations received no genetic input from Neanderthals (Green et al 2010). These findings challenge the pure African replacement model, but do support the idea put forth by the African replacement with hybridization and, to some extent, assimilation models because they show that although some gene flow occurred the majority of the H. sapiens genome is African in origin.

2.2 Archaeological evidence for complexity in Middle Stone Age behaviour

Much of the current research into MSA behaviour is dominated by a single overarching research question: what is the nature and chronology of the emergence of modern human behaviour? Modern behaviour is most often defined in terms of a number of traits that show an increased capacity for complex and abstract thinking, technological and economic innovation, long term planning, and symbol use compared to earlier humans (McBrearty and Brooks 2000; Henshilwood and Marean 2003). A selection of these traits appears in Table 2.1. Although most archaeologists agree that these behavioural traits are indicators of complexity, there remains disagreement on the timing of their emergence and which ones, if any, take precedence as true indicators of modernity. Two theories, the Human Revolution and Long Chronology/Gradualist

15

Table 2.1 – Behavioural traits indicative of Modern Human Behaviour (adapted from McBrearty and Brooks 2000) ______

Ecology

 Range expansion into previously unoccupied regions and ecosystems  Increased diet breadth

Technology

 New lithic technologies, e.g. blades, microliths, backed tools  Standardization within formal tool categories  Hafting and composite tools  Tools in novel materials, e.g. bone, antler  Special purpose tools, e.g. projectiles, geometrics  Increased number of formal tool categories  Geographic variation within formal tool categories  Temporal variation within formal tool categories  Greater control of fire

Economy and social organization

 Long distance procurement and exchange of raw materials  Curation of exotic raw materials  Specialized hunting of large, dangerous species  Scheduling and seasonality in resource exploitation  Site reoccupation  Intensification of resource extraction including aquatic and plant resources  Long-distance exchange networks  Group and individual self-identification through artifact style  Structured use of domestic space

Symbolic behaviour

 Regional artifact styles  Self adornment, e.g. beads, ornaments  Pigment use  Notched and incised objects, e.g. bone, egg shell, , stone  Image and representation  Burials with grave goods, ochre, ritual objects ______

16

theories, consider the complete list of traits in their model for the origins of modern behaviour. Two alternative theories, the Symbolically Mediated Behaviour and Mode 3 theories, place emphasis on their namesake traits as the key signs of behavioural complexity. Subtle variations exist in these theories depending on the specific author writing on the subject, but their basic premises are outlined below:

1. Human Revolution Theory – This theory proposes that the traits associated

with behavioural modernity emerged in a punctuated event referred to as

the “human revolution” near the MSA-LSA boundary ~50-40kya

(Ambrose 1998; Klein 1995, 2000, 2001, 2009). Proponents of this theory

maintain that this behavioural shift may have been caused by either a

neurological change in H. sapiens resulting from a genetic mutation not

visible in fossil morphology or yet detected in palaeo-DNA studies (Klein

1995, 1999, 2000, 2001) or a cultural transition that led to H. sapiens

tapping into existing but previously unused capacity for behavioural

complexity (Stringer and Gamble 1993; Mellars 1996; Mithen 1996; Bar-

Yosef 1998, 2002; Conard and Bolus 2003). In light of the biological

evidence presented above, this theory suggests that the emergence of

modern behaviour did not coincide with the appearance of anatomically

modern humans in the fossil record (Mellars 2005; Stringer 2007).

Instead, this paradigm implies that the behavioural patterns of pre-LSA

African H. sapiens were characterized by the utilization of simple material

17

culture with simple lithics, no non-lithic tools, basic subsistence that did

not include the hunting of large dangerous animals, or exploitation of

marine resources (e.g. fishing, sealing, etc), flying birds, or any other

seasonal resources, and no symbolic behaviours (Klein 1995, 2001; 2009;

Henshilwood and Marean 2003).

2. Long Chronology/Gradualist Theory – This position was summarized by

McBrearty and Brooks in 2000. They cite many of the same traits used

by Human Revolution theorists, but argue that the markers of behavioural

modernity predate the LSA and appear not as a complete suite of

behaviours in a punctuated event, but rather in a gradual, step-wise

manner. This process is seen as beginning in the early MSA and being

fully developed by the MSA-LSA transition (McBrearty and Brooks

2000; Figure 1). Behavioural modernity is not seen as being connected to

any biological change (McBrearty and Brooks 2000). Recent evidence

from Kenya suggests that the origins of some of these traits, such as blade

technology, may have emerged as long as 500kya (Johnson and

McBrearty 2010); thus predating conventional dates for both the MSA

and the origin of H. sapiens which suggests that modern behaviour may

have roots in the ESA amongst non-modern humans. The complete suite

of complex behaviours are not seen as stemming from any single

geographic region in Africa, rather each one emerged independently,

18

perhaps multiple times, in diverse areas of the continent during the MSA

on an “as needed” basis (McBrearty and Brooks 2000). Powell et al

(2009) have suggested that the demographic conditions necessary for

these traits to coalesce into a modern suite of complex behaviours would

have been in place by ~101kya.

3. Symbolically Mediated Behaviour Theory – Supporters of this position

define modern human behaviour as “behaviour that is mediated by

socially constructed patterns of symbolic thinking, actions, and

communication that allow for material and information exchange and

cultural continuity between and across generations and contemporaneous

communities” (Henshilwood and Marean 2003:635). Therefore, they

reject many of the traits cited by proponents of the above theories in

favour of viewing symbolic behaviour as the sole defining feature of

behavioural modernity. To justify its focus on symbolic behaviour alone

as the trait indicative of modern behaviour, this model holds that many of

the behavioural traits cited by the Human Revolution and Long

Chronology/Gradualist theories are too easily explained by resources and

labour intensification behaviours and taphonomic biases (Henshilwood

and Marean 2003). Other arguments against the trait list approaches are

that the traits are derived from the European Palaeolithic record (Deacon

19

Figure 2.1 – Step-Wise Emergence of Modern Behaviours (source McBrearty and Brooks 2000:530).

1995; Henshilwood and Marean 2003) and therefore are not

representative of complex behaviour in Africa or that they lack sufficient

theoretical grounding to make them usable as indicators of behavioural

modernity (Henshilwood and Marean 2003). Symbolically Mediated

Behaviour theory is most popular amongst researchers working in South

20

Africa (e.g. Wurz 1999; Henshilwood et al 2001a; Wadley 2001;

Henshilwood et al 2002; Watts 2002; Henshilwood et al 2004; d‟Errico et

al 2005; Soriano et al 2007; d‟Errico et al 2008; MacKay and Welz 2008)

where the majority of early evidence for symbolic behaviour has been

found (cf. Barham 1998, 2000, 2002a, 2002b; Van Peer 2003 for early

symbolic evidence in and Sudan respectively).

4. Mode 3 Theory – Like the Symbolically Mediated Behaviour theory, the

Mode 3 theory focuses on a single aspect of the archaeological record to

identify behavioural modernity. Put forth by Foley and Lahr (1997,

2003), this theory holds that the emergence of Mode 3 (prepared core)

technologies, as defined by Clark (1977), in prehistoric technological

repertoires marks the origin of modern behaviour. In this theory, Mode

3 technology is used as a proxy for greater cognitive capabilities relative

to earlier hominins that allowed for more complex, and therefore

modern, behaviour. The choice of Mode 3 technology as the criterion

for behavioural modernity has two key implications: 1) modern

behaviour patterns predate the earliest H. sapiens and 2) it implies that

modern behaviour patterns were not limited to H. sapiens. The

Mousterian Industry in Europe and the Levant (emerging ~330kya,

Skinner et al 2007) is associated with Homo neanderthalensis and has

long been known to contain Mode 3 technology. Recently, the origins

21

of Levallois Mode 3 technology have been pushed back to ~700kya

around the time of the transition from Homo ergaster to Homo

heidelbergensis (Bernal and Brooks 2010). As such, modern behaviour

would also be attributed to Neanderthals and H. heidelbergensis under

this model. Moreover, discoidal prepared core technology has been

reported in the Acheulean (Kuman 2001) and even in Oldowan

assemblages (Kuman 1998) opening the possibility the Homo ergaster,

Homo habilis, and possibly even some Australopithecines practiced

some elements of modern behaviour.

Prior to the late 1990s, most archaeologists subscribed to some form of the Human

Revolution hypothesis (Henshilwood and Marean 2003), but since that time continued work in Africa and improved chronologies for the MSA have increasingly showed that essentially all of the commonly cited markers of behavioural modernity (Table 2.1) originated in Africa at dates preceding the LSA. Subsequently, the human revolution theory has been abandoned by most of its one time supporters (cf. Klein 2009).

Likewise, the increasing evidence for the great antiquity of Mode 3 technology and its association with species other than H. sapiens (Kuman 2001; Bernal and Brooks 2010) has proved problematic for the widespread acceptance of the Mode 3 hypothesis. As such, most African Stone Age archaeologists currently subscribe to variations of the

Long Chronology/Gradualist and Symbolically Mediated Behaviour theories.

22

2.3 – Middle Stone Age technological behaviour

As evidenced by the list cited above (Table 2.1), MSA behaviour is characterized by a large number of traits. Many of these traits go beyond the scope of this thesis, but one, MSA technological behaviour, is directly relevant to the analysis presented here and deserves further discussion. Technology is one of the most studied aspects of MSA behaviour. Its popularity as a research topic may be partially explained by the fact that technology, especially lithic technology, is one of the most commonly preserved by- products of human behaviour in the archaeological record, but archaeologists‟ obsession with ancient technology runs deeper than convenience. Technology is intimately linked with cultural and biological evolution (Ambrose 2001; Foley and Lahr 1997, 2003).

Technological studies not only inform us about ancient technical behaviours such as core reduction or tool retouching, but can also provide insight into other aspects of behaviour such as economics (e.g. raw material procurement, subsistence strategies, exchange networks), ecology and land use (e.g. expansion into new eco-geographical regions, territorial range expansions), and symbol use (e.g. regional lithic style). The remainder of this chapter will discuss some key aspects of MSA technology as we understand them today and their implications for other elements of MSA behaviour to provide the necessary context for evaluating the Mvumu study assemblage.

In addition to providing a wealth of behavioural information, technology is often used by archaeologists to organize human prehistory. This can be seen both in terms of broad sweeping categories such as Christian Thomsen‟s Stone, Bronze, and Iron Ages (in

Trigger 2006:123) and through subtler differences within these broad categories. Such is

23

the case in African Stone age archaeology with the ESA, MSA, LSA division.

Goodwin‟s (1929) original definition of the MSA was based upon changes in lithic technology he observed in South Africa that represented a significant shift away from the

Acheulean industry as it was understood at the time. The primary technological markers used in this definition were the transition to flake-based industries utilizing flakes with faceted striking platforms and a tendency toward convergent rather than parallel lateral edges combined with the lack of ESA style core tools and LSA microlithic blade-based industries (Goodwin 1929:97-99).

Since the 1920‟s, continued research has shown MSA lithic assemblages to be more diverse than once thought. For example, large ESA-style core tools do persist in many MSA assemblages. While this pattern is seen throughout the MSA both chronologically and geographically (including in northern Mozambique – see Mercader et al 2008; Mercader et al 2009a), it is perhaps best seen in assemblages associated with the Sangoan industry. Originally defined at Sango Bay, (Wayland and Smith

1923), the Sangoan industry is characterized by the presence of large heavy-duty tools such as core-axes, picks, and core scrapers along with a smaller light duty flake tool component (Wayland and Smith 1923; McBrearty 1988, 1991; Clark 2001a; McBrearty and Tryon 2006; Rots and Van Peer 2006). Sangoan sites are typically found in East

Africa (Wayland 1934, 1935; Cole 1967; McBrearty 1988, 1991, 1992; Van Peer 2003;

Rots and Van Peer 2006), Central Africa (Lanfranchi 1990, 1996), and West Africa (Omi

1977; Nygaard and Talbot 1984; Lioubine and Guede 2000; Mercader and Martí 2003), and tropical Southern Africa (Armstrong 1931; Bond 1948; Cooke 1962, 1963; Clark

24

1962, 1964a, 1965, 1969, 1974, 1982, 2001a). Examples south of the Limpopo River in

South Africa are rare (cf. Kuman et al 2005; Wilkins 2008). This industry is often thought of as a woodland adaptation (Clark 1964a, 1964b, 1965, 1970, 1972, 1975,

1982), but several Sangoan sites have also been discovered in savannah-like environments (MacCalman and Viereck 1967; Sampson 1974; Kalb et al 1982;

McBrearty 1988, 1991; Van Peer 2003; Rots and Van Peer 2006) making it difficult to confidently associate it with any single ecosystem. Although it is widely distributed throughout much of tropical sub-Saharan Africa, the Sangoan‟s absence from the archaeological records of some parts of the continent has been argued by some to mark the earliest expression of regionality in lithic technology (Clark 1970). The chronology of the Sangoan industrial complex is not well understood, but it is often found stratified between final Acheulean and early MSA assemblages and its antiquity could exceed

270kya (Barham 2000). Its age, frequent interstratification between ESA and MSA materials, and mix of archaic and derived technological features has raised questions amongst some archaeologists regarding its inclusion within the MSA proper (McBrearty

1991; Clark 1982, 2001a), but others see no conflict (Sheppard and Kleindienst 1996).

Regardless whether the Sangoan is considered to be MSA proper, its common stratigraphic position above Acheulean levels illustrates that core tool technology carried on into the post-Acheulean era (McBrearty 1988).

Although it is now clear that core tool traditions continued into the MSA they are not present at all sites and where they are found they typically represent a much smaller

25

proportion of tools relative to ESA assemblages. Instead, MSA tool assemblages are primarily composed of smaller flake-based tools as described by Goodwin (1929). The production of flake blanks was accomplished through a number of core reduction strategies ranging from simple amorphous cores to more complex prepared core technologies consistent with Clark‟s (1977) Mode 3 such as Levallois and discoidal.

These prepared core reduction strategies require greater depth of planning and conceptualization of form than is commonly seen in pre-MSA assemblages (Foley and

Lahr 1997, 2003). It has recently been suggested that the roots of prepared core technology may extend back to 700kya thus predating the MSA by as much as 400ky

(Bernal and Brooks 2010), but even if this is the case it is not until the MSA that Mode 3 technology becomes a common and widespread part of human technological repertoires.

Largely due to this increased frequency and distribution of Mode 3 technologies, and the behavioural and cognitive connotations of its greater complexity compared to simple reduction, research into MSA core reduction strategies often focuses on prepared cores as opposed to their simple counterparts (e.g. Van Peer 1991, 1992; Pleurdeau 2005; Tryon et al 2005; Tryon 2006; Wilkins 2008; Wilkins et al 2010).

MSA prepared core technologies are widely distributed throughout Africa.

Levallois technology has been recorded across North Africa from Egypt to Morocco

(Wendorf et al 1987; Van Peer 1991, 1992, 1998, 2003; Hawkins and Kleindienst 2002;

Garcea 2004; Rose 2004; Nespoulet et al 2008), in East Africa including several sites in

Ethiopia (Pleurdeau 2005; Yellen et al 2005; Shea 2008), Kenya (McBrearty et al 1996;

26

Tryon et al 2005; Tryon 2006; Tryon et al 2008), and Tanzania (Willoughby and Sipe

2002), in West Africa (Robert et al 2003), in Central Africa (Cornelissen 2002; Mercader and Martí 2003; Williams 2005), and in Southern Africa including Angola (Clark 1963),

Mozambique (Mercader et al 2008; Mercader et al 2009a), Zambia (Barham 2000, 2002a;

Clark 2001b; Clark and Brown 2001), and South Africa (Singer and Wymer 1982; Wurz

2002; Kuman et al 2005; Villa et al 2005; Marean et al 2007; Soriano et al 2007; Wilkins

2008; Villa et al 2010; Wilkins et al 2010). Discoidal approaches are equally widely distributed throughout North Africa (Van Peer 2003; Rose 2004), East Africa (Yellen

2005; Tryon 2006; Shea 2008; Tryon et al 2008; Willoughby and Sipe 2002), West

Africa (Boriskovsky and Soloview 1978; Nygaard and Talbot 1984; Robert et al 2003),

Central Africa (Bayle des Hermens 1975; Cornelissen 2002; Mercader and Martí 2003;

Williams 2005), and Southern Africa (van Riet Lowe 1945; Clark 1963; Clark and

Haynes 1970; Singer and Wymer 1982; Barham 2000; Clark 2001b; Clark and Brown

2001; Barham 2002a; Villa et al 2005; Mercader et al 2008; Wilkins 2008; Mercader et al

2009a).

Goodwin‟s (1929) original definition of the MSA excluded blade technology and microlithic/geometric industries, Clark‟s (1977) Mode 4 and Mode 5, but this notion has changed since his time and they are now known in some MSA assemblages. Blade technology employs specially planned cores designed to allow for the production of a series of morphologically standardized blanks with parallel or nearly parallel lateral edges and a length to width ratio of >2:1 (Bordes 1961; Inizan et al 1992; Inizan et al

27

1999; Kooyman 2000; Andrefsky 2005). Such blanks are referred to as blades or bladelets depending on their size. The identification of formal blades can be complicated in some assemblages because even the simplest reduction strategies will inevitably produce some parallel sided elongated flakes. Therefore, the key to identifying early examples of blade technology is not found in the blades themselves but rather in the presence of systematically produced blade cores. MSA blades were produced by a number of techniques including from cylindrical and pyramidal blade cores and specialized Levallois cores (McBrearty and Brooks 2000; Tryon et al 2005; Shea 2008;

Johnson and McBrearty 2010).

Although not present in all MSA assemblages, blade technology can be found in many parts of Africa including parts of North Africa such as (Chazan 1995), the

Western Desert and Nile Valley in Egypt (Van Peer 1998; Hawkins and Kleindienst

2002), and northern Sudan (Rose 2004), East Africa at sites in Ethiopia (Wendorf and

Schild 1974; Clark et al 1984; Pleurdeau 2005; Yellen et al 2005; Shea 2008) and Kenya

(Tryon and McBrearty 2002, 2006; Tryon et al 2005; Tryon et al 2008), Central Africa in

D.R. Congo (Williams 2005), and in southern Africa in Zambia (Barham 2000, 2002;

Clark 2001; Clark and Brown 2001) and South Africa (Singer and Wymer 1982;

Thackeray 1989, 1992; Kuman et al 1999; Wurz 1999, 2002; Henshilwood et al 2001b;

Villa et al 2005; Marean et al 2007; Soriano et al 2007). Like Mode 3 technology, blade technology may predate the MSA. Excavations in Kenya have shown blades to also be associated with Acheulean industries (Leakey et al 1969; Tryon and McBrearty 2002;

28

Tryon et al 2005). In fact, recent work has pushed the origins of blade technology back to 500,000 years ago (Johnson and McBrearty 2010), well before commonly accepted dates for the earliest MSA. But as is the case with Mode 3, examples of Mode 4 technology do not begin to appear in lithic assemblages with any degree of regularity until the MSA.

Mode 5 technology, including macrolithic and microlithic backed blades and geometrics, was once thought to be one of the principal markers of the LSA (Goodwin

1929; McBrearty and Brooks 2000), but is now also known in many MSA assemblages.

Typically made on blade/bladelet fragments, Mode 5 tools are characterized by steep retouch, or “backing”, on one or more sides of the tool while the side opposing this retouch remains sharp (Clark and Kleindienst 2001). In geometrics, this retouch forms the tool into commonly recognized shapes such as triangle, trapezes, and crescents/lunates that are used to further classify the tool (McBrearty and Brooks 2000; Clark and

Kleindienst 2001). Non-geometric tools generally retain the basic shape of the blank and are referred to as backed blades (Clark and Kleindienst 2001). The size threshold beneath which a tool is considered macro- or microlithic varies by author, but is typically between 20mm (Andrefsky 2005) and 30mm (Clark and Kleindienst 2001). Regardless of whether they meet the technical size criterion to be considered microlithic, Mode 5 tools tend to be too small to be held and used comfortably in the hand and they are therefore interpreted as hafted tools (Clark 1977; McBrearty and Brooks 2000; Barham

2002a; Lombard 2005, 2007, 2008; Soriano et al 2008; Lombard and Pargeter 2008;

29

Wadley et al 2009; Lombard and Phillipson 2010; see below for more on hafting technology).

Unlike prepared cores and blades, microliths have not yet been found in pre-MSA contexts, but instead appear midway through the period typically thought of as the MSA.

Amongst the best known MSA microlithic industries is a South African regional subphase known as . The Howiesons Poort industry is characterized by backed geometrics on blades or blade fragments the are similar, but generally larger than

LSA microliths with average lengths of between 30-50mm (Goodwin 1929; Wurz 1999;

Thackeray 2000; Soriano et al 2007). It was originally defined by Goodwin (1929), in the same volume that defined the MSA, but at that time it was thought to post date the

MSA and represent the transition into the LSA (McBrearty and Brooks 2000). Since then, excavations have revealed Howiesons Poort levels interstratified with typical MSA levels at sites such as Klasies River (Singer and Wymer 1982; Wurz 2002) and improvements in dating techniques now place it between approximately 80kya and 55kya

(Miller et al 1999; Lombard 2005; Lombard and Pargeter 2008). In Central Africa Mode

5 technologies predate the Howiesons Poort. Early examples of backed blades from

Zambia that are approaching microlithic size but lack the morphological standardization of the Howiesons Poort are associated with the Lupemban industry at dates as old as

300kya (Barham 2002a). Rare microlithic artifacts have also been reported in northern

Mozambique with dates of up to 105kya (Mercader et al 2008; Mercader et al 2009a).

30

These technological innovations provided a basis for the diversification of tool types seen during the MSA. Humans were no longer relying on large, multipurpose tools such as handaxes. Instead, they were making multiple types of small specialized tools including scrapers, awls, and points (McBrearty and Tryon 2006). Points especially have been considered to be diagnostic of the MSA (Goodwin 1929; McBrearty and Brooks

2000; McBrearty and Tryon 2006). Points are also the most commonly cited evidence of techno-typological diversification in the MSA. Despite forming a single artifact category with presumed functional equivalence throughout Africa, they often show distinct regional stylistic differences (McBrearty and Brooks 2000). Such regional variation can be seen in tanged Aterian points (Clark 1988; Debénath 1994; Garcea 2004; Shea 2009) and Levallois-based Nazley Khater points (Van Peer 1991, 1998) in North Africa, foliate and Levallois points from East Africa (Clark 1988; Tryon et al 2005; Shea 2008), and

Lupemban lanceolate points in Central Africa (Pommeret 1966; Cole 1967; MacCalman and Viereck 1967; Nenquin 1967; Van Moorsel 1970; Van Noten et al 1972; Van Noten

1982; Cahen 1978; McBrearty 1988; Mercader and Martí 1999, 2003; Barham 2000;

Mercader et al 2002). Southern Africa claims the richest diversity of regional point styles including bifacial Still Bay points (Goodwin 1928; Volman 1984; Henshilwood and

Sealy 1997; Henshilwood et al 2001b; Lombard 2006; Villa et al 2008), foliate and triangular corner-struck Bambata points (Armstrong 1931; Cooke 1966; McBrearty and

Brooks 2000; Brooks et al 2006), Howiesons Poort points including unifacial and bifacial retouched foliate points and possible composite points composed of backed geometrics

31

(Singer and Wymer 1982; Wurz 1999; Lombard 2008; Lombard and Pargeter 2008; Shea

2009; Lombard and Phillipson 2010).

It has been argued by some researchers that the regional differentiation seen in points, as well as industrial subphases such as the Howiesons Poort, represent symbolic behaviour through the intentional impartation of style (Sackett 1977, 1982, 1990; Conkey

1990; Wurz 1999; Wadley 2001; Henshilwood and Marean 2003; Soriano 2007; cf.

Chase and Dibble 1987). It is conceivable that some of the variation seen in lithic style can be explained by changes in technological approaches over time, but as chronologies for many MSA sites have improved it has become clear that several contemporaneous geographically distinct tool traditions are present. This phenomenon may have its roots in the early MSA Sangoan industry (Clark 1970). While the Sangoan does have a fairly broad distribution compared to later MSA industries like the Howiesons Poort, it is not found in many parts of the continent and therefore represents a clear departure from ancestral tool industries such as the Acheulean that were essentially homogenous throughout Africa. Proponents of a symbolic basis for this regionality argue that it shows evidence for greater cultural complexity that includes concepts of social identity and ethnicity (Sackett 1990; Wynn 1996; Wurz 1999; Wadley 2001; Soriano et al 2007).

MSA technology also suggests greater complexity in subsistence behaviour.

Points are the most frequently cited evidence supporting this. It is possible MSA points may have served as cutting and scraping tools (Kuman 1989 as cited in McBrearty and

Brooks 2000; Wendorf and Schild 1993), but many archaeologists also argue that they

32

were used to tip weapons used in hunting activities (Milo 1998; McBrearty and Brooks

2000; Lombard 2005; Shea 2006). Strong evidence for points as hunting tools can be seen at sites such as ≠Gi, where over 600 points, accounting for ~41% of the

MSA lithic assemblage, were found in association with fauna remains from large potentially dangerous game (McBrearty and Brooks 2000). The manner in which points were used in hunting activities is a matter of debate amongst some researchers, but the small size and retouch aimed at creating a symmetrical mass distribution and an aerodynamic shape may point to the emergence of projectile technologies during the

MSA in the form of spearthrower darts or perhaps even the bow and arrow (Gresham and

Brandt 1996; McBrearty and Brooks 2000; Shea 2009; Lombard and Phillipson 2010).

Studies conducted on points from ≠Gi in Botswana and South African sites including

Blombos Cave (Kuman 1989 cited in McBrearty and Brooks 2000) and

(Lombard 2005) have identified impact damage on the tips of points consistent with their use as projectiles. The development of these types of projectile technologies would have allowed hunters to kill potentially dangerous game from a safe distance and may well have played a role in broadening subsistence and ecological patterns of MSA people

(Shea 2009; Steele and Klein 2009).

As suggested by their presumed use on spears/darts/arrows, both points and the microlithic industries discussed above are often associated with another technological innovation that is connected to the MSA: hafting (Singer and Wymer 1982; Wurz 1999;

Barham 2002a; Lombard 2005, 2007; Wadley 2005; Villa and Lenoir 2006; Soriano et al

33

2007; Lombard and Pargeter 2008; Wadley 2009; Lombard and Phillipson 2010).

Hafting is often facilitated by basal thinning or, in the case of Aterian tools, the creation of a tang (McBrearty and Brooks 2000). While the benefit of hafting small tools is clear,

MSA hafting behaviour may not have been limited to flake tools. Rots and Van Peer

(2006) have presented evidence that heavy duty core-axes were also hafted for use in

Sudan. The precise method used to fix artifacts to shafts or handles is unclear, but both archaeological studies and the ethnographic record suggest that a combination of plant- based twine and adhesive resins was likely involved (Wendt 1976; Lombard 2005, 2007;

Wadley 2005; Lombard and Pargeter 2008; Wadley and Mohapi 2008). Ochre may have also played a role in hafting as an adhesive and/or a symbolic part of the practice

(Wadley et al 2004; Lombard 2005, 2006a, 2006, 2007; Wadley and Mohapi 2008).

Another key aspect of MSA technology is tools made on non-lithic raw materials such as bone. The use of bone in tool manufacture is not a MSA innovation per se. Rare bone implements, including pieces flaked in a similar manner to that used in lithic reduction and digging sticks, may date back nearly two million years and the

Australopithecines (Dart 1960; Leakey 1971; Backwell and d‟Errico 2000, 2001, 2008).

Bone tools remain relatively rare in the MSA, but they are more abundant relative to ESA assemblages. This phenomenon could be related to taphonomy rather then behaviour, but regardless of the reason for the proliferation of bone tools during the MSA there are recognizable differences between ESA and MSA bone tool technology. MSA bone tools show a much greater degree of sophistication than older examples and were produced by

34

novel techniques including scraping, grinding, and notching (Yellen et al 1995; d‟Errico and Henshilwood 2007). Many MSA examples seem to be related to projectile technology (McBrearty and Brooks 2000; McBrearty and Tryon 2006). Cylindrical bone points have been recovered at a number of South African sites including Klasies River

(Singer and Wymer 1982), Sibudu Cave (Backwell et al 2008), Peers Cave (d‟Errico and

Henshilwood 2007) and (Henshilwood and Sealy 1997; d‟Errico and

Henshilwood 2007). The bone assemblage at Blombos Cave also includes tanged points, awls, and pins (d‟Errico and Henshilwood 2007). Evidence for bone points in Central

Africa predates these South African tools. Several barbed and unbarbed bone points dating to ~90kya have been recovered from Katanda, D.R. Congo (Brooks et al 1995;

Yellen et al 1995; Yellen 1996, 1998). Barbed points have also been reported at White

Paintings Rock Shelter in Botswana (Robbins et al 1994; Robbins et al 2000) and Omo,

Ethiopia (Trapani 2008). Despite the broad geographic range of the sites containing these barbed points they were all found in association with fish remains including those of large catfish (Clarias) which can weigh upwards of 35kg. As such, they are commonly interpreted as harpoons and cited as evidence the further broadening of subsistence behaviour through the incorporation of aquatic resources (Robbins et al 1994; Brooks et al 1995; Yellen et al 1995; Yellen 1996, 1998; McBrearty and Brooks 2000; Robbins et al 2000; Trapani 2008). Moreover, evidence from Katanda and Omo suggests that fishing was a seasonal activity (McBrearty and Brooks 2000; Trapani 2008) adding further complexity to subsistence behaviour in the form of long term planning and scheduling.

35

The preceding is by no means a comprehensive review of MSA technological innovation, but does serve to highlight many of the key technological features that define our current understanding of the MSA. The most important point to take away from this discussion is that MSA technology represents a broad techno-typological diversification from both a diachronic and geographic point of view that shows increasing behavioural complexity. Not all of the technologies discussed above are MSA innovations per se, but their proliferation throughout all parts of the continent represents a marked departure from the relatively homogeneous technology of the ESA. It is this greater diversity and complexity that identifies MSA technology as a marker of behavioural modernity in the eyes of many archaeologists. Moreover, this technological paradigm would have been a prerequisite for many of the other frequently cited indicators of behavioural modernity including economic and ecological diversification.

Although this chapter made a conscious effort to present a pan-African review of the evidence for modern human origins, the emergence of modern behaviour, and some key points of MSA technology, our knowledge of MSA archaeology is certainly not evenly distributed throughout the continent. There is a clear bias towards East and

Southern Africa in the current research. However, even within these most thoroughly investigated regions there are several large areas that remain archaeologically underexplored. The remainder of this thesis deals with one of these understudied areas in

Southern Africa: Niassa, Mozambique. In truth, all of Mozambique, not just Niassa, has received scant archaeological attention from international research projects compared to

36

many of its neighbours, but this trend is beginning to change. The following chapter will present a brief history of previous work in Mozambique and then outline the details of a recent project based out of the University of Calgary and the Universidade Eduardo

Mondlane in Maputo that has conducted the most detailed investigation of the MSA in

Niassa to date.

37

Chapter Three: Archaeology in Mozambique

Mozambique is located on the east coast of Africa at the southern end of the Great

African Rift system. It is bordered by Tanzania to the north, South Africa and Swaziland to the south, and , Zambia, and Malawi to the west (Figure 3.1). As such, it represents a corridor connecting Southern, Eastern, and Central Africa (Clark 1966;

Meneses 1988, 2004; Mercader et al 2009a). Mozambique‟s north-western province of

Niassa shares a long coastline with Lake Niassa; the southernmost of Africa‟s Great

Lakes. Such lakeside margins are thought to have maintained ecological stability (Basell

2008) throughout times of intense MSA aridity (Scholz et al 2007). It has therefore been suggested that they served as refugia and conduits that guided human migrations within and out of Africa (Cohen et al 2007). These geographic traits mark Mozambique, and

Niassa, as an important location for bettering our understanding of MSA populations.

Archaeological research in Mozambique began in the 1930s during the

Portuguese colonial period (see below), but although the prehistory of many of the surrounding countries has been extensively studied, Mozambique‟s archaeology remains amongst the least explored in Sub-Saharan Africa (Meneses 1988, 2004; Sinclair et al

1993). This is especially true for provinces north of the Zambezi River including

Zambezia, Tete, Cabo Delgado, and Niassa (Morais 1984, 1988). A number of causes may be cited for this lack of archaeological attention, not the least of which is a 16 year period of civil strife following independence from in 1975 and the slow process of post-conflict rebuilding combined with a lack of resources and trained personnel

38

______

______

Figure 3.1 – Study Area Location (Mercader et al, in prep)

39

(Sinclair et al 1993; Mercader et al 2009a). However, recent work has begun to fill this gap (see Rodrigues 2006a, 2006b, 2007; Mercader et al 2008; Mercader et al 2009a).

The remainder of this chapter will give a brief history of this previous work and present the details of a recent project based out of the University of Calgary that has excavated a number of MSA sites in Niassa including Mvumu, the site upon which this thesis is based.

3.1 – History of archaeology in Mozambique

The Portuguese colonial authorities in Mozambique began recording ethnographic information on the indigenous peoples they encountered immediately following their arrival in the territory during the late 15th century (see Morais 1988). However, these early documents were concerned primarily with identifying possible sources of valuable resources such as gold and ivory than they were about recording the intricacies of local cultures and their prehistory (see Morais 1988). It was not until 1721 when the Bishop of Mozambique presented a number of rock paintings to the Academia Real das Ciências de Lisboa that the first report on pre-colonial cultural resources from Mozambique was made (Morais 1984, 1988). Despite this report, and in contrast to developments in neighbouring British colonies, interest in Mozambican antiquities remained minimal

(Morais 1988). With the exception of rare investigations made by foreign scholars (e.g.

Carl Wiese at Chifumbaze Cave in 1907, in Phillipson 1976:17; Wieschoff in Manica province in 1930, Wieschoff 1941) few further reports on the prehistory of the country were made until the mid-1930s (e.g. Correia 1934; Santos Júnior 1941) when the

40

Portuguese administration under Salazar took an interest in Mozambican prehistory as part of their colonial reforms (Morais 1984, 1988). Morais (1984:114) suggests that the creation of the Missão Anthropológica de Moçambique in 1936, under which much of the early 20th century Mozambican anthropology was conducted, was a result of these reforms.

Many of these early studies dealt with the general culture traits of indigenous ethnic groups, but some also focused on aspects of physical anthropology and comparative biology (Morais 1984). By the 1940s interest in Mozambican Stone Age archaeology increased amongst colonial scholars (e.g. Borges 1944; Bettencourt Dias

1948; Barradas 1949, 1962; Simões 1951, 1958; Senna-Martinez 1968a, 1968b; see

Rodrigues 2007 for summary of Santos Júnior‟s work in 1955) and the number of recorded sites began to increase (Morais 1988), but archaeological research continued to be scarce. A number of foreign archaeologists also took a greater interest in the Mozambican Stone age during this period (e.g. van Riet Lowe 1943; Wells 1943;

Breuil 1944; Derricourt 1975; Smolla and Korfmann, n.d.). The focus on Stone Age archaeology remained until 1975, when the total number of Stone Age sites recorded in the national inventory of archaeological sites at the Eduardo Mondlane University was

133 (Madiquida 2010, personal communication; cf. Morais 1984, 1988). Due to the proximity to the capital city of Maputo, a bias toward the southern part of the country was prevalent with only seven (6%) sites recorded in the northern provinces (Morais

1988; Sinclair et al 1993). Even into the 1980s nearly 60% of known sites were located in the southernmost province of Maputo with almost 80% being south of the Save River

41

(Meneses 1988, 1999, 2004). Beginning in the 1960s, articles on Iron Age sites and rock art by authors such as de Oliveira and Rita-Ferreira began to appear in newspapers and magazines, but these reports were limited to the colonial media and reached only a narrow portion of the population (Morais 1984, 1988; Sinclair et al 1993).

Once Mozambique gained independence from Portugal in 1975, archaeological and ethnohistorical research began to shift away from the priorities of the colonial agenda in a perfunctory attempt to serve the needs of post-colonial Mozambican society (Sinclair et al 1993). These efforts grew out of the work of the Archaeological Section of the

Departmento de Ciências de Terra de Instituto de Investigação Científica (Earth Sciences

Department of the Scientific Research Institute of Mozambique) which was formed just prior to independence in 1974 under the direction of Professor G. Soares de Carvalho to conduct archaeological salvage work ahead of the construction of the Massingir Dam.

Key in this refocusing of archaeological research in the post-colonial era was a shift to the study of more recent Iron Age farming cultures (Sinclair et al 1993). In addition to expanding knowledge of the Mozambican Iron Age, work done during this period contributed to a greater understanding of the complexities of languages spoken in the country, the origins of farming, the slave trade, and to the training of Mozambicans

(Medeiros 1997; Serra 2000; Omar and António 2004; Zimba et al 2005; Ngunga 2009;

Madiquida 2010, personal communication; Ngunga et al 2010; Raja 2010, personal communication). However, this work was limited in scope.

Post-independence research began with a large scale survey program which ran from 1976-1984. By 1978, this survey program began receiving financial support from

42

the Swedish Agency for Research Corporation (SIDA-SAREC) and in 1982 the Swedish

Board of Antiquities came onboard with analytical, personal, and logistical support

(Morais 1988). This survey program sought to address the temporal and geographical biases in previous archaeological work and led to a doubling of the number of known archaeological sites in Mozambique (Morais 1988). Important excavations during this time include those at the Zimbabwe culture site Manyikeni (Garlake 1976), the early coastal mercantile site of Chibuene (Sinclair 1982, 1985, 1986), the Early Iron Age site

Matola (Cruz e Silva 1976, 1980), and at the Eduardo Mondlane University campus in

Maputo (Sinclair et al 1987). SIDA-SAREC has also provided funding for other projects including, but not limited to, work by Adam (1993) at Zitundo and Teixeira Duarte at

Somomá, Ilha de Moçambique. However, despite these efforts most of the country remained unexplored from an archaeological perspective with very little work being done outside of Maputo (Madiquida 2010, personal communication; cf. Adamowicz 1987,

1988, 1990).

Along with the commencement of the survey program in 1976, Eduardo

Mondlane University (UEM) began offering courses in African prehistory for the first time under the direction of the Archaeological Section (Morais 1988). By 1980 the

Archaeological Section became the Department of Archaeology and Anthropology

(Sinclair et al 1993). In addition to training Mozambicans in prehistory at the university level, the Department of Archaeology and Anthropology was also involved in training cultural delegates and museum and antiquities personnel abroad beginning with Solange

Macamo in the former Soviet Union in 1987. These educational efforts were also

43

instrumental in the fostering of the current generation of Mozambican archaeologists who remain focused primarily on Iron Age archaeology (e.g. Macamo and Saetersdal, 2004;

Macamo, 2006; Madiquida 2006, 2007) and play key roles in current patrimony protection, archaeological research, education, and historic resource conservation in

Mozambique. A new generation of foreign archaeologists, primarily working out of

Portuguese (e.g. Rodrigues 2006a, 2006b), Swedish (e.g. Ekblom 2004; Kohtamaki, personal communication; Sillén, personal communication), and Norwegian (e.g.

Saetersdal 2003) institutions have also continued making contributions to Iron Age archaeology in Mozambique.

Although Iron Age archaeology has been the focus of most archaeologists working in post-colonial Mozambique, Stone Age research has not completely stopped.

Paula Meneses (1988, 1992, 1996, 1999, 2004) has published the most extensive body of of post-colonial material on the Mozambican Stone Age. The majority of her work has been focused on ESA Acheulean sites in the southern part of the country, but she has also published summaries of the history and current (as of 1988) status of Stone Age archaeology in Mozambique. Other examples of post-colonial Stone Age reporting include brief discussions of MSA and LSA lithics encountered by chance in Iron Age excavations at sites such as Caimane Rock Shelter in southern Mozambique (Morais

1988:113-116) and Chinde, Marromeu, and Sena in central Mozambique (Madiquida

2006). Recently, attention has once again been paid to Stone Age archaeology in

Mozambique. Portuguese archaeologist Rodrigues (2007) published an analysis of lithic assemblages collected in Tete province by Santos Junior in 1955 and archaeologists from

44

Uppsala University in Sweden are conducting new analyses on Acheulean lithics originally collected by Meneses and are also in the early stages of examining new lithics found in the Maputo area (Sillén, personal communication).

3.2 – Archaeology in the Niassa Rift

The renewed interest in Mozambican Stone Age archaeology is also represented by the research conducted in western Niassa that grew out of a University of Calgary based project called Projecto Património Arqueológico e Cultural (PAC; the Cultural and

Archaeological Patrimony Project). The excavation of Mvumu, the site upon which this thesis is based, was conducted as part of this project. PAC began working in

Mozambique in 2003 when Julio Mercader first conducted survey work in Niassa province (Figure 3.2). Since the initial survey work, PAC has grown to become an international research effort including members based in Canada, the , and

Mozambique and has produced a number of articles and conference presentations focused on Stone Age archaeology, palaeobotany, and human ecology (e.g. Mercader and

Fogelman 2006; Mercader et al 2008; Mercader 2009; Mercader et al 2009a; Mercader et al 2009b; Bennett et al 2010; Mercader et al 2010; Mercader et al, in press; Mercader et al, in prep). This chapter will now turn its focus to PAC‟s work in Niassa beginning with a brief description of the physiography of the greater study area. This is followed by more in depth descriptions of the two districts in Niassa, Sanga and Lago, where PAC has conducted archaeological investigations and some details of the work that led up to the excavation of Mvumu. And lastly, it will present the site of Mvumu and the work done

45

______

______Figure 3.2 – PAC Survey Area and Study Area Geology (Mercader et al, in prep)

46

there in 2007-2008.

3.2.1 – Physiography of the study area

On the Mozambican side of the basin, recent geological cartography conducted by

Norwegian geologists at 1:250000 (Norwegian Geological Survey 2004) shows that the study area spans the Proterozoic and Phanerozoic eras back to 2500mya (Figure 3.2). It falls within the “Bandawe/Metangula/Nkotakota” structural rift province (Chapola and

Kaphwiyo 1992) and is dominated by Proterozoic (2500-600mya) granitoids

(granodiorite, granosyenite, and granite) with the crystalline basement outcropping in the southeastern half of the study area. During the Neoproterozoic (~600mya) small patches of carbonate deposition were overlain on the granitoid basement forming the Malulu formation. The second most abundant materials (gres, siltstone, and sandstone) come from the lacustrine and fluviatile basin known as the Maniamba carboniferous basin (300-150mya) (Verniers et al 1978). This geological unit also includes volcanics such as from the Ecca, Beaufort, and Stromberg lithostratigraphic units

(Catuneau et al 2005). The Mikindani formation, dating to the Mio-Pliocene (10-2mya), is the third most common geological unit in the study area. It runs northeast to southwest and consists of onshore sediments from an upper deltaic complex characterized by red cross-stratified sands and arkoses with conglomerate facies (Salman and Abdula 1995;

Alfonso et al 1998; Lächelt 2004).

During the Quaternary (the last 2my), portions of the landscape have been capped by Pleistocene colluvial pediments and Holocene alluvial and colluvial infill. Due to

47

their chronology and the MSA focus of PAC‟s research, the Pleistocene pediments are of particular interest here. Despite their discontinuous nature, they make up a significant portion of the landscape as it is today (Mercader et al, in prep; cf. Reid and Frostick

1986). These pediments show significant variation in their sedimentology across the study area. Four sediment profiles, named for the type sites at which they were defined, have been described to explain this variability (Figure 3.3):

Type 1: Mvumu – A clast-supported fine cap sits upon a clast-supported diamicton

gravel bed that forms a stone line. This is followed by semi-compact gravels

that sit unconformably on the gneiss basement.

Type 2: Mitumbati – A coarse clastic bed sitting unconformably upon a Karoo

siltstone basement.

Type 3: Mbalanenga – A clast-supported fine cap sits atop a clast-supported

diamicton gravel bed (stone line). Below this there is a calcimorphic layer

sitting unconformably on a Karoo sandstone basement.

Type 4: Lucambo – Fine clast-supported sediments separated by two clast-

supported diamicton gravel beds (stone lines). These sediments rest

unconformably on Mikindani sandstone.

Archaeological materials have been found in all four sedimentary domains, but the most in depth work has been done in Type 1 – Mvumu sediments.

Work was conducted by PAC in two districts within this region: a highland area called

Sanga and a lowland area called Lago. Elevation varies from 1841m a.s.l. in Sanga to

48

______

Figure 3.3 – Niassa Sediment Column Profiles (Mercader et al, in prep)

465m a.s.l. in Lago (Figure 3.4). Major river systems flowing into Lake Niassa include the Fubue, Lunho, and the Luchamange. In the highlands soils are clayey, oxic, and deep with pH varying from alkaline to acidic (classified as Rhodic Ferrasols by FAO

1998). Lowland soils are shallower and have a silty sand texture and neutral pH

(classified as Ferric Lixisols by FAO 1998; Instituto Nacional de Investigação

Agronómica 1995). Precipitation is unimodal in both districts with a rainy season spanning January to May and a dry, drought-prone season from June to December.

Annual rainfall is higher in the highlands at ~1400mm compared to ~700mm in the lowlands (Gama 1990). Vegetation throughout the area is classified as Zambezian

49

woodland (White 1983). The majority of trees are semi-deciduous, but wetter areas also support some evergreen species. Trees in this region primarily come from the genus

Brachystegia (Bloesch and Mbago 2006) although in some localities it is co-dominant with Julbernardia. This arboreal dominance by members of the Caesalpinioideae sub- family of the Fabaceae provides ecological coherence between the highlands and lowlands and allows both to be classified as Miombo woodland (White 1983; Frost

1996).

______

______

Figure 3.4 – Niassa Basin Profile through Study Area (adapted from Mercader et al 2010)

50

3.2.2 – Archaeology in Sanga District

PAC‟s investigation of the Stone Age archaeology on the Mozambican portion of the Niassa basin began in the Sanga district of Niassa and was focused on discovering bearing MSA human occupations. The Sanga district is a highland area above the

Lake Niassa basin. The study area is bounded by four mountain ranges (Mercader et al

2009a) – Geci (north), Moombela (south), Chipilua (east), and Dilombe (west) – which are part of the Neoproterozoic crystalline granitoid basement (Lächelt 2004). These mountains constitute the third highest elevation in Mozambique (~1800m a.s.l.). During the late Proterozoic (~600mya) this granitoid formation was capped with crystalline limestones (grading from pure limestone to dolomite to marble) that form a geological unit referred to as Malulu in Mozambique‟s geological cartography (Lächelt 2004).

Survey work in Sanga was focused in the vicinity of a village called Njawala.

Several caves with archaeological potential were identified within pyramidal limestone massifs associated with the Malulu formation (Mercader et al 2009a) and one, known as

Nankambe, was selected for excavation during the 2005 and 2007 field seasons. U-series and ESR analyses conducted at Nankambe showed a mid- to late-Pleistocene chronology for the sediments in the cave (Grün 2006). The oldest dates come from chamber three where excavations reached 235cm in depth and revealed at least two distinct sedimentary phases. The older phase pre-dates ~350kya (by U-series on a capping speleothem) and the younger dates to the last glacial cycle at ~80kya (by ESR on tooth) (Grün 2006).

Other dates from the talus and mouth reflect a younger chronology. Excavations in the talus reached 445cm in depth and showed two phases of overhang retreat and roof

51

collapse. Two minimum dates were produced from within the talus, one of ~18kya by U- series on a capping speleothem and one of ~55kya by ESR on tooth, suggesting that the talus sediments pre-date the Last Glacial Maximum (Grün 2006). Testing at the mouth of the cave reached a maximum depth of 100cm and encountered two phases of sedimentation separated by a flowstone which produced a U-series date of ~18kya (Grün

2006). Sediments throughout the cave contained large numbers of quartz fragments and faunal remains (Mercader and Fogelman 2006; Russell and Jamniczky 2006; Brehm

2007). However, although clear products of intentional lithic reduction were found throughout the cave (with the highest concentration near the mouth), many are ambiguous and possibly the product of natural geological processes within the cave. Like the lithics, the faunal materials possess some markers of human modification (e.g. cut marks, long spiral fractures, and pressure flakes), but the majority (~75%) show no indications of human activity (Brehm 2007). Species identified in the faunal collection include many that do not typically inhabit caves such as hartebeest, buffalo, addax, eland, warthog, rhinoceros, leopard, lion, and an unidentified primate (Russell and Jamniczky

2006).

From the latter part of the 2007 field season through 2009 archaeological efforts in Sanga shifted to a nearby cave called Ngalue. Ngalue cave is located in a dolomitic marble massif of the same name in the Chitete River valley (S12˚51.5‟ E35˚11.9‟)

(Mercader et al 2009a). Access to the cave is gained through a mouth ~14m above the current valley floor that leads into an 18m long (2-3m wide) gallery that connects a single main chamber (~13x7m). The roof of the main chamber is open forming a chimney

52

through the top of the massif. Like Nankambe, Ngalue contains a long sedimentary sequence spanning the mid- to late-Pleistocene and into the Holocene (Mercader et al

2009a). The oldest deposits in the cave, the Basal Beds, have been dated to 470±48kya by U-series on speleothem and contain no archaeological materials (Mercader et al

2009a). The Basal Beds are followed by the Lower Beds which are subdivided into

Lower Beds 1 and Lower Bed 2. Lower Bed 1 was deposited prior to 251±51kya when it was capped by speleothem formation (Mercader et al 2009a). At some point in the cave‟s history the majority of the Lower Bed 1 sediments became subject to erosion and transportation such that little of the package remains preserved today (Mercader et al

2009a). Lower Bed 2 consists of a subsequent phase of sedimentation ending 105±13kya

(U-series on speleothem) when it was capped by a stalagmitic conglomerate (Mercader et al 2009a). Lower Bed 2 contained faunal and MSA lithic materials. MSA artifacts and faunal remains continue in the overlying Middle Beds which begin at the speleothem that capped Lower Bed 2 and continue until they are capped themselves by a laminated flowstone dated to 55±5kya by U-series (Mercader et al 2009a). This date is further supported by two ESR dates (on tooth) of ~50kya from the base of the following Hearth

Beds (Mercader et al 2009a). In addition to their namesake hearths, the Hearth Beds preserve faunal remains and contain a continuation of Ngalue‟s MSA lithic component.

The upper margin of the Hearth Beds is marked by an AMS 14C date from a hearth of

>42kya (Mercader et al 2009a). The most recent phase of sedimentation is referred to as the Capping Beds and has been dated to AD900-1040 by AMS 14C (Mercader et al

2009a).

53

The MSA artifacts in Lower Bed 2, the Middle Beds, and the Hearth Beds at

Ngalue proved to be much richer in unambiguous lithic artifacts compared to Nankambe.

The excavated lithic assemblage from these deposits consists of >1000 artifacts, 555 of which were selected for techno-typological analysis (see Mercader et al 2009a). Further discussion of the lithics from Ngalue can be found in chapter 6.4 of this thesis. The

Capping Beds also contained Iron Age artifacts including iron tools, a complete pot, and numerous pot sherds. Faunal preservation is good in the Lower, Middle, Hearth, and

Capping Beds and a total of 14958 bone and tooth fragments were recovered. No faunal remains were found in the Basal Beds.

3.2.3 Archaeology in Lago District

Although excavations were first carried out in the highland region, PAC has also conducted large scale survey work in Lago district that began in 2003 and continued intermittently through 2009. The survey covered the area between Cobue and Meluluca;

>100km north-south along the coast of Lake Niassa and extended >30km inland spanning elevations from 507-871m a.s.l. (Mercader et al, in prep). This work identified 19 localities where lithic surface collections were made totalling >10000 artifacts (Mercader et al, in prep). These surface scatters occur in areas possessing all four sediment profiles presented above for the Pleistocene pediments in the study area. Sites associated with profile Type 1 have been the subject of the most detailed analysis. Two in particular,

Mikuyu and Mvumu, have been most thoroughly investigated.

54

No name had yet been proposed to refer to the Type 1 pediments in the

Mikuyu/Mvumu area when PAC began investigating them so we have taken the liberty of temporarily naming them the Luchamange Beds in reflection of their proximity to the

Luchamange River basin (Mercader et al, in prep). The Luchamange Beds range from

480-580m a.s.l. Data from over 50 geological pits show that sedimentation phases, layer thickness, and basements are variable throughout these beds (Mercader et al, in prep).

The Luchamange Beds range from ~20-105cm in depth.

The first Lago district site to be investigated in detail was Mikuyu. It was identified as having high archaeological potential near the end of the 2005 field season when significant surface finds of MSA lithics were made by Mercader and PAC cultural anthropologist Arianna Fogelman (see Mercader et al 2008; Raja 2008; Bennett, n.d.) in the cassava field of a local potter. Mikuyu was excavated by PAC in 2006. The site is located on a hillside 865m inland of Lake Niassa (E34˚49.5‟ S12˚43.4‟. Test pits range in elevation from 480-535m a.s.l. The highest point of the site is near a hilltop divide that is 67m above current lake levels. Sediments at Mikuyu are classified as silty sand (58% sand, 39% silt, 3% clay) and have a mean particle size of 89.93µm. Clay distribution remains consistent throughout the sequence.

Seven archaeological and geological excavations (T1-T7) were opened at

Mikuyu. The largest and most productive from the archaeological point of view was T6

(see Mercader et al 2008). In total, approximately 8000 lithic artifacts were recovered during these excavations. A preliminary study of ~1000 of these lithics was conducted by Raja (2008), 76 were examined by Bennett (n.d.), and 33 were subjected to starch

55

residue analysis by Mercader et al (2008). The assemblage is dominated by quartz lithics with no faunal or other organic artifacts being recovered. A technological summary of the Mikuyu lithics is presented in chapter 6.4. The quartz used in manufacturing these tools is abundant in the hills surrounding Mikuyu (see chapter 5.2 for detail on quartz sources in the Luchamange Beds).

While working at Mikuyu it became clear that the surface finds that allowed the initial identification of the site were not limited to the Mikuyu locality itself. This prompted more detailed pedestrian survey in the hills to the south during the 2007 field season that identified an essentially continuous lithic scatter along the Luchamange Beds between Mikuyu and Nkolongwe with variable artifact density (Figure 3.2). Sites were identified in areas where the concentration of surface artifacts is higher than average.

One of these sites, called Mvumu, had an especially rich surface scatter and was selected for excavation beginning in the 2007 field season. The following section of this chapter will provide details of this work.

3.3 – Mvumu

Mvumu (E34˚ 49.149‟ S12˚ 44.099‟; Figure 3.1; Figure 3.5) is located south of the village of Mikuyu approximately 200m inland from the shore of Lake Niassa at the top of a steep escarpment. The site was first identified through a dense scatter of quartz lithics covering approximately 10ha (500m east-west, 200m north-south) that occurs between

65-115m above current lake level (530-580m a.s.l.). In addition to the artifactual quartz, unmodified blocks of pegmatitic quartz and sub-rounded quartz cobbles are present with

56

______

Figure 3.5 – Main Excavation Trench, Study Assemblage Source Units, and In-Situ Artifact locations (Mercader et al, in prep)

increasing abundance at higher elevations (Mercader et al, in prep). This abundance of readily available raw materials may have played a role in drawing people to the site to conduct both quarrying and tool production activities. However, the lithic assemblage lacks some attributes commonly associated with quarry sites (e.g. large numbers of hammerstones and a high ratio of unfinished tools relative to finished tools – Renfrew and Bahn 2008). Moreover, we see a similar diversity of lithic technology at Ngalue which is not interpreted as a quarry site. Therefore, while raw material procurement was likely one reason MSA people visited Mvumu, it seems likely that the site also served

57

other functions as well. One possibility is that Mvumu was a campsite that was stragegically located to allow easy access to lithic raw materials.

The MSA deposits at Mvumu occur within the Luchamange Beds‟ sediment profile Type 1 which was defined based on work at the site and consists of three stratigraphic units (Mercader et al, in prep): a compact gravel layer at the base (Figure

3.6:3), a clast-supported massive diamicton to poorly sorted massive gravel bed above it

(Figure 3.6:2), and massive clast-supported matrix with a fine cap at the top (Figure

3.6:1a, 1b). All three units contain archaeological materials, but the majority of the MSA lithics discussed in this thesis were recovered from a stone line 5-12cm thick horizon that

______

Figure 3.6 – Mvumu Stratigraphy (Mercader et al, in prep)

58

undulates 24-44cm below the current ground surface (Figure 3.6:2; Figure 3.7). Organic artifacts such as bones and teeth are not preserved in these sediments.

Field methodologies were devised by Mercader, Bennett, and Gosse. They were adapted from methods used previously at Mikuyu in 2006. Work at Mvumu began with a series of 10 geological test pits dug around the periphery of a dense area of surface artifacts to evaluate the sedimentary conditions and subsurface archaeological potential.

The data gained from these test pits was used to guide the location of a 10x10m grid that formed the main excavation area. A total of 26m2 were excavated within this grid during the 2007 field season (Figure 3.5). Initially, a series of eight 1x1m test pits located in units A1, A5, A10, E1, E6, J1, J4, and J10 were excavated in 10cm spits were dug down to the C-horizon, sieving ¼ of excavated sediments from each spit, to gain a better understanding of the morphology and richness of the subsurface deposits. These tests were followed by two additional 1x1m test pits in units B6 and H8 that employed a comprehensive excavation methodology that included 5cm spits, sieving of all sediments, and planview drawings of pit floors that plot in situ artifact locations. The knowledge gained through these test pits was used to choose the location for the main trench excavated at Mvumu. The main trench accounted for 16m2 of the total excavation and included units A7, A8, B7, B8, C6, C7, C8, D6, D7, D8, E7, E8, F6, F7, G6, and G7.

Each 1m2 unit in the main trench was divided into sixteen 25cm2 squares and excavated in arbitrary 5cm spits. Sediments from each square were wet sieved separately to minimize relocation error for artifacts recovered in the sieve using a 1.87mm screen.

Artifacts recovered in situ (n=193) were plotted in three dimensions and drawn in

59

______

Figure 3.7 – Mvumu Stone line (photograph by J. Mercader, modified by T. Bennett)

60

planview sketches for spits two to five. Detailed plotting was abandoned from spit six onward due to a reduction in artifacts and time constraints as the season was coming to an end. Additional geological trenches were opened in 2008 by Mercader and John Gosse to better characterize the geological history of the site and retrieve samples for terrestrial cosmogenic nuclide dating (see below for more on dating work).

Sediments at Mvumu are classified as brown silty sand (68% sand, 26% silt, 6% clay) with a mean particle size of 161µm (Mercader et al, in prep). Clay increases with depth. This differs from nearby Mikuyu where clay distribution is homogeneous. The sediments have a neutral pH with ~6% organic carbon and increasing water content with depth in the column (Mercader et al, in prep). X-ray diffraction conducted on sediments from Mvumu (University of Calgary‟s Microbeam: JEOL JXA-8200; 5 WDS) show them to be primarily quartz with some k- and traces of zircon, Ti-oxide, and biotite

(Mercader et al, in prep). Cobbles within the matrix are irregularly oriented with horizontal, inclined, and vertical dips (Mercader et al, in prep).

Soil at Mvumu is classified as a Ferric Lixisol. Within this three horizons were distinguished above the bedrock at Mvumu (Mercader et al, in prep). The uppermost is humic, followed by a leaching zone, then clay accumulation. Several values of brown are represented in the soil profile. The transitions between soil horizons vary from sharp to gradual generally becoming more gradual with depth. The uppermost zone of clay accumulation coincides with a sedimentary unconformity (the stone line). Cementation of soils increases with depth beginning at this boundary. Above the stone line particle size fraction becomes increasingly finer.

61

Given the lack of organic preservation at Mvumu, and the lack of other datable materials such as volcanics and carbonate formations such as speleothems, determining the chronology of Mvumu was problematic. None of the absolute dating methods commonly applied in MSA research – electron spin resonance, thermoluminescence, optically stimulated luminescence, uranium-series, potassium-argon/argon-argon – were viable options for the site so an alternative solution was required. This solution came in the form of terrestrial cosmogenic nuclide (TCN) dating, a technique that is revolutionizing historical geomorphology (Siame et al 2006; Dunai 2010). TCN dating is relatively uncommon in African archaeology, but has also been applied in several other

African contexts including dating the Oldowan-Acheulean boundary in South Africa

(Chazan et al 2008; Gibbon et al 2009) and to date fossil hominin sites in Chad (Lebatard et al 2008) and a number of other well known Palaeolithic sites including Atapuerca

(Carbonell et al 2008) and (Shen et al 2009). This is, however, its first application to MSA research.

In the analysis of Mvumu‟s geochronology, John Gosse of the Dalhousie

Geochronology Center calculated the accumulation and decay of 10Be and 26Al in quartz from multiple column samples from Mvumu following the depth profile method (Granger and Muzikar 2001:277). These two nuclides are analyzed independently and the combined results give greater accuracy than single nuclide studies. The analysis of these nuclides allowed for the chronological reconstruction of the depositional history of sediments at Mvumu. In some ways this technique is analogous to using the redness on a person's skin – the amount of cosmogenic nuclides present – to estimate the duration of

62

exposure to sunlight – the age of the sediment being dated (Gosse and Phillips 2001).

The analysis showed that the stratigraphic unit underlying the stone line was deposited rapidly ~34.5kya (Figure 3.6:3). This deposition was followed by a period of erosion lasting until ~33kya. The stone line containing the cultural horizon (Figure 3.6:2) formed shortly after this erosional event and therefore the human occupation responsible for the lithic assemblage must be <33ky old. The youngest well dated MSA sites in Africa are

~30-28ky old (see chapter 6.4 for detailed discussion). I use these dates are used to further constrain the age of the Mvumu assemblage and therefore maintain that it dates to between ~33-28ky old. Although nearby Mikuyu has not been directly dated it is stratigraphically correlated with Mvumu and therefore this date is arguably also valid for

Mikuyu.

63

Chapter Four: Analytical Methods and Lithic Classification Scheme

4.1 – Study assemblage and methodological approach

The lithic assemblage studied here consists of 1977 stone artifacts from four of the sixteen units in the main trench at Mvumu: A7, A8, B7, and B8 (Figure 3.5). These excavation units were selected near the end of the 2007 field season when it was clear that they represented a thick portion of the Mvumu MSA horizon containing a dense concentration of artifacts. The collection includes a wide selection of cores, debitage, and tools that preliminary field classification suggested make up a good representative sample of the lithic variation seen at Mvumu. Upon the completion of the fieldwork component of this study the artifacts were brought to the University of Calgary Tropical

Archaeology Laboratory for in-depth analysis including the identification of lithic raw materials (chapter 4.2), techno-typological classification (chapter 4.3), and the recording of metric and qualitative traits (chapter 4.4). The details of these methods and the techno-typological classification framework are described below. The results of these analyses are presented in chapter 5. The final section of this chapter discusses experimental flintknapping done in conjunction with the main techno-typological analysis (chapter 4.5).

4.2 – Raw material analytical methods

Raw material analysis of the Mvumu study assemblage was done on a macroscopic scale. First, the type of stone represented by each artifact was identified through visual inspection and comparison with known rocks from the area. Each type

64

was quantified and the range of colours observed was recorded. Colour determinations follow the Munsell Soil Colour Chart. This basic level of identification was sufficient for non-quartz raw materials, but several varieties of quartz were identified that required additional levels of differentiation. The main attribute used to distinguish between these varieties of quartz was the texture of the stone. Texture here refers to the size of the crystal structures that form a specific piece of quartz. Texture can be evaluated with the naked eye (Seong 2004) and through tactile evaluation of a piece of quartz. Using these methods each piece was recorded as fine, coarse, or very coarse. Lastly, attempts were made to identify whether pieces of quartz were obtained from pegmatitic outcrops or lacustrine/alluvial/Karoo sources based on cortex development; pegmatitic quartz is typically rough and irregular whereas lacustrine/alluvial/Karoo quartz develops smooth rounded cortex. Such determinations proved difficult with most of the assemblage because artifacts lack sufficient cortical coverage, but some cores did retain adequate amounts of cortex to allow raw material source identification.

4.3 Lithic Artifact Analytical Methodology

Lithic artifacts were classified into three broad categories: cores, debitage, and tools. Each broad category was then split into a number of subtypes to account for the variability seen within the category. The classification is primarily based on technological criteria that reflect the organization of lithic reduction at Mvumu. In a few cases, elements of typological classification were also used to help distinguish between subtypes. These typological criteria were especially useful in the classifications of tools

65

and typically focus on the location of retouch or working edges (Bordes 1961, Chazan

1997). To help facilitate inter-site comparison, efforts were made to use existing techno- typological classification schemes from other published MSA assemblages to classify the artifacts from Mvumu. The sources of these techno-types are cited below in the techno- typological definitions. In cases where suitable techno-types where not already present in published literature new classifications were created. These created types are cited as

“this study” in the techno-typological definitions presented below. Photographs of lithics representing these types are presented in chapter 5.

4.3.1 – Core classification scheme

Inizan et al (1992:84) define cores as a block of raw material from which flakes, blades, or bladelets have been removed to create blanks for tools. In general, the Mvumu cores fall into two broad classes: simple reduction and prepared core technology. These distinctions closely follow Clark and Kleindienst‟s (2001:60-61) “unspecialized” (simple) and “specialized” (prepared) core types. Simple reduction cores are those that show no signs of their surfaces being intentionally modified prior to the removal of a flake to influence its morphology. Conversely, prepared cores do show deliberate modification to the flake release face or faces and sometimes to the striking platform or platforms aimed at producing a more consistently shaped flake blank (Clark and Kleindienst 2001:61).

These general classes are then further divided into a total of eight techno-typological categories that include four simple reduction techno-types – bipolar cores, elongated high-backed cores, simple cores, and testing cores – and four prepared core types –

66

discoidal, Levallois, preferential discoidal (single and recurrent), and unifacial radial.

Core techno-types are defined below.

4.3.1.1 – Simple Cores

Simple core reduction strategies at Mvumu are defined as follows:

Bipolar cores (Andrefsky 2005:153-155; Kamuanga 1982:212; Kuhn

1995:97; Clark and Kleindienst 2001:46, 61): These cores are unique at

Mvumu in their use of a hammer and anvil reduction technique where the

core is rested upon the anvil and struck from above at approximately a 90˚

angle. When properly executed, this blow results in a flake being detached at

the point of impact as with other types of cores, but the anvil also serves to

reflect some force back up into the core from beneath causing a flake to be

detached from the area opposed to the striking platform. Cores will therefore

often show two opposed removals with one originating at a primary striking

platform where the percussor struck the objective and the other from a

secondary platform where the objective meets the anvil. This secondary

platform may look like a typical striking platform or merely appear as a

battered end. If the opposed flake fails to detach, the battering on the anvil

end may be the only indication that the bipolar technique was applied to a

core (Clark and Kleindienst 2001). Bipolar cores are usually small in size

(Andrefsky 2005) and often take on a lenticular profile as a result of the

paired removals from the ends thinning as they approach the center

67

(Kooyman 2000). This technique can also be useful in the initial splitting of cobbles, but such a use does carry with it a risk of causing the cobble to shatter into wedge-shaped pieces.

Elongated high-backed cores (this study): Unlike other simple cores these artifacts do have a consistent morphology. They have a triangular cross section with long faces that taper at the ends and come together to form a trihedral pick-like point. Removals are made either bifacially or trifacially with the angle between sides serving as a striking platform. In cases where flake scars are only present on two surfaces the third is flat, featureless, possibly the result of the removal of a platform rejuvenation flake to allow continued flaking on the adjacent sides. Some of these cores retain morphological features that show that they were made on large, thick flakes such as intact striking platforms and bulbs of percussion. Cores that possess these flake features are considered a subtype of elongated high-backed cores and referred to as elongated high-backed cores on flake. Those without flake features are simply called elongated high-backed cores.

Simple cores (this study): Characterized by irregular blocky polyhedral morphology that shows no deliberately imposed form, these cores provide an expedient source of non-standardized flakes. As suggested by the name, they require minimal pre-planning to create; selection of a cobble or chunk of

68

suitable size that possesses an edge with a suitable angle for flint knapping is

all that is required. Minimal preparation to create this first platform in the

form of primary splitting may be necessary if the raw material being utilized

is a rounded cobble (Singer and Wymer 1982). Once the initial platform

becomes unusable, the knapper would either move on to a different part of

the core or discard it in favour of another. Simple cores closely follow

several core types described by Clark and Kleindienst (2001) including single

platform, double platform, two platforms on different planes, and

multidirectional. Shea (2008) refers to similar cores as “polyhedral”.

“Testing” cores (this study) are cobbles or chunks that have had a minimal

number of flakes removed as if the knapper was testing the quality of the raw

material, then discarded it without further reduction. It is possible that a few

useful flakes were removed from these cores, but it appears that they were

abandoned for some reason before they were developed into a more formal

type.

4.3.1.2 – Prepared cores

Prepared core technologies at Mvumu are defined as follows:

Discoidal cores (Barham 2000; Clark 2001b; Shea 2008; Tryon 2006; Yellen

1996): These roughly circular cores are lenticular in profile and defined by

the presence of two distinct, non-hierarchically exploited flaking volumes

69

divided along the equatorial axis. The striking platforms for both volumes run along the lenticular margin of the core with flakes being struck bifacially and inward towards the core‟s center. Each subsequent series of removals from a given volume also serves to prepare a striking platform for the next series of removals from the opposing volume. As reduction continues the disc will become smaller in diameter and each volume will become more conical until further reduction becomes impossible. Sometimes a core that has become overly conical can be rejuvenated by removing the apex of the cone which allows reduction to continue (Barham 1995a). Some discoidal cores from Mvumu have a small facet on or near the lenticular margin, referred to as a side facet (this study), which has not been reduced according to the same method as the rest of the core. This facet often remains cortical.

Levallois cores (Boëda 1995; Chazan 1997; Pleurdeau 2005; Tryon et al

2005; Wilkins et al 2010): Levallois cores were originally defined by Bordes

(1961) on largely typological grounds. However, this study follows a more recent technological definition put forth by Boëda (1995; Chazan 1997;

Pleurdeau 2005). Technological criteria for Levellois technology include: 1) two opposed volumes with a non-reversible hierarchical relationship where only one is used to produce flake blanks, 2) lateral and proximal-distal convexities on the utilized flaking volume, 3) flake detachment planes parallel to the intersection of the two volumes, and 4) the use of hard hammer

70

percussion on prepared platforms oriented at 90˚ to the flaking axis. These technological features are aimed at producing a prepared flake with a specific predetermined morphology. In most cases a single prepared flake blank is removed from the core before the preparation cycle begins again, but occasionally recurrent removals are evident. Levallois cores at Mvumu can be high-backed or relatively flat with bases that are cortical or partially reduced, but if bases are reduced removals are limited to small flakes aimed at preparing the shape of the core and not at producing usable blanks.

Preferential discoidal cores (Mercader et al 2009a; definition expanded in this study): These cores represent a hybrid between the discoidal and

Levallois approaches described above. Like discoidal cores they are roughly circular with two distinct flaking volumes and a lenticular profile, however, like Levallois cores there is evidence of a hierarchical relationship between these volumes. Although both volumes of the core have been used to produce useable flake blanks, one volume is preferentially exploited and used to produce one or more large, regular flakes. Flake scars on the non- preferential side tend to be smaller and less regular. This exploitation of the non-preferential side sets them apart from Levallois cores that show similar preferential treatment of one volume, but no evidence for blank production from the opposing side. There are two variants within this subtype: single and recurrent. Single preferential discoidal cores possess only a single

71

preferential removal from the preferentially treated volume whereas recurrent

preferential discoidal cores have two or more. In the recurrent variety

preferential flakes most often continue to follow a radial pattern, but may also

be struck from parallel or opposed platforms.

Unifacial radial cores (this study): Examples of this subtype exhibit the

same type of radial flaking present on discoidal cores, but reduction is

unifacial with the opposing side remaining cortical. The periphery of these

cores is often prepared to create steep angled striking platforms and they are

typically quite flat in cross section. Considering their cortical bases, their

flatness suggests that flat circular cobbles were preferred for these cores.

Some researchers (see Clark and Kleindienst 2001) include unifacially

reduced radial cores in the same category as bifacial radial cores such as

those defined as discoidal in this study. However, in the case of these

unifacial racial cores there is a pronounced difference in the technological

approach to core reduction that sets them apart from all other radial techno-

types from Mvumu.

4.3.2 – Debitage classification scheme

Debitage includes all by-products of core reduction that are not cores themselves or formal tools (Andrefsky 2005; Debénath and Dibble 1994). In this study debitage is divided into three general categories: flakes, core preparation products, and angular

72

waste. Flakes and core preparation products represent intentional by-products of flaking activities. This sets them apart from angular waste which results from the unintentional shattering of a core during reduction. The key criteria for discriminating intentional from unintentional by-products are the presence of formal flake features (Andrefsky 2005) including identifiable dorsal and ventral sides, proximal and distal ends, striking platforms, and bulbs of percussion. It is not necessary for an artifact to possess the full suite of these features to be considered an intentional product of reduction. Amongst intentional debitage, flakes are differentiated from core preparation products in that they are produced for use as blanks for retouched tools, to modify or re-sharpen the edge of a retouched tool, to shape a core tool, or simply to create a sharp cutting edge. Core preparation products are the result of attempts to remove undesirable portions of a core such as protruding ridges or to rejuvenate a striking platform to allow continued reduction (Clark and Kleindienst 2001). Each general category is also subdivided into more refined subcategories that capture further techno-typological details. The details of these subtypes are presented below:

4.3.2.1 – Flakes

Two general categories of flakes are recognized at Mvumu: whole flakes and flake fragments. Within these categories flake debitage is divided into several subtypes that capture the technological and morphological variation seen in the assemblage.

Flakes are defined as follows:

73

Whole flakes (this study) are those that retain both proximal and distal ends with observable striking platforms and terminations (Inizan et al 1999;

Kooyman 2000; Andrefsky 2005). Within the whole flake class artifacts are further divided into two varieties: whole flakes and blade-like flakes.

Artifacts that fall under the general whole flakes label possess a large range of morphologies, but those that are considered blade-like must meet a set of specific morphological criteria. Blade-like flakes (this study) have parallel to subparallel lateral edges that measure at least twice as long proximal to distal as they do laterally (Bordes 1961; Inizan et al 1999; Kooyman 2000).

Although these are the same criteria commonly used to define formal blades by most lithic researchers (e.g. Inizan et al 1999; Kooyman 2000; Andrefsky

2005; Johnson and McBrearty 2010), they are considered “blade-like” rather than true blades because no blade core technology has been identified at

Mvumu.

Flake fragments (this study) are pieces of debitage that possess sufficient formal flake features to allow their identification as flakes, but are broken or damaged in a way that renders them incomplete. Flake fragments are subdivided based on the portion of the flake that they represent. Proximal flake fragments are distinguished by the presence of a striking platform, medial flake fragments show breaks on two ends and possess neither a striking platform nor termination, and distal flake fragments are the terminal

74

end of a broken flake. Some broken flakes show signs that they were intentionally truncated by laying the flake flat against an anvil and striking them with a hammerstone (Moore et al 2009; Brumm et al 2010; Figure 4.1).

Brumm et al (2010:460) refer to this method of truncation as cross-axis percussion. Evidence of cross-axis percussion manifests in several ways including recognizable impact areas at the point of percussion, crushing and shattering around the impact point, and the presence of “demi-cones” (Moore et al 2009). A “demi-cone” is a negative conchoidal scar that is sometimes

______

______

Figure 4.1 – The Truncation Technique (adapted from Moore et al 2009)

75

______Figure 4.2 – Demi-Cone on Truncated Piece

produced on the fractured facet of a truncated flake (Figure 4.2). Only flakes

that clearly show at least one of these traits were recorded as truncated flakes.

Truncated flakes were also subdivided based on the portion of the original

flake that they represent: proximal, distal, and lateral. Proximal and distal

follow the same criteria used for flake fragments. Lateral refers to flakes that

were truncation proximal to distal.

4.3.2.1 – Core preparation products

Core preparation products fall into two subtypes: core trimming spalls and core preparation tablets. They are defined as follows:

76

Core trimming spalls (this study) are the result of attempts to remove

undesired ridges and protrusions, often related to hinge and step

terminations, from a core that prevent the knapper from removing a desired

flake. They are characterized by a thick triangular cross section with three

relatively equal sides. Some examples that were detached through

percussion possess formal flaking features, but experimental knapping (this

study) showed that undesirable ridges could often be removed from cores by

simply rubbing/grinding the core with a hammerstone.

Core rejuvenating spalls (Clark and Kleindienst 2001; referred to as core

trimming elements by Shea 2008) result from the removal of all or a part of

an existing platform which has become too battered or rounded through use

to allow for the creation of the desired platform angle. The resulting flakes

share many of the features seen on other types of flake debitage, but are set

apart by a thick tabular cross-section. The outer margin of the dorsal face

often shows signs of battering from its use as a striking platform and the

sides of the tablet will retain significant amounts of residual core edge (Shea

2008).

4.3.2.2 – Angular waste

Angular waste (this study), also referred to as chunks (Clark and Kleindienst

2001:62) or shatter (Kooyman 2000:14-15), appears as irregular, polyhedral pieces of

77

stone that resulted from the unintentional fracture of a core during reduction. These artifacts lack any flake features, but are distinguished from naturally occurring rocks by the presence of many fracture facets related to the knapping process.

4.3.3 – Tool classification scheme

Tools are defined in this study as pieces that possess one or more edges or areas that are suitable for performing a specific task and do not appear to have been primarily used as cores. Working edges on tools may or may not be retouched (Inizan et al

1992:99; cf. Kuhn 2007:268). Usewear analysis was not conducted as a part of this study so while it is presumed that at least some of these tools were used, use cannot be confirmed. The majority of tools from Mvumu fall into three broad categories that are typical of the MSA: scrapers, awls, and points. Each of these categories contains a number of subtypes that further define the techno-typology of specific artifacts. These subtypes include 12 varieties of scrapers, five varieties of awls, and four types of points and are described below. In addition there are eight additional tool techno-types that account for less common, but distinct, tools bringing the total number of tool techno- types to 29. Tools techno-types are defined below.

4.3.3.1 – Scrapers

Scraper techno-types are primarily differentiated based on the location and morphology of their working edges. The assessment of these edge criteria follows the structure laid out in Clark and Kleindienst (2001) with the modification that their

78

“retouch category” can refer to the morphology of either a retouched or unretouched working edge. Within the general edge type categories several varieties are also recognized. In some cases techno-typological attributes of scrapers bodies set them apart as unique and are instead used as the basis of classification. These general techno-types also include several varieties within them. Scraper techno-types are described here by the traits used to classify them – edge classified and body classified –and include:

Edge classified scrapers

Convex scrapers (Clark and Kleindienst 2001) are made on either a flake

or core fragment blank and have a convex scraping edge. The working edge

may be retouched or simply take advantage of an existing edge angle.

Convex scrapers are further divided into three subtypes based on the

location of the working edge: side, end, and side-and-end. Side and end

varieties are characterized by a single working edge. In the case of

examples made on flake fragments, convex side scrapers have their working

edge on a lateral edge of the flake and convex end scrapers have the

working edge located on the distal portion of the blank (Clark and

Kleindienst 2001). Examples made on core fragments are differentiated by

whether the working edge is located on the long or short axis of the scraper;

those with the working edge on the long axis are referred to as convex side

scrapers and those with the working edge on the short axis are convex end

79

scrapers. Side-and-end scrapers combine both side and end scraping edges as described above on a single tool (Clark and Kleindienst 2001).

Concave scrapers (Clark and Kleindienst 2001) can be made on flake blanks or core fragments and are characterized by concave working edges created through retouch. They occur as both side and end varieties that are distinguished by the same criteria as side and end convex scrapers.

Concave-convex scrapers (this study) combine one concave and one convex scraping edge on a single tool.

Convergent scrapers (Clark and Kleindienst 2001) have two unifacially retouched edges or one retouched edge and another suitable working edge created by a naturally occurring edge angle that come together to form a point. Made of both flake blanks and core fragments, these artifacts tend to have a thick cross section that prevents them from being considered any kind of projectile point despite their triangular planform (Shea 2006).

Denticulate scrapers (Clark and Kleindienst 2001) possess a serrated or

“toothed” convex working edge created by retouch. Retouch can occur on either the side or end of a denticulate scraper and they can be made on flake blanks or core fragments.

80

Semi-circular scrapers (this study) are made on both flake blanks and core

fragments. They are distinguished by the presence of unifacial retouch

around at least 50% of their periphery to create a functional scraping edge.

Often these scrapers are broken in half such that they possess a “D”

morphology. Examples with this planform tend to be retouched along the

entire convex margin. They are similar to Clark and Kleindienst‟s (2001)

“circular” category, but the retouch does not necessarily continue around the

blank to the same extent. An additional variant on this subtype is the

denticulate semi-circular scraper which simply adds denticulation to the

retouched edge.

Body classified scrapers

Core scrapers (Coon et al 1968; Clark and Kleindienst 2001) are high-

backed scrapers with a flat ventral surface – likely the product of the

removal of a single tabular flake similar to a rejuvenation tablet – and steep

retouch along at least one scraping edge. Their name suggests that they are

often made on extinguished cores or core fragments, but it is possible that a

core scraper was made deliberately from a cobble without ever serving as a

core. Most core scrapers at Mvumu have convex working edges, but

denticulate core scrapers were also recorded.

81

Bevel-based scrapers (this study) are generally similar to convex scrapers,

except that their bases – the surface that faces the surface that is being

scraped during use – have been modified to create a bevelled plane. The

bevelled base is created by removing one or more flakes from the base of

the scraper. They can be made on either flake blanks or core fragments.

The working edge is typically, but not exclusively, convex and oriented as a

side scraper. The closest to an equivalent type in the existing MSA

literature is a bevel-based core scraper presented by Clark and Kleindienst

(2001) in their typology for Kalambo Falls, but except for the bevelled base

there are no techno-typological connections between the two types; even

Mvumu bevel-based scrapers made of core fragments cannot be considered

a type of core scraper because they lack several of the other defining

characteristics of core scrapers (e.g. high-backed, steep retouch).

Scraper/Awls (this study) are a hybrid tool type that combines the traits of

both a convex side scraper and a single truncation awl (see below). These

tools are made on both flake blanks and core fragments and all examples

show signs of being retouched to form the working edges.

4.3.3.2 – Awls

Most awls are primarily classified according to the technological approach used in their production: truncation or retouch. A third category, topknot awls, is distinguished

82

by the use of topknot flakes (Barham 1995a; see below) as blanks. Awl techno-types are defined as follows:

Truncation awls (this study) are made on flake blanks or core fragments

and are characterized by a thick, relatively blunt “drill bit” at their working

end that is produced using a variant of the truncation technique described

above in the debitage techno-typology. For further discussion of the

technique used to manufacture these tools see chapter 6.2.3. Truncation

awls are further subdivided by the number of truncation blows involved in

their manufacture; single truncation awls use one truncation and one

naturally occurring edge to create a “drill bit”, double truncation awls use

two truncations to that come together to create a “drill bit”), and triple

truncation awls the use three adjacent truncations to create two “drill bits”.

Double retouched awls (this study) are made by applying unifacial retouch

to the lateral sides of a flake blank to narrow them down to a thin “drill bit”

that ends in a sharp point. These tools are similar to artifacts referred to as

“borers” by Clark and Kleindienst (2001).

Topknot awls (this study) are made on topknot flakes (as defined by

Barham 1995a) which are a type of rejuvenation flake used to extend the

life of a radially reduced core by removing the apex where flakes meet in

83

the middle of the reduction surface when it becomes too cone shaped or is

affected by undesirable hinge and step terminations. Topknot flakes are

distinguished by a pyramidal morphology. The arrises that radiate from the

peak of the pyramid act as guides that create a kite-shaped planform in these

flakes. To create the awl the longest corner of the kite is typically retouched

to create the final working end.

4.3.3.3 – Points

Points are a characteristic artifact of the MSA (McBrearty and Tryon 2006) and were one of the key criteria in its original definition (Goodwin 1929:98). At Mvumu, points fall into three general categories that are based on the type of flake blank on which they were made: base-struck, Levallois, and corner-struck. Point techno-types observed in the Mvumu study assemblage are defined as follows:

Base-struck points (this study) are made on triangular flakes where the

striking platform forms the base of the point. The base may be thinned or

remain at its full thickness, but regardless of thinning it is typically faceted.

These points tend to be relatively short with a length to width ratio

approaching 1:1. Retouch on these points is done unifacially and is only

seen on the dorsal surface of the flake blank. Similar points are called by a

variety of names and widely reported throughout the MSA at sites including

Kalambo Falls (Clark and Kleindienst 2001) and Twin Rivers (Clark and

84

Brown 2001) in Zambia, Klasies River (Singer and Wymer 1982;

Thackeray 1989) and Florisbad (Kuman et al 1999) in South Africa, and

Aduma in Ethiopia (Yellen et al 2005).

Levallois points (Inizan et al 1992; Clark and Kleindienst 2001) are convergent flakes produced deliberately from specially prepared Levallois cores (Inizan et al 1992:55). Striking platforms are located at the base of the point and remnant scars from core preparation prior to detachment create a trademark triangular basal facet and guiding arris on the dorsal side of the proximal end of the point. Points made using Levallois technology are a common feature of many MSA assemblages throughout Africa (see

Garcea 2004; Pleurdeau 2005; Rose 2004; Shea 2008; Tryon 2006; Tryon el al 2005; Van Peer 1991, 1998; Yellen et al 2005; Wurz 2002). There are two varieties of Levallois point at Mvumu: simple-based and tanged.

Simple-based Levallois points are roughly triangular with their maximum width occurring at the base which is formed by a facetted striking platform.

Tanged Levallois points have a pentagonal planform where the base is narrowed laterally to create a tang or stem. The tang feature could have been created as part of the pre-detachment preparation of the point or after the point was removed from the core using a truncation technique similar to that used in awl production.

85

Corner-struck points (Brooks et al 2006; Phillipson 2007a) have a roughly

kite-shaped planform with the tip of the point being formed by the

intersection of the two long sides and the base by the short sides. The name

“corner-struck” comes from the orientation of the striking platform which is

located on one of the short sides of the “kite”. These points are produced

from radially reduced cores and take advantage of arrises and flake scars

from previous removals to control the morphology of the points (Brooks et

al 2006). Limited unifacial retouch is often applied to the dorsal surface of

the blanks for these points to achieve the desired final morphology. MSA

points similar to these are generally restricted to central southern Africa and

have been reported from Bambata Cave, Zimbabwe (Armstrong 1931) and

≠Gi (Brooks et al 2006), Rhino Cave, and White Paintings Rock Shelter,

Botswana (Phillipson 2007a, 2007b).

4.3.3.4 – Other tools

In addition to the major tool categories of scrapers, awls and points, a number of uncommon and/or expedient tool types are also recognized at Mvumu. These techno- types are defined as follows:

Pointed pieces (this study) represent a morphologically and technologically

diverse collection of artifacts that possess a pointed working end suitable for

boring, drilling, and perforating activities but do not meet the techno-

typological criteria to be considered points or awls. They are typically

86

made on flake blanks and working ends may be formed by retouch, truncation, or unmodified convergent edges. Rare examples on core fragments do also exist. With a few exceptions pointed pieces are too thick and asymmetrical in cross section to have functioned as projectile points

(Shea 2006). Similar rough pointed tools are reported from MSA sites throughout Africa under various names including pointed tools (Clark and

Kleindienst 2001), borers (Singer and Wymer 1982), and pointed pieces

(Yellen et al 2005).

Notches (Shea 2008) have a working edge characterized by a deep, marked concavity created through the removal of one or more flakes. This concavity is deeper and narrower (~1cm deep by ~1cm wide) than that seen on concave scrapers and is more akin to the morphology seen in LSA spokeshaves (see Wadley 1992; Klatzow 1994). All notches at Mvumu are made on flake blanks. Notched areas on all notches reported here are consistent in shape and size.

Pebble tools (Mercader and Brooks 2001; definition expanded in this study) resembling small choppers are also present in this assemblage. Made on rounded cobbles, these tools are modified unifacially or bifacially along one edge with a chopping edge as defined by Clark and Kleindienst (2001) in

87

their Kalambo Falls typology. The surface away from the retouched edge remains cortical.

Crescents, sometimes referred to elsewhere as lunates (Clark and

Kleindienst 2001), are small, half-moon shaped retouched flake fragments.

The convex portion of the planview is created by steep retouch or “backing” while the straight edge remains sharp. Artifacts similar to these have been reported at other nearby sites such as Ngalue, Mozambique (Mercader et al

2009a) as well as Zambian sites such as Twin Rivers Kopje (Barham 2000;

Clark and Brown 2001), Mumbwa (Barham 2000), and Kalambo Falls

(Clark and Kleindienst 2001), and at Lake Rukwa in Tanzania (Willoughby

2001, 2002).

Snapped retouched pieces (this study) are made on medial flake fragments that are created by snapping or truncating a wide, thin flake laterally into wedge shaped sections. The wider lateral side is retouched with steep, scraper-like retouch. Similar tools have also been observed by the author at other Niassan sites including Mikuyu and Ngalue.

Retouched pieces follow Clark and Kleindienst‟s (2001) modified pieces.

They are flakes or flake fragments that show clear signs of being intentionally retouched that do not fit into any other techno-type. Most

88

often this category is in reference to modified flakes, but there are a few

examples of modified core fragments also included here. Such expedient

tools are present in most assemblages where they are referred to by various

names including “other retouched flake” (Shea 2008) or “flaked pieces”

(Clark and Kleindienst 2001).

Hammerstones (Clark and Kleindienst 2001; Shea 2008) are stones of a

size easily held in a hand that are distinguished from other rounded cobbles

that were never used in the lithic reduction process by the presence of

discrete, localized patches of battering damage.

4.4 – Metric and qualitative data

Once classified, the assemblage was analyzed from a primarily quantitative perspective. In building the quantitative database a range of metric measurements and qualitative observations were made for each artifact. Metric attributes and the specific methodologies used to take measurements varied slightly between cores, debitage, and tools. They are presented separately below. Recorded qualitative traits also varied by techno-type and are detailed in the relevant sections below.

Measurements pertaining to physical dimensions were taken with a digital sliding calliper (Mastercraft, #58-6800-4) and recorded in millimetres to two decimal places.

Mass was measured on a digital balance (Ohaus Navigator, Model N1D110) and

89

recorded to one decimal place. Angles were measured with a contact goniometer

(Ward‟s Natural Science, Model 13-0561) with accuracy to whole degrees.

4.4.1 Core metrics and qualitative traits

Metric attributes recorded for each core in the Mvumu study assemblage were adapted from the methods used by Shea (2008). Measurements include:

 Length – The core‟s longest dimension (Shea 2008).

 Width – The core‟s longest dimension measured perpendicular to length (Shea

2008).

 Thickness – The longest dimension measured perpendicular to the plane created

by length and width (Shea 2008).

 Mass – The core‟s mass in grams.

 Platform angle – The residual platform angle at the proximal end of a flake scar.

On multiple platform cores this measurement was taken for all platforms. On

discoidal cores the platform angle is the angle between the two flaking volumes.

If a platform was deemed to be damaged to an extent that a platform angle

measurement would not accurately reflect the original residual platform it was not

measured.

90

Morphological and qualitative attributes recorded for each core in the Mvumu study assemblage include:

 Number of striking platforms – The number of residual striking platforms

remaining on a core. Discoidal cores are considered to have two platforms – one

for each volume – that run around the core‟s circumference.

 Number of negative scars – The number of flake scars present on a core.

Accurately counting the flake scars on many cores was difficult due to the nature

of the raw materials used. Therefore, negative scar counts should be considered a

minimum number of removals.

 Cortex – Cortex was recorded on a presence/absence basis.

 Raw material type – The type of rock used in making the core.

 Weathering – The state of weathering of a core was recorded as fresh, slightly

weathered, or heavily weathered. A core was considered to be in fresh condition

if it retained sharp edges (similar to those on recently fractured stone) and showed

no signs of smoothing on its surface. Slightly weathered refers to cores that show

minor signs of edge rounding and smoothing. Heavily weathered cores are those

that have pronounced rounding on all edges and a smooth, polished surface.

91

4.4.2 – Debitage metrics and qualitative traits

Metric measurements taken for debitage were derived from Andrefsky (2005).

They include:

 Length – The distance from the proximal to distal end of the flake measured

perpendicular to the striking platform. In the case of angular waste length is

simply the maximum dimension of a piece in any plane (Andrefsky 2005).

 Width – The maximum dimension measured perpendicular to length (Andrefsky

2005).

 Thickness – The maximum dimension measured perpendicular to the plane

created by length and width (Andrefsky 2005).

 Mass – The mass of the debitage.

 Platform width – Maximum dimension of the striking platform measured along

the lateral axis of a flake (Andrefsky 2005).

 Platform thickness – Maximum dimension of the striking platform measured

along the dorsal-ventral axis (Andrefsky 2005).

 Platform angle – The exterior platform between the dorsal surface and the striking

platform in degrees (Patterson 1983). When this measurement was not possible

the interior platform angle (between the ventral surface and striking platform) was

recorded. Exterior angles were used whenever possible in this analysis. When an

92

interior angle must be used it will be noted. Because exterior and interior

platform angles are not necessarily related, only data from external angles are

presented in the results section.

Morphological and qualitative attributes recorded for each flake in the Mvumu study assemblage include:

 Platform preparation – The number of facets present on a striking platform

(Andrefsky 2005).

 Platform cortex – Residual cortex on a striking platform. Recorded as absent,

partial, or complete (Andrefsky 2005).

 Dorsal scar pattern – The patterning of the negative flake scars on the dorsal

surface of a flake. Recorded as radial, parallel, or irregular (Andrefsky 2005).

 Dorsal cortex – Residual cortex on a dorsal surface of a flake. Recorded as

absent, partial, or complete (Andrefsky 2005).

 Termination – The morphology of termination on the distal end of a flake.

Termination types include: feather (Andrefsky 2005:20), hinge (Andrefsky

2005:20), step (Andrefsky 2005:20), overshot (Andrefsky 2005:20), and concave

step (Isaac and Isaac 1977: Figure 57).

 Raw material type – The type of rock.

93

Only minimal metric and morphological data was collected for angular waste including length, mass, presence/absence of cortex, and raw material.

4.4.3 – Tool metrics and qualitative traits

Tool measurement methodologies varied depending on the type of blank the tool was produced on; tools made on core fragments essentially follow the methodology used for cores and flake blank tools follow debitage methods. Measurements taken include:

 Length – For tools made on flake blanks, length is the distance from the proximal

to distal end of the flake blank measured perpendicular to the striking platform

(adapted from Andrefsky 2005). Length on core fragment based tools is simply

the tool‟s longest dimension (adapted from Shea 2008).

 Width – The width of both flake blank and core fragment blank tools is the

maximum dimension measured perpendicular to length (adapted from Andrefsky

2005; Shea 2008).

 Thickness – Taken in the same way for both flake and core fragment blank tools,

thickness is the maximum dimension measured perpendicular to the plane created

by length and width (adapted from Andrefsky 2005; Shea 2008).

 Mass – The mass of the tool.

Morphological and qualitative attributes recorded for each tool include:

94

 Blank type – Whether the tool is made on a flake blank or core fragment.

 Retouch – Whether the tool is retouched or simply takes advantage of natural

edges.

 Edge angle – Measured on scrapers, edge angle is the angle of the working edge.

Measured using a contact goniometer.

 Platform orientation on corner-struck points – The side to which the platform is

oriented when the point is viewed from the dorsal side with the tip pointing away.

 Cortex – Cortex was recorded on a presence/absence basis.

 Raw material type – The type of rock used in making the tool.

 Weathering – The state of weathering of a tool was recorded as fresh, slightly

weathered, or heavily weathered. A tool was considered to be in fresh condition

if it retained sharp edges (similar to those on recently fractured stone) and showed

no signs of smoothing on its surface. Slightly weathered refers to tools that show

minor signs of edge rounding and smoothing. Heavily weathered tools are those

that have pronounced rounding on all edges and a smooth, polished surface.

4.5 – Experimental reduction

Small scale experimental flintknapping trials were conducted in the University of

Calgary Tropical Archaeology Laboratory during the spring of 2010. This work served

95

three purposes: 1) to gain a better understanding of quartz technology in Niassa, 2) to assess the theoretical compositional fidelity of the archaeological assemblage at Mvumu, and 3) to produce replica tools for use in a lithic usewear feasibility study that was conducted by PAC colleague Steven Simpson (Simpson et al 2010). Lithic raw materials for these trials were collected from the shore of Lake Niassa near Mvumu by Julio

Mercader during PAC‟s 2009 field season.

Trials began with a rounded quartz lake cobble (Figure 4.3) which was split into several pieces through bipolar percussion. Reduction of these pieces continued primarily through hard hammer percussion with a quartzite hammerstone (Figure 4.3). Soft hammer percussion with an elk antler billet was also used to a lesser extent to test the effectiveness of soft hammers on hard quartz and apply fine retouch to tool edges.

Reduction began with an opportunistic approach akin to that used in simple core technology aimed at relatively large flakes to serve as blanks for replica scrapers and awls. Once scraper and awl blanks were produced, reduction methods shifted to a discoidal approach with the goal of replicating corner-struck points. Cores were reduced until they became too small to be effectively reduced by free hand percussion. In addition to tool blanks a large number of waste flakes and shatter were also produced.

These pieces were collected and classified by size: <10mm, 10-20mm, and >20mm. The number of pieces and mass in each size category was quantified and these data were used to create a theoretical baseline of what should be expected from a complete quartz debitage assemblage. Scraper blanks were finished with soft and hard hammer retouch to create convex scraping edges. Awls were created using a truncation technique where the

96

blank was placed on a quartzite anvil and struck at ~90˚ angle (Figure 4.3). Corner- struck point replicas were not modified after being detached from the core.

______Figure 4.3 – Experimental Knapping. A) Lake Cobble, B) Hard Hammer Percussion, C) Truncation

97

Chapter Five: Middle Stone Age Lithics from Mvumu

5.1 – The Mvumu Lithic Study Assemblage

The Mvumu study assemblage contains 1977 lithics including 128 cores (6.47%),

214 tools (10.82%), and 1635 pieces of debitage (82.70%) (Table 5.1). Lithics were found throughout the entire depth of the excavation (~1m) with the highest concentration being found in an undulating horizon of variable thickness that occurs between ~25-35cm below the modern surface in units A7 and A8 and ~32-47cm in units B7 and B8 (Figure

3.6 and 3.7). A secondary lighter concentration is also present in the A units at ~60cm, but there is no correlate in the B units. Considered together, these units contained an average of 494 lithics per m2.

On the whole, this assemblage is in good condition with very little evidence of taphonomic alteration. Weathering was assessed on cores and tools and found to be minimal with none showing signs of heavy weathering, 16 (4.62%) showing slight weathering, and 330 (95.37%) being in fresh condition. Cores and tools that did show signs of weathering were all recovered from the upper half of the excavation with two coming from spit one, three from spit two, ten from spit three, and one from spit four. As

Table 5.1 – Assemblage Composition ______Category n= % Cores 132 6.67 Debitage 1631 82.49 Tools 214 10.82 Total 1977 100 ______

98

a whole assemblage, pieces with patches of residual cortex total 17.15% (n=339). By general typology cortex is present on 62.88% (n=83) of cores, 14.59% (n=238) of debitage, and 8.41% (n=18) of tools.

The remainder of this chapter presents detailed results of the raw material classification (chapter 5.2), artifact techno-typological classification and analysis (chapter

5.3), and experimental reduction conducted to test and support the analysis of the archaeological sample (chapter 5.4).

5.2 – Raw Materials

The lithic assemblage at Mvumu is heavily dominated by quartz (n=1884,

95.27%; Table 5.2). Small quantities of quartzite (n=87, 4.40%; light olive brown,

Munsell 2.5Y 5/6), rhyolite (n=4, 0.20%; yellowish brown/weak red, Munsell 10YR 5/4,

2.5YR 4/2), and (n=2, 0.10%; white, Munsell 2.5YR 8/1) are also present in the assemblage. All lithic material identifications were made through naked eye observation and consultation with the published geological cartography of Mozambique

(Lächelt 2004). This is consistent with the assemblages known from other sites in the

Niassa area (Mercader et al 2008; Mercader et al 2009a; Bennett et al 2010).

While quartz is the most common raw material in this collection, it must be noted that there is considerable variability within the category both in texture and colour.

Milky quartz is most abundant, but a small amount of quartz rock crystal is also present.

Milky quartz is divided into three subtypes based on texture: fine, coarse, and very coarse. Fine milky quartz is the overall most common raw material found at Mvumu. It

99

Table 5.2 – Raw Material Types ______Full New Name n = % Munsell Code Munsell Colour Fine Milky Quartz 1444 73.03 10YR 7/1 Light Grey Very Course Milky 195 9.86 2.5Y 8/1 White Quartz Course Milky Quartz 220 11.12 10YR 8/1; 10R 4/4; White; Weak Red; Dark Yellowish 10YR 4/4 Brown Quartzite 87 4.40 2.05Y 5/6 Light Olive Brown Quartz Rock Crystal 25 1.26 n/a n/a Rhyolite 4 0.20 10YR 5/4; 2.5YR 4/2 Yellowish Brown; Weak Red Chalcedony 2 0.10 2.5Y 8/1 White ______

represents 73.03% of the total assemblage (n=1444). Typical examples of this quartz are light grey in colour (Munsell 10YR 7/1). Coarse milky quartz is the second most common variety making up 11.12% (n=220) of the assemblage. This quartz tends to be white in colour (Munsell 2.5Y 8/1) and mottled with clear patches. The third most common variety is very coarse milky quartz which comprises 9.86% (n=195) of the assemblage. Very coarse quartz shows greatest variability in colour ranging from white

(Munsell 10YR 8/1), to weak red (Munsell 10R 4/4), to dark yellowish brown (10YR

4/4). Quartz rock crystal is the fifth most common material overall accounting for 1.26%

(n=25) of the total assemblage. This quartz is clear and colourless with the exception of scarce inclusions.

A total of 60 quartz artifacts (all cores) had sufficient patches of residual cortex to allow the origin of the raw material to be determined. Of these pieces, 38 were made on angular, rough surfaced quartz derived from local pegmatitic outcrops. The remaining 22

100

display smooth cortex representing either lake cobbles collected from the shore of Lake

Niassa, hillside deposits associated with dismantled Phanerozoic Karoo conglomerates, or ancient alluvial deposits.

5.3 – Artifact analysis results

The results of the lithic classification are divided into three sections: cores, debitage, and tools. Each section begins with a summary for the general category of artifact then, presents detailed results for each techno-type recognized at Mvumu. The complete data sets for all artifacts are included in Appendices A – Cores, B – Debitage, and C – Tools.

5.3.1 – Cores and core reduction

The Mvumu study assemblage includes 128 cores (6.47% of the total assemblage) that fall into eight techno-typological categories (Table 5.3). As with all categories of lithic artifacts from Mvumu, cores were primarily made on fine milky quartz (76.56%, n=98). The remaining cores were made on coarse milky quartz (8.59%, n=11), quartzite

(8.59%, n=11), and very coarse milky quartz (6.25%, n=8). Cores from Mvumu come in a wide range of sizes. They range in maximum dimension from 116.21mm to 26.26mm with a mean length of 49.64mm and a standard deviation of 14.46mm. Core mass ranges from 545.7g to 7.0g with an average value of 48.55g and a standard deviation of 67.06g.

The maximum dimensions and masses of cores from Mvumu are relatively continuously distributed from the lowest to highest value. Patches of residual cortex are common

101

Table 5.3 – Core Types ______Type n = % Simple Core 61 47.66 Discoidal 19 14.84 Preferential Discoidal - Recurrent 9 7.03 Levallois 8 6.25 Elongated High Backed Core 7 5.47 Low Backed Unifacial Radial 6 4.69 Preferential Discoidal - Single 6 4.69 Elongated High Backed Core on Flake 5 3.91 Testing Core 4 3.13 Bipolar Core 3 2.34 Total 128 100.00 Type n = % ______

being present on 63.28% (n=81) of cores. Most cores (98.44%, n=126) show no signs of weathering with the remainder (1.56%, n=2) appearing slightly weathered.

Simple reduction techniques outnumber prepared core technologies at Mvumu 80

(62.50%) to 48 (36.36%). The predominance of simple reduction strategies over prepared core technology is largely due to one techno-type: simple cores. Simple cores

(n=61, 47.66% of the total core assemblage; Figure 5.1) are the most common core techno-type at Mvumu accounting for nearly half of all cores. They show the greatest metric variability amongst the Mvumu cores lengths ranging from 116.21mm to

26.26mm (mean 50.27mm) and masses from 545.7g to 8.0g (mean 56.84g). Residual platform angles on simple cores range from 116˚ to 53˚ with an average of 82.09˚. Out of

141 residual platform angles measured on simple cores 36 (25.53%) had angles greater

102

than 90˚. Some simple cores are quite extensively reduced with up to four individual striking platforms (minimum 1, mean 2.38) and bearing as many as 10 flake scars

(minimum 1, mean 4.48). Despite this, the majority retain patches of cortex (75.86%, n=44). Fine milky quartz is the most common raw material utilized for simple cores

(79.31%, n=46) with coarse milky quartz (13.79%, n=8), quartzite (5.17%, n=3), and very coarse milky quartz (1.72%, n=1) making up the remainder of the specimens.

Discoidal cores (n=19, 14.84% of the total core assemblage; Figure 5.2) are the second most frequently recovered core techno-type. The largest discoidal core in terms of maximum dimension measured 76.74mm with the smallest measuring 32.76mm (mean

46.96mm). Their masses range from 152.6g to only 9.5g with an average of 46.56g. The average angle between the flaking surfaces, and therefore the effective platform angle, is

72.30˚ with a range of 50˚ to 104˚. Typical discoidal cores have nine flakes struck from these platforms with the highest number of residual flake scars being 13 and the lowest being six. The most common raw material exploited to produce discoidal cores is fine milky quartz (78.95%, n=15), followed by very coarse milky quartz (10.53%, n=2), coarse milky quartz (5.26%, n=1), and quartzite (5.26%, n=1). More than half have patches of cortex (52.63%, n=10) remaining and 15 of the 19 examples display side facets (78.95%).

The next best represented techno-type is elongated high-backed cores (n=12,

9.38% of the total core assemblage; Figure 5.1). Elongated high-backed cores are further subdivided into two categories: elongated high-backed cores and elongated high-backed cores on flake. The standard variety of these cores (n=7, 5.47% of the total core

103

Figure 5.1 – Simple Cores: A) Simple Core, B) Elongated High Backed Core, C) Simple Core, D) Simple Core

104

Figure 5.2 – Prepared Cores: A) Levallois, B) Preferential Discoidal Recurrent, C) Preferential Discoidal Single, D) Discoidal, E) Unifacial Radial

105

assemblage) range from 71.99mm to 34.38mm (mean 56.60mm) in maximum dimension and 33.3g to 9.3g (mean 27.04g) in mass. The average example has 6 flake scars (range

4 to 11) struck from platforms ranging in angle from 99˚ to 53˚ (mean 80.36˚). Fine milky quartz is the most common raw material (n=5, 71.43%) with the remaining cores being on quartzite (n=2, 28.57%). Only one elongated high-backed core has cortex remaining. Examples made on flakes (n=5, 3.91% of the total core assemblage) possess similar metrics with maximum dimensions between 62.14mm and 36.67mm (mean

50.89) and masses from 40.2g to 10.0g (mean 26.54g). Platform angles measure from

102˚ to 58˚ (mean 75.31˚) and on average four flake scars remain on each core

(maximum nine, minimum three). All examples of elongated high-backed cores on flakes are on fine milky quartz and two of the five have patches of cortex remaining.

Fourth most prominent are recurrent preferential discoidal cores (n=9, 7.03% of the total core assemblage; Figure 5.2). As with the non-preferential discoidal cores reported above, some of these cores also possess side facets (33.33%, n=3) on or near the lenticular margin. Typical recurrent preferential discoidal cores have a mean maximum dimension of 55.39mm (maximum 78.19mm, minimum 31.07mm) and a mean mass of

57.9g (maximum 147.9g, minimum 8.3g). Preferentially flaked volumes have an average of three flake scars (maximum 6, minimum 2) with non-preferential sides averaging four

(maximum 7, minimum 3). The angle between the two volumes ranges from 55˚ to 80˚ with an average of 56.44˚. Fine milky quartz is again the most commonly exploited raw material (66.67%, n=6) with coarse milky quartz (22.2%, n=2) and very coarse milky

106

quartz (11.11%, n=1) making up the remainder of the assemblage. Three cores (33.33%) retain patches of cortex.

Levallois cores (n=8, 6.25% of the total core assemblage; Figure 5.2) are the fifth most frequent. Mvumu‟s Levallois cores range from 62.71mm to 34.78mm (mean

45.22mm) in maximum dimension and 49.3g to 9.3g (mean 23.0g) in mass. All but one of the Levallois cores have only a single preferential flake removed from the flaking volume. The exception has two removals. Basal volumes have an average of five flake scars (maximum 9, minimum 2) associated with the preparation of the flaking surface and striking platform. Residual striking platform angles range from 121˚ to 59˚ with a typical platform measuring 86.29˚. Six examples are made on fine milky quartz (75.00%), one

(12.5%) is on coarse milky quartz, and one (12.5%) on quartzite. Six of the eight

Levallois cores retain some cortex on the base, but coverage tends to be minimal.

Unifacial radial cores (n=6, 4.69% of the total core assemblage; Figure 5.2) are the next most abundant techno-type at Mvumu. Maximum dimensions of these cores fall between 70.08mm and 47.26mm (mean 57.98mm) with masses ranging from 84.1g to

23.4g (mean 57.15g). Platforms angles range from 92˚ to 62˚ (mean 76.1˚) and produced between three to five residual negative flake scars per core. Raw material choice is split evenly between fine milky quartz and quartzite. The bases of five (83.33%) of these cores remain completely cortical. The non-cortical base of the single exception to this rule seems related to where the core was broken off of a larger cobble prior to being used as a core rather than intentional reduction.

107

Single preferential discoidal cores are also represented by six examples (4.69% of the total core assemblage; Figure 5.2) in the Mvumu study assemblage. Metrically, they are very similar to the other discoidal types reported in this study with maximum dimensions ranging from 78.83mm to 29.24mm (mean 48.22mm) and masses between

174.4g and 8.3g (mean 47.42g). While preferentially flaked sides by definition have one preferential flake, residual scars from previous rounds of removals can be seen on half of the specimens examined. Including these residual scars from previous stages of reduction preferential volumes possess between one and three removals. Non- preferential volumes range from three to six flake scars with an average of four. Platform angles for the preferential flakes show significant variation ranging from 37˚ to 92˚ with an average of 74.83˚. Most cores of this type were made on fine milky quartz (n=4,

66.67%) with the rest being made on very coarse milky quartz (n=2, 33.33%). Only one core has any cortex remaining.

Four “testing” cores (3.13% of the total core assemblage) were identified in the

Mvumu study assemblage. These cores have a maximum dimension ranging from 73mm to 36mm with a mean of 49mm. Their masses range from 237g to 24g with an average value of 83g. Showing only minimal reduction, they are limited to either two or three removals and all examples retain patches of cortex. Residual platform angles range from

120˚ to 65˚ (mean 87˚). All “testing” cores are made on fine milky quartz.

Bipolar cores are the least represented core techno-type at Mvumu with only three examples being recognized (2.34% of the total core assemblage). Metric analysis shows the bipolar cores are the smallest and least variable techno-type in this assemblage.

108

Maximum dimensions range from 39mm to 30mm (mean 34mm). The greatest mass of a bipolar core is 9g with the minimum being 7g (mean 8g). Striking platforms measure between 107˚ and 77˚ (mean 91˚). Negative scars are consistent at two on all bipolar cores with one initiated at the striking platform and the second from the end opposite the platform where the core sat on the anvil. Two examples of this techno-type were made on fine milky quartz (67%) and one on very coarse quartz (33%). Cortical patches remain on two (67%) cores.

5.3.2 – Debitage

Debitage totals 82.70% (n=1635) of the study assemblage. It is dominated by fine milky quartz (71.99%, n=1177) with the next most frequent raw materials being coarse milky quartz (11.19%, n=183) and very coarse milky quartz (10.83%, n=177). As such quartz makes up 94.01% (n=1537) of the debitage presented here. Remaining pieces are on quartzite (4.22%, n=69), quartz rock crystal (1.53%, n=25), rhyolite (0.12%, n=2), and chalcedony (0.12%, n=2). In terms of overall size distribution, 11.56% (n=189) of the debitage measures <10mm in maximum dimension, 48.32% (n=790) falls between 11-

20mm, and 40.12% (n=656) is >20mm. Size fractions were also evaluated by the portion of the total mass of debitage they represent. These observations showed 1.49% (76.1g) of the debitage belonging to the <10mm fraction, 20.74% (1062.8g) represented by pieces measuring between 11-20mm, and 77.77% (3984.6g) being from pieces >20mm.

The majority of the debitage is in the form of flakes and flake fragments (Table

5.4). Whole flakes account for 44.10% (n=721) of debitage. A typical whole flake is

109

Table 5.4 – Debitage Types ______Type n = % Whole Flake 717 43.96 Blade-Like Flake 7 0.43 Proximal Flake Fragment 173 10.61 Medial Flake Fragment 108 6.62 Distal Flake Fragment 140 8.58 Proximal Truncated Flake 7 0.43 Distal Truncated Flake 4 0.25 Lateral Truncated Flake 6 0.37 Core Prep Tablet 4 0.25 Core Trimming Element 47 2.88 Hammerstone Fragment 1 0.06 Shatter 417 25.57 Total 1631 100.00 ______

20.84mm (range 67.91mm-6.60mm) in maximum dimension and 3.62g (range 60.8g-

<0.1g) in mass. A slight majority of whole flakes (56.15%, n=402) are longer than they are wide or “under-square”, but a significant portion (43.85%, n=314) are “over-square” with their widths exceeding their lengths. Blade-like flakes (n=7, 0.43% of total debitage assemblage) are necessarily longer than they are wide, and on average they are longer and heavier than regular whole flakes with a mean maximum dimension of

30.94mm (range 38.43-18.88mm) and an average mass of 3.67g (range 8.5-0.4g).

Broken and truncated flakes are the second most abundant type of debitage at Mvumu

(n=432, 26.42% of total debitage assemblage). As a general group, flake fragments and truncated flakes have a mean maximum dimension of 15.59mm (range 48.85-4.82mm) and a mean mass of 1.97g (range 26.1-0.1g). More specifically, proximal flake fragments (n=173, 10.58% of total debitage assemblage) have a mean maximum

110

dimension of 17.53mm (range 48.85-6.10mm) and a mean mass of 2.6g (range 20.0-

0.4g), medial flake fragments (n=108, 6.61% of total debitage assemblage) average

13.83mm (range 32.05-4.82mm) and 1.44g (8.4-0.1g), and distal flake fragments

(n=140, 8.56% of total debitage assemblage) typically measure 14.13mm (range 41.1-

4.94mm) in maximum dimension and have an average mass of 1.33g (range 8.0-.01g).

Clearly truncated flakes show similar measurements. Proximal truncations (n=7, 0.43% of total debitage assemblage) averaging 22.33mm (range 34.88-10.05mm) in maximum dimension and 7.31g (range 26.1-0.9g) in mass, distal truncations (n=4, 0.24% of total debitage assemblage) being a mean of 18.03mm (range 24.46-13.37mm) in length and

2.65g (range 4.3-1.3g), and typical lateral truncations (n=6, 0.37% of total debitage assemblage) measuring 25.05mm (range 33.65-14.698mm) and 2.82g (range 5.9-1.4g).

Striking platforms on intact flakes and proximal fragments from Mvumu range in width from 42.45mm to 2.30mm (mean 12.36mm) and from 17.11mm to 0.97mm (mean

5.05mm) in thickness. Cortex is rare on striking platforms with 95.38% (n=908) being cortex free, 1.68% (n=16) retaining partial cortical cover, and 2.94% (n=28) being completely cortical. Beyond cortex removal, intentional platform preparation is rare;

85.70% (n=815) show no preparation, 11.88% (n=113) possess two preparation facets,

2.31% (n=22) are three faceted, and a single platform (0.11%) has four facets.

Cortex on the dorsal surfaces of flakes and flake fragments is minimal. It is completely absent from 86.13% (n=1000) of the examples studied here with 11.02%

(n=128) showing partial coverage and 2.84% (n=33) having fully cortical dorsal surfaces.

Broken down by size category the 1-10mm fraction of flakes, flake fragments, and

111

truncations represents 2.20% (n=2) of the cortical pieces, the 11-20mm fraction provides

11.57% (n=73), and the >20mm fraction contains 26.61% (n=116). Most dorsal surfaces show an irregular pattern of removals (95.51%, n=1128), 3.05% (n=36) show parallel removals, and 1.44% (n=17) retain signs of a radial removal pattern.

Flake terminations are also dominated by a single termination type: feather

(71.09%, n=654). Step terminations are the next most common at 10.65% (n=98) followed by concave step (sensu Isaac and Isaac 1977: Figure 57) at 10.22% (n=94), hinge (5.33%, n=49), and overshot (2.72%, n=25).

Core preparation byproducts are not overly abundant at Mvumu. Core trimming spalls make up 2.87% (n=47) of the debitage studied here. The typical core trimming spall has a maximum length of 21.05mm (range 48.77-12.46mm) and a mass of 2.53g

(range 11.0-0.4g). Three quarters (74.47%, n=35) of the artifacts of this type are completely cortex free while 23.40% (n=11) have patches of cortex remaining on their dorsal surfaces. A single specimen (2.13%) has a completely cortical dorsal face. Core rejuvenating spalls are even rarer than core trimming spalls totalling only four (0.24% of total debitage assemblage). They range from 88.18mm to 35.77mm (mean 57.08) in maximum dimension and from 99.9g to 9.2g (mean 45.3g) in mass. Most (75%, n=3) are devoid of cortex, but one example is completely cortical on its dorsal side.

The final type of debitage recorded at Mvumu is angular waste. It accounts for

25.50% (n=417) of the total debitage assemblage. Pieces of angular waste range from

67.62mm to 5.46mm (mean 20.45mm) in maximum dimension and from 40.9g to 0.1g

112

(mean 2.80g) in mass. The majority of angular waste retains no cortex on its surface

(85.13%, n=355), with the remainder showing partial coverage.

5.3.3 – Tools

Tools constitute 10.82% (n=214) of the Mvumu study assemblage (Table 5.5).

Fine milky quartz is again the most abundant raw material (79.44%, n=170) used in tool manufacture. Coarse milky quartz is the second most utilized (10.75%, n=23) and very coarse milky quartz comes third representing 4.67% (n=10). The remainder of tools are made on quartzite (4.20%, n=9) and rhyolite (0.93%, n=2). Tools in general range in maximum dimension from 85.57mm to 16.01mm with a mean length of 40.31mm and in mass from 429.7g to 0.9g with a mean of 22.01g. Maximum length values are evenly distributed throughout the range of variation. Masses are slightly skewed towards the lower end of the continuum with only 16 artifacts above 60g. The high end of the mass variation at 429.7g is the result of a single hammerstone that is 287.6g larger than the next largest artifact. Flakes and flake fragments are the most common blank type

(57.48%, n=123), but core fragment based tools are also well represented in the assemblage (42.52%, n=91). Retouch is common across the tool techno-types (87.38%, n=187), but it is generally unifacial and confined to small areas at the margins of tool edges. Patches of cortex remain on 8.41% (n=18) of the tools studied here. The majority of tools are in fresh condition (93.46%, n=200), with the remainder showing slight signs of weathering.

113

5.3.3.1 – Scrapers

Scrapers make up 53.74% (n=115) of the total number of tools and are divided into 12 techno-types (Table 5.5). Most scrapers were produced on flake blanks (64.34%, n=74) with the remainder on core fragments. However, it should be noted that despite flake blanks outnumbering core fragment blanks no subtype represented by greater than two examples is made exclusively on flakes or core fragments with the exception of core scrapers which are by definition made on core fragments.

The most abundant category of scraper is convex scrapers (n=41, 19.16% of total tool assemblage) which are further divided by retouch orientation into three subtypes: convex side scrapers (n=24, 11.21% of total tool assemblage), convex end scrapers

(n=17, 7.94% of the total tool assemblage), and convex side-and-end scrapers (n=2,

0.93% of the total tool assemblage; Figure 5.3). Convex side scrapers have a mean length of 44.15mm (range 65.79-25.90mm) and a mean mass of 20.12g (range 76.7-5.7g). The end scraper variety follows closely with a mean length of 45.41mm (range 62.20-

25.78mm) and an average mass of 21.76g (range 36.0-5.0g). Core fragment blanks slightly out number flake blanks in both varieties (side = 14:10, end = 10:7). Retouch is present on all convex side and convex end scrapers except for three of the side variant that exploit naturally occurring angles. Edge angles on convex side scrapers range from

80˚ to 33˚ (mean 61.8˚). Convex end scraper edges are similar with a range of 82˚ to 45˚

(mean 59.05˚). Fine milky quartz accounts for 66.67% (n=16) of side and 82.35% (n=14) of end subtypes. The balance of convex side scrapers are made on coarse milky quartz

(12.5%, n=3), very coarse milky quartz (8.3%, n=2), rhyolite (4.17%, n=1) and the

114

Table 5.5 – Tool Types ______Tool Type n= % Scrapers 115 53.74 Scraper, Bevel Based 24 11.21 Scraper, Concave 6 2.80 Scraper, Convergent 6 2.80 Scraper, Convex End 17 7.94 Scraper, Convex Side 24 11.21 Scraper, Convex-Concave 1 0.47 Scraper, Core 16 7.48 Scraper, Denticulate 3 1.40 Scraper, Denticulate Core 1 0.47 Scraper, Denticulate Semi-Circular 2 0.93 Scraper, End and Side 2 0.93 Scraper, Semi-Circular 10 4.67 Scraper/Awl 2 0.93 Scraper/Burin 1 0.47

Awls 31 14.49 Awl, Double Retouched 1 0.47 Awl, Double Truncation 8 3.74 Awl, Single Truncation 18 8.41 Awl, Topknot 3 1.40 Awl, Triple Truncation 1 0.47

Points 15 7.01 Point, Corner Struck 10 4.67 Point, Base-Struck 2 0.93 Point, Levallois-Like - Shouldered/Tanged 2 0.93 Point, Levallois-Like - Simple Base 1 0.47

Other 53 24.77 Crescent 2 0.93 Hammerstone 1 0.47 Notch 3 1.40 Pebble Tool 3 1.40 Pointed Piece 9 4.21 Pointed Piece, Distal Point Fragment 2 0.93

115

Table 5.5 – Tool Types (continued) ______Tool Type n= % Retouched Piece 23 10.75 Retouched/Prepared Snapped Piece 8 3.74 Split Pebble 2 0.93 Total n = 214 100

remaining convex end scrapers utilize quartzite (8.3%, n=2) and coarse milky quartz

(4.17%, n=1). Cortex remains on four and two of side and end types respectively. Five side (20.83%) and one end (4.17%) scraper show signs of slight weathering with the rest being in fresh condition. Side-and-end scrapers are represented by two (0.93% of the total tool assemblage) artifacts. These scrapers fit within the range of metric variation seen in the single edged side and end convex scrapers with maximum dimensions of

55.41mm and 36.28mm and masses of 49.4g and 15.2g. Both are on fine milky quartz flake blanks and have retouched edges with an average angle of 69.33˚. There are no signs of weathering or cortex on either piece.

Second most common at Mvumu are bevel-based scrapers (n=24, 11.21% of the total tool assemblage; Figure 5.3). The average bevel-based scraper is 45.97mm in maximum dimension with a range from 68.99mm to 27.01mm. Masses range from 91.7g to 5.7g (mean 26.98g). Core fragments outnumber flakes as blanks by 17 (70.83%) to seven (29.17%). All pieces are retouched to set the edge angle, create the bevelled base, or both. Edge angles range from 78˚ to 43˚ with an average edge measuring 59.2˚. The working edge is oriented as a side scraper in all but one case which shows an end scraper

116

Figure 5.3 – Scrapers: A) Convex Side on Flake, B) Convex Side on Flake, C) Semi- Circular, D) Denticulate, E) Bevel-Based on Flake, F) Bevel-Based on Core Fragment, G) Convex Side on Core, H) Denticulate Core, I) Convex End on Flake, J) Convex End on Core Fragment, K) Core

orientation. The surfaces of all but one (4.16%, slightly weathered) example are in fresh condition and only two (8.33%) possess remnants of cortex. Fine milky quartz is the

117

dominant raw material type used (79.17%, n=19) with coarse milky quartz (12.5%, n=3) and very coarse milky quartz (8.33%, n=2) making up the remainder.

Core scrapers represent 7.94% (n=17) of the tools from Mvumu (Figure 5.3).

Most examples are convex core scrapers (n=16, 94.12% of total tool assemblage), but a single example is considered a denticulate core scraper (0.47% of total tool assemblage). Edge angles on the convex edged examples average 72.41˚ (range 90˚ to

55˚) and the denticulate core scraper‟s working edge is more acute at an average of

70˚across its width. Maximum dimensions for core scrapers range from 81.22-37.58mm

(mean 56.36mm) for convex specimens. The denticulate specimen is also within this range measuring 66.18mm. Mass ranges from 140.6-17.7g (mean 60.82g) for convex edged and is 66.7g for the denticulate. All examples are made in fine milky quartz with cortex remaining on only four convex edged artifacts (25.00%). Three convex core scrapers are slightly weathered, but all others are in fresh condition.

Next most common are semi-circular scrapers (n=12, 5.61% of the total tool assemblage; Figure 5.3). As with core scrapers, this type is split between examples with convex (n=10) and denticulate (n=2) working edges. The convex variety ranges in maximum dimension from 44.76-29.46mm (mean 36.58mm) with the denticulates being slightly smaller measuring 35.68mm and 22.11mm (mean 28.90mm). Masses are also higher for the convex variety ranging from 19.6-7.0g and 9.6-6.5g (means 11.97g and

8.05g) respectively. All pieces are retouched. Convex edge angles average 58.3˚ (range

75˚ to 42˚) and denticulates average 46˚ (range 51˚ to 41˚). Convex semi-circular scrapers are made on core fragments (n=6) slightly more often than flakes. All

118

denticulate examples are made on flake blanks. Fine milky quartz is the most common raw material with six convex and one denticulate utilizing the variety. The other denticulate semi-circular scraper is made on quartzite as is one additional convex edged example. The final three convex pieces are on coarse milky quartz. No artifacts in this category retain cortex and weathering is minimal with only a single convex example showing slight weathering.

Six concave scrapers are present in the Mvumu study assemblage (2.80% of the total tool assemblage). Made on both flake blanks (n=2) and core fragments (n=4) concave scrapers have maximum dimensions ranging from 62.32-34.55mm (mean

43.30mm) and masses from 33.8-9.0g (mean 17.02g). The angle of their working edges ranges from 76˚ to 53˚ with an average of 65˚. All examples are made on fine milky quartz and are in fresh condition with one piece retaining patches of cortex.

Convergent scrapers account for an additional 2.80% (n=6) of the tools in this study. Blank types for convergent scrapers are evenly split between flakes and core fragments. They range in maximum dimension from 39.76-27.20mm (mean 32.03mm) and have masses from 15.3-5.4g with an average of 9.4g. Most pieces have two retouched edges to form the convergent morphology, but one example takes advantage of a naturally occurring angle along one of its edges. Edge angles range from 72˚ to 48˚ with a mean angle of 58˚. Again, fine milky quartz dominates the collection being the raw material used for four examples. The remaining two convergent scrapers are made on coarse milky quartz. All tools of this type are in fresh condition and none have cortex remaining on their surfaces.

119

Three examples of denticulate scrapers (1.40% of the total tool assemblage) are present in the Mvumu study assemblage (Figure 5.3). These tools have a maximum dimensions ranging between 58.97mm and 23.25mm (mean 37.53mm). Their masses span from 55.0g to 5.0g (mean 22.37g). Two are made on flake blanks with the third on a core fragment. All are made on fine milky quartz. Working edges, all retouched to achieve their denticulate morphology, have angles that are extremely consistent compared to other scrapers at Mvumu with an average angle of 63˚ (range 66˚ to 58˚).

None of these scrapers show signs of being weathered and all are devoid of cortex.

Three scraper/awls (1.40% of the total tool assemblage) are present in the

Mvumu study assemblage. Average scraper/awls have a maximum dimension of

62.31mm (range 76.04-46.92mm) and a mass of 50.37g (range 68.9-27.2g). Two are made on flake blanks with the third on a core fragment. Fine milky quartz and coarse milky quartz are both used as raw materials with the finer variety outnumbering coarse by

2:1. One piece retains patches of cortex. Signs of weathering are absent.

Lastly, a single concave-convex scraper (0.78% of the total tool assemblage) was observed. Made on a coarse milky quartz core fragment, it has a maximum dimension of 49.73mm and a mass of 28.5g. Its working edges are retouched to achieve the morphology with an average angle of 76.5˚. It possesses no cortex and shows no signs of weathering.

120

5.3.4.2 – Awls

Following scrapers, awls (n=31, 14.49% of the total tool assemblage) are the next most common tool category recorded at Mvumu. Awls are further divided into five subtypes (Table 5.5). The most abundant awl type is single truncation awls (n=18,

58.06% of the total tool assemblage; Figure 5.4). There is a preference for flake blanks amongst single truncation awls (n=12, 66.67%) with only six examples utilizing core fragments (33.33%). This is the only subtype within the awls to make use of core fragments as the basis of their manufacture. Typical examples have a mean maximum dimension of 43.19mm (range 65.36-23.27mm) and an average mass of 16.23g (range

42.9-3.0g). All are retouched with a single truncation at the working end. The primary

Figure 5.4 – Awls: A) - D) Single Truncation, E) - G) Double Truncation

121

raw material used is fine milky quartz (n=13) with four on coarse milky quartz and one on very coarse milky quartz. Residual patches of cortex are common (n=15, 83.33%).

All examples are in fresh condition.

The Mvumu study assemblage contains eight double truncation awls (3.74% of the total tool assemblage; Figure 5.4). This subtype averages 30.47mm in maximum dimension (range 40.75-21.73mm) and 7.26g (range 13.3-2.1g). All pieces are retouched with double truncations to shape the working end. A wide variety of raw material types are represented including fine milky quartz (n=5), coarse milky quartz (n=1), very coarse milky quartz (n=1), and quartzite (n=1). Cortex is absent from all double truncation awls and all are in fresh condition.

Topknot awls (n=3, 1.40% of the total tool assemblage) are the third most common awl type in this collection. These awls take advantage of the natural shape of the topknot flake blanks on which they are made to provide their morphology, but they all also possess minor retouch to further shape their working end. They average 27.55mm in maximum dimension (range 29.20-24.97mm) and 3.97g (range 6.4-2.7g) in mass. Cortex is absent from all examples. Raw materials include fine milky quartz (n=2) and very coarse milky quartz (n=1). All three topknot awls are in fresh condition.

The triple truncation awl (n=1, 0.48% of the total tool assemblage) is the final awl type based on truncation and is represented by a single example. This awl has a maximum dimension of 37.99mm and a mass of 14.3g. Retouch is present in the form of three truncations that form two working points. It is made on coarse milky quartz with no cortical patches and is in fresh condition.

122

The double retouched awl type is also represented by a single artifact (0.48% of the total tool assemblage). Made on a flake blank, the working end is formed by conventional retouch along the blank‟s lateral edges. It measures 23.76mm in maximum dimension and has a mass of 1.8g. This awl is made on fine milky quartz and its surface is cortex free. There are no signs of weathering.

5.3.3.2 – Points

Mvumu‟s point techno-types make up 7.01% of the tool assemblage (n=15) and fall into three flake-based categories (Table 5.5). Most prominent amongst these types are corner-struck points (n=10, 4.67% of the total tool assemblage; Figure 5.5).

Figure 5.5 – Points: A) - C) Corner-Struck, D) Base-Struck, E) - F) Levallois

123

Corner-struck points have an mean maximum dimension and mass of 32.17mm and

6.18g; however these numbers are skewed by a single piece that is considerably larger than all others (63.71mm and 24.6g). With this outlier removed, typical corner-struck points are 28.67mm (range 37.11-19.96mm) and 4.13g (range 5.7-1.2g). The distinct corner-struck platform that gives these points their name is oriented most often to the right (n=8) when viewed from the dorsal aspect. All examples are retouched and no cortex remains on the dorsal surfaces of the flake blanks or platforms of any of them. All artifacts of this type are made on fine milky quartz and all except one are in fresh condition.

Levallois points (n=3, 1.40% of the total tool assemblage; Figure 5.5) are the second most common points. Two of these points are of the tanged variety whilst the other has a plain base. Despite this technological difference, the metrics behind the points are all similar. Their maximum dimensions average 34.77mm (range 41.14-

29.38mm) and their mean mass is 5.57g (range 6.6-5.0g). As with corner-struck points, all examples of these types are made on fine milky quartz. Although these points are the product of Levallois prepared core technology which generally does not require retouch to achieve its final form, one piece does show minor retouch. No cortex remains on platforms or dorsal surfaces and all artifacts are in fresh condition. Base-struck points are the least common point type with only two examples (13.33% of the total tool assemblage; Figure 5.5). As with all other points, this type is made exclusively on fine milky quartz. They possess maximum lengths of 29.44mm and 25.92mm (mean 27.68)

124

with masses of 5.8g and 3.1g (mean 4.45g). The platforms on each point are prepared and bear two facets each. Cortex is absent and both pieces are in fresh condition.

5.3.4.4 – Other tools

The remainder of the Mvumu tool assemblage is composed of rare and/or informal tools (Table 5.6). Considered as a group these tools account for 24.77% (n=53) of the tools in this study. Most common amongst these tools are retouched pieces

(n=23, 10.75% of the total tool assemblage). Retouched flakes make up the majority of these tools (n=19, 82.61%), but four (17.39%) retouched core fragments/chunks are also present. These expedient tools average 30.30mm in maximum dimension (range 82.4-

17.18mm) and 8.19g in mass (range 62.8-2.1g). Made primarily on fine milky quartz

(n=18, 78.26%), no retouched pieces retain patches of cortex. Other raw materials represented include coarse milky quartz (n=2, 8.67%), very coarse milky quartz (n=2,

8.67%), and rhyolite (n=1, 4.35%). Twenty one (91.30%) examples are in fresh condition with two showing slight weathering (8.67%).

Pointed pieces (n=11, 5.14% of the total tool assemblage) are next most common in this category. Typical pointed pieces measure 32.01mm in maximum dimension

(range 42.70-22.35mm) and have an average mass of 6.06g (range 13.6-1.7g). Ten are made from flake blanks (90.90%) and one began as a core fragment (9.09%). Fine milky quartz is the most common raw material (n=9, 81.82%) with a single example on coarse milky quartz (9.09%) and one on quartzite (9.09%). Six of the 11 (54.55%) were formed with retouch while the others are pointed by virtue of their morphology when struck from

125

Figure 5.6 – Other Tools: A) and B) Crescents, C) Notch, D) - H) Snapped Retouched Pieces

the core. One pointed piece on a flake fragment retains a small patch of cortex and all are in fresh condition.

Snapped retouched pieces (n=8, 3.74% of the total tool assemblage Figure 5.6) are the most common microlithic-like tool type. All made on medial flake fragments, these tools range from 33.87mm to 16.01mm (mean 22.95mm) in length and 4.2g to0.9g

(mean 2.09g) in mass. Six are retouched with two having working edges defined by unmodified lateral edges of the blank. All but one snapped retouched piece, made on quartzite, are on fine milky quartz. None have any remaining cortex and all are in fresh condition.

Notches are represented by three (1.40% of the total tool assemblage Figure 5.6) artifacts. They average 46.17mm (range 52.58-38.30mm) and 11.07g (range 15.9-4.9g).

126

All notches were made on fine milky quartz flake blanks, have no residual cortex, and are in fresh condition. The notched portion of each tool is retouched.

Also with three examples are pebble tools (1.40% of the total tool assemblage).

These tools are made on minimally reduced fine milky quartz cobbles with an average length of 57.84mm (range 67.68-51.74mm) and mean mass of 105.97g (range 142.1-

65.8g). The surfaces of all three tools are mostly cortical and appear in fresh condition.

Two crescents (0.93% of the total tool assemblage; Figure 5.6) were found in the

Mvumu study assemblage. They have lengths of 29.51mm and 19.32mm and masses of

2.4g to 1.7g. One is made on a fine milky quartz flake fragment and the other on very coarse milky quartz. Both are in fresh condition with no cortex remaining. The crescent morphology was created on both examples by backing retouch.

Only one hammerstone (0.47% of the total tool assemblage) was identified at

Mvumu. This oblong cobble measures 85.57mm at its widest point and has a mass of

429.7g. It is made of fine milky quartz and was identified as a hammer by two distinct battered patches on its surface.

5.4 – Experimental reduction results

Experimental reduction trials began with a 1380g cobble collected from the shore of Lake Niassa. The raw material was consistent with the fine milky quartz described in the archaeological assemblage. Reduction of this cobble produced a total of 990 pieces of debitage. This assemblage included 495 (50.00%) pieces measuring <10mm, 363

(36.67%) pieces between 10-20mm, and 132 (13.33%) pieces being >20mm. Most of the

127

mass of the experimental assemblage was included in the >20mm category (1080g,

78.26%) with the 10-20mm category representing 14.49% (200g) of the mass and the remaining 7.25% (100g) being composed of pieces <10mm in maximum dimension.

Pieces consistent with all types of debitage were produced during this experiment.

Reduction by freehand percussion remained possible until the core was reduced to 45.7g

(Figure 5.7:B).

Once suitable blanks were produced, scrapers were manufactured by applying retouch to an edge. Comparable edges to those seen on convex scrapers in the study assemblage were produced with ease and with very little time investment (<2 minutes).

The replication of truncation awls required more time and effort than scrapers, but after a short period of experimentation close replicas to those present in the study assemblage

Figure 5.7 – Experimental Replica Tools: A) and B) Discoidal Cores, C) Chert Corner-Struck Point, and D) Quartz Corner-Struck Point

128

were created. The key variable in producing the desired truncation morphology was found to be the angle at which the blank was rested upon the anvil. The best results were achieved by resting the end of the blank that was to become the working end of the tool directly on the anvil with the opposing end supported above the anvil using a piece of folded leather. Attempts to reproduce corner-struck points only resulted in limited success. The main difficulty encountered was producing a flake that possessed the desired kite shaped morphology; flakes on quartz ended short and therefore lacked a proper point at the distal end (Figure 5.7:D). However, the base morphology that defines corner-struck points was successfully reproduced and the failure to achieve the full point morphology was likely related to the author‟s skill as a flintknapper and not to problems with the production technique. More success was had when replicating corner-struck points on chert (Figure 5.7:C).

129

Chapter Six: Discussion

The Mvumu lithic assemblage studied in this thesis has provided the largest and most detailed techno-typological assessment of lithic technology in northern

Mozambique. Previous work on lithics from Niassa includes the classification of 33 stone artifacts from Mikuyu (see Mercader et al 2008) and 555 artifacts from Ngalue (see

Mercader et al 2009a). In general terms it is consistent with these and other MSA industries found throughout south-central Africa, but it also contains several unique techno-typological elements that make it distinct. With a chronology set at ~28-33kya it also represents one of the youngest dated MSA assemblages that have been analyzed.

This chapter draws from the results of the analysis to characterize the assemblage and assess its techno-typology relative to other MSA lithic assemblages from southern Africa.

It begins with a discussion of the depositional context from which the lithics were excavated to evaluate the completeness of the cultural component at Mvumu. The focus then turns to the artifacts themselves to create a profile of lithic technological behaviour at Mvumu from core reduction to finished tool manufacture and consider the connotations of the lithic technology on other aspects of behaviour at Mvumu. And finally, it considers the Mvumu lithics in the context of current MSA research through comparisons with geographically and temporally related sites.

6.1 – Mvumu depositional context and site fidelity

As noted in chapter 3, the geological package containing the MSA at Mvumu can be interpreted as a stone line. Stone lines are often viewed as post depositional features

130

that may affect the integrity of the archaeological record (Mercader et al 2002). This is because they typically represent interruptions in sedimentation followed by deflation brought about by the loss of the smaller fractions in the deposit that results in a lag deposit. A variety of agents can be responsible for this loss of the finer fractions during stone line formation often including chemical or mechanical weathering combined with fluvial, alluvial, colluvial, and/or aeolian transport. Problems can arise with the fidelity of the archaeological record at stone line sites as a result of the winnowing effect of these geological processes that may remove archaeological materials as well as natural sediments. There is no geological evidence for fluvial or any other high energy transport agents at Mvumu (Gosse, personal communication). Instead, stone line formation at

Mvumu seems to be related to lower energy processes involving colluvial movement and slope wash. The unweathered condition of the assemblage suggests rapid burial of the artifacts as the package was created.

Although some winnowing is necessarily in the formation of a stone line, the compositional fidelity of the Mvumu assemblage is quite good. Size fraction data from the debitage component of the Mvumu assemblage was compared to a debitage assemblage (n=990) created during the production of experimental replicas of Mvumu- like tools (Table 6.1). Comparisons were made both in terms of the mass represented by the fraction and the number of pieces represented. The results showed that the mass of the <10mm fraction of the archaeological assemblage is under represented by a factor of

4.87. Likewise, it is numerically underrepresented by 4.33:1. Results from the 10-20mm fractions show a slight over representation in the archaeological collection by factors of

131

Table 6.1 - Size and mass distributions for the experimental reduction of quartz cobbles ______Experimental Assemblage Debitage Size Fraction Mass, g % Debitage, n % 1-10mm 100 7.24638 495 50.00 11-20mm 200 14.4928 363 36.67 >20mm 1080 78.2609 132 13.33 1380 990

Mvumu Assemblage Debitag e Size Fraction Mass, g % Debitage, n % 1-10mm 76.1 1.48531 189 11.56 11-20mm 1062.8 20.7436 790 48.32 >20mm 3984.6 77.7711 656 40.12 5123.5 1635 ______

1.43 in mass and 1.32 in numbers. Data for the <20mm pieces is conflicting with the mass of the archaeological debitage being slightly under represented by a factor of 1.01 and a numerical overrepresentation of 3.00:1. This comparison suggests that the Mvumu stone line represents a filtered palimpsest of final MSA lithic technology that has been compressed vertically and lost some of its smaller fraction while its larger fraction was slightly inflated. Nonetheless, all size categories remain well represented in the archaeological assemblage and it is considered to have retained good compositional fidelity capable of accurately informing us on the nature of Niassan lithic technology

~28-33kya. The energies involved in creating the stone line were incapable of removing any artifactual evidence larger than 10mm; the slight underrepresentation in the mass of the <20mm fraction is small enough to easily be accounted for in simple knapping

132

differences between the archaeological and experimental assemblages. The underrepresentation of the <10mm fraction does not compromise this analysis because these small pieces, mostly shatter and nondescript flake fragments, provide little information on the aspects of techno-typology that are the focus of this thesis. In short, the archaeological assemblage from Mvumu is sufficiently intact to allow it to be used to characterize the lithic technology of the final MSA in Niassa.

6.2 – Technological behaviour at Mvumu

6.2.1 – Raw Materials

The raw material utilized in the manufacture of any lithic assemblage is a major element of its technological character. In the case of Mvumu, the assemblage is heavily dominated by varieties of quartz. The overwhelming use of quartz is the most immediately apparent technological decision made by MSA knappers at the site, although small amounts of quartzite, rhyolite, and chalcedony are also present in the assemblage.

There is no recegnizable difference in the technological approach toward quartz and non- quartz raw materials.

Quartz is one of the most common forms of silica found on Earth (Dickson 1977) and is readily available in the Mvumu area (Afonso and Marques 1998:80; Lächelt

2004:352) suggesting that raw material procurement/quarrying could have been one of the key activities that drew people to Mvumu. Four potential sources were identified through survey. The first source is quartz associated with pegmatitic outcrops that occur throughout the hills surrounding Mvumu (Figure 6.1). These outcrops would have made

133

ideal quarry sites for the inhabitants at Mvumu because quartz can be easily gathered in large quantities from several locations within 1km of the site. Quartz in these outcrops ranges in size from large boulders (>1m) to small pebbles (<5cm). Pieces tend to have an angular morphology and rough cortex that identifies their origin. The second source of quartz is the shore of Lake Niassa. Palaeoclimatic evidence suggests that the level of

Lake Niassa, and therefore the proximity of its shoreline to Mvumu, has fluctuated greatly throughout the Pleistocene (Finney et al 1996; Cohen et al 2007; Scholz et al

Figure 6.1 – Quartz Coverage on the Slopes near Lake Niassa (Photograph J. Mercader)

134

2007). Lake core data indicates that lake levels during the Mvumu occupation ~28-33kya were 200-300m below current levels (Finney et al 1996). While this would have meant the shoreline was further away from Mvumu than its current 200m, it still would have been within sight and accessible due to the steep morphology of the lake basin. Beaches along the lake‟s shoreline provide a rich source of rounded quartz cobbles and pebbles ranging in size from >30cm to <5cm. The lacustrine source of these stones is reflected in the smooth cortex on their surfaces. The final two possible sources also contribute smooth, rounded pebbles associated with either a weak residual signal from dismantled

Phranerozoic Karoo conglomerates or ancient alluvial deposits that occur on the slopes throughout the area.

This wealth of quartz may seemingly make it an obvious choice for MSA toolmakers simply based on convenience. However, within much of the archaeological community there is a commonly held notion that quartz is a poor raw material for flaking that would only have been exploited when other “superior” stones were not available

(Dickson 1977; Kamminga 1982:24). At the very least it is typically not seen as a suitable raw material for the manufacture of anything but the crudest tools and the detailed analysis of quartz technology is often portrayed as an exercise in futility. This is largely due to the flaking characteristics possessed by quartz. Internal flaws in quartz cobbles and chunks are common due to heat and pressure experienced during their geological history (Dickson 1977). This may cause unpredictable fracturing during reduction of a quartz core. Even if such flaws are absent, quartz possesses relatively low tensile and compressive strength compared to other common lithic raw materials which

135

makes it brittle (Cotterell and Kamminga 1990; Tallavaara et al 2010) and it rarely possesses detectable conchoidal fracture characteristics (Dickson 1977; Knight 1991;

Seong 2004). This has led to quartz gaining a reputation for producing only nondescript, undiagnostic assemblages and has resulted in some researchers making only minimal comments on the quartz components of assemblages or even to refrain from trying to apply techno-typological classifications to quartz at all (Cornelissen 2003).

However, as with most stereotypes this is not necessarily the case as it does not consider variation in the qualities of different quartzes or the full complexity of quartz technology (Hiscock 1982). Like all raw materials, quartz presents a number of specific challenges to a flintknapper or a lithic analyst, but once the properties of the material are understood it can be reduced/analyzed with no greater difficulty than obsidian or chert.

Indeed, the results of the classification presented in this thesis also show that with such understanding quartz technology is not limited to “crude” tool types as the Mvumu study assemblage contains a full range of tool techno-types including types made using sophisticated prepared core technologies. One key property of quartz that must be assessed to exploit it to its potential is texture. The texture is determined by the size and arrangement of the crystals that make up a piece of milky quartz (Hiscock 1982; Seong

2004). Finer internal crystal structures in a piece of quartz produce better flaking characteristics (Seong 2004). Such attributes in quartz can be assessed through naked eye observation and tactile examination of a piece being considered for reduction (Seong

2004). Even should the initial evaluation of a cobble fail a knapper during raw material selection, the good or bad qualities of a given piece of quartz would quickly become

136

apparent after only a few attempted removals. The preferential selection of such finer textured quartzes, and their superior flaking characteristics, is clearly expressed at

Mvumu where 71.99% of the lithic artifacts were made on fine milky quartz; the finest textured variety present. The beneficial attributes of fine textured quartz were also confirmed through the experimental knapping conducted in conjunction with this study.

In light of these facts regarding the true potential of quartz as a raw material for the production of stone tools it should be noted that it is indeed a commonly chosen raw material not just at Mvumu but throughout south-central Africa. There is ample evidence to suggest it was even preferentially exploited in many MSA and LSA assemblages over other available materials that have “superior” flaking characteristics in the eyes of most archaeologists (see Clark and Brown 2001; Cornelissen 1997, 2003; Mercader and Martí

1999, 2003, Rose 2004; Williams 2005; Mercader et al 2008; Mercader et al 2009a).

Moreover, there is abundant evidence that quartz was used in the creation of many lithic assemblages that show equal diversity and technological sophistication to those utilizing non-quartz materials not only at Mvumu but throughout south-central Africa (e.g.

Mikuyu, Mozambique: Mercader et al 2008; Ngalue, Mozambique: Mercader et al 2009a;

Mumba, Tanzania: Mehlman 1987, Domínguez-Rodrigo et al 2007, Diez-Martín et al

2009; Mwanganda, Malawi: Clark and Haynes 1970; Twin Rivers, Zambia: Clark and

Brown 2001, Barham 2000; Mumbwa, Zambia: Barham 2000; see regional comparisons below). Therefore, quartz cannot be considered a raw material of last resort and there is no reason that a detailed analysis of the Mvumu assemblage is any less informative than a similar analysis on an assemblage dominated by any other raw material.

137

The reason behind the almost exclusive use of quartz at Mvumu is unclear. The most parsimonious reason for the overwhelming use of quartz is simply convenience.

The hillsides surrounding Mvumu contain an incredible wealth of easily accessible pegmatitic quartz that was exploited as the primary source of raw materials for this assemblage. Rounded quartz cobbles from riverine and lacustrine sources, as well as residual Karoo conglomerates, were also exploited as a secondary source of quartz. This convenience seemingly took precedence over the fact that other lithic materials which typically possess better flaking properties are also available in the Mvumu area.

Quartzite, although rarer than quartz, can be sourced from the same slopes that quartz was collected from and rhyolite and chalcedony are available within 10-15km of the site in the Lunho River basin (Bennett et al 2010). From a geological perspective, rhyolite may also be present in dismantled Karoo and alluvial deposits, but this has not been directly observed. Despite this close proximity, these “superior” materials represent <5% of the study assemblage. This trend toward the preferential exploitation of quartz is also apparent at other Niassa sites including Mikuyu (Mercader et al 2008) and Ngalue

(Mercader et al 2009a).

The abundance and convenience of quartz would surely have been an attractive feature of the area to MSA people in Niassa, but this local abundance does not help explain why quartz was preferentially utilized at many sites throughout central Africa during the MSA and into the LSA (e.g. Yellen 1996; Mercader and Martí 1999; Barham

2000, 2002a; Mercader and Brooks 2001; Mercader et al 2002; Cornelissen 2003;

Williams 2005; Dominguez-Rodrigo et al 2007; Diez-Martín et al 2009). Sadly, few

138

authors have discussed the potential reasons behind this phenomenon. One alternative explanation for the preferential selection of quartz is that the qualities of the stone offered some adaptive advantage over other available raw materials. Much of central Africa is dominated by woodland and forest environments (White 1983) so it could be reasonable to hypothesize a link between these types of ecosystems and quartz use. However, research conducted in northwestern (Cornelissen 2003) and Equatorial Guinea

(Mercader and Brooks 2001) has documented the preferential reduction of quartz in both forested and open environments. While a preference for quartz in open environments does not necessarily imply it is not ideally suited to forested environments, it does weaken the connection. The preference for quartz could also be culturally driven with no underlying ecological adaptive rationale, but this notion is difficult to effectively test given the available archaeological data.

6.2.2 – Lithic Reduction Strategies at Mvumu

The preference for quartz as a raw material is clearly visible in the cores from

Mvumu; 91.41% are made on quartz with quartzite being the only other raw material represented in the core component of the study assemblage (8.59%). Within the general category of quartz the tendency to favour fine textured stones is clear with 76.56% being produced on fine milky quartz. Despite of the challenges interpretations of quartz assemblages such as this can present, several clear and consistent reduction strategies are identifiable in the cores from Mvumu. Amongst these modes of reduction are several approaches that are consistent with prepared core strategies (as per Clark and Kleindienst

139

2001) seen in sites across Africa during the MSA. All prepared cores at Mvumu share one key technological feature: radial reduction. Radial reduction is a common feature in

MSA lithic assemblages throughout Africa, including those dominated by quartz (e.g.

Katanda – Yellen 1996, Williams 2005; Mikuyu – Mercader et al 2008); Mumba –

Domínguez-Rodrigo et al 2007, Diez-Martín et al 2009; Mumbwa – Barham 2000;

Mwanganda – Clark and Haynes 1970; Ngalue – Mercader et al 2009a; Twin Rivers –

Clark and Brown 2001, Barham 2000). It was employed for two purposes at Mvumu. In the case of discoidal and unifacial radial cores it was used to remove a continuous series of flakes from a platform running around the periphery of a core. In Levallois and preferential discoidal cores radial flaking was used in the shaping of the preferentially reduced volume of the core to create the desired convexity and therefore to control the cross section of the prepared product. Discoidal cores are the most common of prepared core types, with preferential discoidal types being second most common and Levallois making up the remainder. Prepared cores are generally small with an average mass of

50.8g. It is worth noting that this value is skewed by a few larger cores; 70.8% of prepared cores are under 50g (Appendix A).

While clear techno-typological differences do exist between the prepared core types defined in this thesis, it should be noted that these classifications are based solely on the technology as seen in the cores at the time when they were discarded and entered the archaeological record. It is arguable that the reality of prepared core reduction at

Mvumu was not that clear cut and that the three discoidal variants discussed grade into each other and simply represent differences within the same general conceptual approach.

140

For example, a core that began its use life as a preferential discoidal core could easily become a standard, non-preferential discoidal core as the needs of the toolmaker changed or as an adaptation in method to the changing character of the core itself as it was reduced. Likewise, a Levallois core could easily become a preferential discoidal core with only a slight change in the reduction methodology – the removal of usable flakes from its base – or a single preferential discoidal core could be re-prepared and reduced with recurrent removals. Such problems in the classification of similar cores have been reported at sites throughout the greater region (Clark et al 1970:341; Clark and

Kleindienst 2001:62). Refitting studies could provide better insight into this issue by reconstructing the reduction sequence leading up to the final form as excavated, but no refits were completed as a part of this analysis. Regardless of the exact life histories behind prepared cores at Mvumu, they are consistent with typical MSA prepared core reduction technology.

Although prepared cores are more interesting from a technological point of view, they are in the minority at Mvumu where simple reduction strategies account for 62.5% of the cores studied. Most simple cores are amorphous, multiplatform cores referred to in this study as simple cores. Such cores are present in essentially all Stone Age assemblages regardless of chronology or geography and they account for nearly half

(47.7%) of the total number of cores at Mvumu. They provide a fast and easy source of unspecialized, non-standardized flakes for use when a simple edge is required or as blanks for the manufacture of more complex tools with very little energy or time investment. Also common amongst simple reduction cores at Mvumu are the elongated

141

high backed cores described in this thesis. They employ more sophisticated reduction methods compared to simple cores where a series of flakes are removed from long continuous platforms, but their surfaces show no signs of deliberate preparation aimed at controlling flake morphology. The initial time and energy investment involved in producing these cores is greater than that required for simple cores due to the need to locate a cobble with a suitable striking platform, but once such a stone was found striking a series of flakes from a single long platform would be quite efficient. Like prepared cores, simple cores are also typically small with an average mass of 49.3g. This average is also skewed by a few large cores; 75% of simple cores have a mass of less than 50g

(Appendix A).

The near absence of bipolar cores (2.27%, n=3) at Mvumu is an interesting aspect of the core technology reported here considering what is known from other nearby quartz dominated assemblages. Bipolar reduction is often a common feature of quartz dominated MSA and LSA assemblages at many sites in southern Africa (e.g. Mumba –

Diez-Martín et al 2010; Mumbwa – Barham 2000; Twin Rivers – Barham 2000) and indeed throughout the world (e.g. – Hiscock 1982; Barber 1981; Sweden –

Knutsson 1988, Knutsson and Lindgren 2009). Its utility in the reduction of quartz cores, especially small quartz cores, is largely related to the fact that quartz is a hard stone

(Mohs‟ hardness 7) and therefore requires a strong blow to detach flakes (Dickson 1977).

Such blows are most easily delivered through hard freehand percussion, but this technique becomes increasingly difficult as core mass decreases. Cores with low masses possess insufficient inertia to allow for the efficient transfer of energy from the hammer

142

into the core. As a result, instead of the impact transferring energy into the core and detaching a flake the blow merely pushes the core away from the hammer and delivers a painful shock to the knapper‟s hand (Barham 1987). Bipolar reduction overcomes the tendency of a small core to move away from the percussor by bracing it upon an anvil.

Experimental studies have been conducted by several researchers to determine the mass threshold below which freehand reduction of quartz becomes inefficient/impossible.

This research has failed to produce a definitive answer, but most authors have arrived at reasonably coherent values. Dickson (1977:99) suggests the lower limit lies between 60g and 90g. Similarly, Barham (1987:48) suggests that a core must have a mass of at least

75g and adds that it must also be at least 50mm long for freehand percussion to be viable.

Knutsson (1988:89) cites a minimum mass of 50g. During experimental trials the author was able to reduce cores down to a mass of ~45g provided relatively acute platform angles (<60˚) were available (Figure 5.7: B). Regardless of which threshold is used, these results pose a problem for the Mvumu assemblage where many of the cores fall beneath the proposed lower mass and size threshold. Accepting Barham‟s (1987) figures of 75g and 50mm as minimums, only 13.63% (n=18) of the cores studied would have been reducible through direct freehand hard percussion. Even if Knutsson‟s (1988:89) more liberal 50g minimum is used the number of cores suitable for freehand reduction is only 35 (27.3%).

Although cores displaying formal bipolar reduction characteristics at Mvumu are rare it does not rule out the use of an anvil in the reduction of small cores. One possibility is that some form of “soft” anvil was employed to brace the core. A “soft”

143

anvil could be a tree trunk, exposed root, the knapper‟s leg, or even a hard patch of ground. The leg-braced method was used during the experimental trials when a discoidal core was successfully reduced to a mass of 45g. However, attempts to work beyond 45g in this way typically ended with a bruised thigh more often than a successfully detached flake. The use of a “soft” anvil could have allowed the Mvumu knappers to stabilize a small core enough to facilitate flaking without the bracing effect being strong enough to reflect the impact energy back up into the base of the core and cause an opposed removal as seen with the bipolar technique (Kuhn 1995). Clark and Kleindienst (2001) suggest that tightly wrapping a core in leather or bark before knapping may also help direct energy into the core and facilitate a removal. However, there is no evidence that such a practice was followed at Mvumu.

Although prepared core techno-types make up a significant portion of the cores at

Mvumu, the overall character of the assemblage suggests an expedient approach to reduction with little concern for the economic use of raw materials. This is most evident in the fact that simple cores, which require the least amount of investment in terms of planning and preparation and return the least standardized yield of usable blanks, are the most abundant techno-type recorded at Mvumu, but the notions of expediency and low regard for material conservation are also supported by a number of additional lines of evidence. Considered as a single category, cores following simple reduction strategies bear an average of only three flake scars. Prepared core techno-types double this value with an average of six removals, but six is still an exceptionally low number considering that 87.5% of prepared cores are bifacially flaked meaning an average of only three

144

removals per reduction volume. These negative scar figures, especially those from the simple core component, fall well short of those from a comparable quartz dominated

MSA assemblage from the West Bay locality at Lake Eyasi, Tanzania which shows an average of 7.6 removals for both polyhedral (simple cores in this study) and discoidal cores (Domínguez-Rodrigo et al 2007:63). While low negative flake scar counts combined with small size such as that discussed above have been used in some studies as indicators of intensive reduction to the point of core exhaustion (e.g. Diez-Martín et al

2009), this is not likely the case at Mvumu where 70% of simple cores and 52.1% of prepared cores (63.3% total) still possessed patches of cortex when they were discarded.

Bipolar cores, which have been interpreted by some as a response to the need to conserve raw materials (Hiscock 1996) are exceedingly rare at Mvumu (n=3) and blade cores, argued by some to be related to attempts to maximize the yield of usable edge from a given amount of raw material (Bar-Yosef and Kuhn 1999; cf. Eren et al 2008), are completely absent from the study assemblage. In short, it appears that little effort was made to extract the maximum number of blanks or amount of usable edge from cores at

Mvumu which given the abundance of easily accessible quartz in the area is not a surprising finding.

6.2.3 – Debitage

Debitage makes up the majority of all lithic assemblages with good compositional fidelity (Schick and Toth 1993). This holds true at Mvumu and the debitage portion of the assemblage provides additional insight into lithic reduction at Mvumu. As with

145

cores, debitage is dominated by quartz (95.54%). Again finer textured materials are most common with fine milky quartz outnumbering coarse and very coarse varieties by a ratio of 3.2:1. However, unlike the cores several non-quartz materials are also represented including quartzite, rhyolite, and chalcedony. Rock crystal quartz is also present.

Presumably if a larger sample of cores was examined these raw materials would also be represented in that portion of the assemblage. Although present, the non-quartz raw materials make up a minuscule portion of the debitage assemblage. Combined they are outnumbered by even the coarsest milky quartzes by 3.7:1.

One subtype of whole flakes, blade-like flakes, requires additional mention.

These pieces share a strong affinity to artifacts classified as blades at other MSA sites

(see regional comparisons below), but cannot be considered formal blades at Mvumu due to the lack of blade cores. True blades by definition must be struck from a specialized type of core set up to facilitate the continuous production of elongated flakes (Andrefsky

2005:253; Kooyman 2000:12, 170). Without blade cores, it is possible that the blade-like flakes were simply fortuitously struck long flakes and do not represent a deliberate technological product. Experimental knapping confirmed that long parallel sided flakes similar to the Mvumu blade-like flakes were sometimes produced from non-blade cores.

Since these pieces cannot be shown to be intentional technological products, the similarities between true blades and blade-like flakes can only be considered to be a case of equifinality and therefore, this study can only consider them to be blade-like; to do otherwise would be at the risk of attributing a technological approach to the Mvumu toolmakers that they did not consciously apply.

146

Whole flakes are the most common category of debitage at Mvumu. This finding is unusual for a quartz based assemblage due to quartz‟s tendency to shatter and fragment upon the impact of a percussor (Rose 2004; Tallavaara et al 2010). However, the high frequency of whole flakes is not without precedent. For example, whole flakes outnumber fragments in both C Block and F Block at Twin Rivers Kopje, Zambia (Clark and Brown 2001) and at the West Bay 9 locality on Lake Eyasi, Tanzania (Domínguez-

Rodrigo et al 2007). Experimental knapping aimed at testing quartz shattering tendencies conducted by Tallavaara et al (2010) has also demonstrated this possibility; one of the four testers produced primarily whole flakes when testing hard hammer reduction. The preferential use of fine textured quartzes may have also contributed to the high frequency of whole flakes at Mvumu.

The flake fragment component makes up the next largest category of debitage with proximal and distal fragments outnumbering medial fragments. This suggests that the majority of broken flakes possess only one fracture that essentially broke the flake into two pieces. Although unintentional flake fragmentation is a common feature in quartz debitage, a small percentage of flakes in this assemblage show signs of intentional breakage or truncation in the form of demi-cones and percussive impact points near the truncated margin. Demi-cones and impact marks are considered to be good indicators of intentional truncation as they represent evidence of fracture by percussion as opposed to other possible agents such as flaws in raw material (Tallavaara et al 2010) and trampling

(Gifford-Gonzales 1985; McBrearty et al 1998). Demi-cones are also found in other prehistoric assemblages featuring truncation (Moore et al 2009; Brumm et al 2010) and

147

on 19th century gun-flints that utilized truncation in their production (Knowles and

Barnes 1937:207), thus further strengthening their connection to the technique. Although the truncation technique occasionally causes flakes to shatter, experimental knapping

(this study) showed that with practice flakes can consistently be split cleanly into two pieces. Varying the angle of the flake on the anvil was also shown to be an effective method of controlling the shape of the fracture (e.g. straight or curved) during the experimental trials. Fragments showing these signs of intentional breakage total only 17

(3.88% of fragments), but this number is likely an underrepresentation of the true use of the truncation technique due to the poor capacity of quartz to display features of conchoidal fracture such as demi-cones.

The prevalence of whole flakes compared to fragments could be related to flake thickness. In the debitage studied here the average thickness of a whole flake is 6.74mm whereas fragments average only 5.28mm. Tallavaara et al (2010) offer a more sophisticated explanation stating that it is a higher relative thickness, the ratio of thickness to length, which affects the likelihood of fragmentation. Their theory holds that greater thickness in relation to length creates a higher incidence of intact flakes being produced. This idea seems to hold true at Mvumu where short “over-square” flakes make up nearly half of the flake assemblage (43.85%). A tendency toward thick flakes is also evident in striking platform thickness; typically the thickest portion of these flakes.

Platform thickness combined with generally short length creates a proximal to distal wedge-like morphology in many of the flakes analysed. In addition to contributing to the high percentage of whole flakes, this morphology also explains the high residual platform

148

angles seen on several cores. Thick platforms also suggest the use of hard hammer percussion as the primary mode of reduction.

The distal ends of flakes are dominated by feather terminations with a relatively even split between step and concave step (sensu Isaac and Isaac 1977: Figure 57) being the next most common types. Step terminations are followed by hinge terminations and overshot flakes are the least abundant. This is an interesting observation because step and hinge terminations are usually quite common in quartz-based assemblages due to the flawed nature of many quartzes (Dickson 1977). The high percentage of feather terminations may seem to run contrary to this, but could be a reflection of the preferential selection of fine textured quartzes that contain fewer internal flaws than coarser varieties with more manageable flaking characteristic at Mvumu.

The debitage component of the assemblage is consistent with the predominance of the simple, expedient reduction strategies seen in core reduction. The vast majority of flakes show irregular flake scar patterns on their dorsal surfaces. Radial and parallel patterns combined account for only 4.49% dorsal scar patterns. Platform preparation data follow suit; non-faceted platforms are by far the most common followed by a sharp drop in frequency to two faceted platforms and another significant decrease in frequency to three faceted platforms. Only a single four faceted platform was observed. Given the projected good compositional fidelity of the assemblage, the fact that whole flakes and proximal fragments (fractions of the flake component which possess striking platforms to prevent double counting of the same fragmented flake) outnumber flake scars on cores by a ratio of only 1.29:1 (901 flakes / 697 flake scars) suggests a short use life for most

149

cores. This strengthens the notion that reduction at Mvumu was expedient with little concern for maximizing the productivity of a given core.

Although the majority of flakes and flake fragments show no traces of cortex on their dorsal surfaces, all reduction stages are represented in the assemblage. Analysis of cortical coverage shows that the degree of coverage increases as debitage size increases

(Table 6.2). Cortex on striking platforms is also rare, but again although the majority of pieces have non-cortical platforms, examples with partial and complete coverage are also present. The degree of cortical coverage reported here should be considered a minimum indicator of the amount of primary reduction performed at Mvumu; survey of the local area indicates that much of the quartz on the slopes has only weak cortex development.

The low occurrence of cortex could be due to the quartz reduced at Mvumu eroding or being quarried from local pegmatites as large chunks that were then broken apart into core-sized cobbles without allowing time for significant cortex to develop on their surfaces. However, while there is a strong case for quarrying quartz at Mvumu, the presence of primary, secondary, and tertiary flakes as well as finished tools in the study assemblage shows that all stages of core reduction were conducted at Mvumu and site use was not limited to quarrying activities.

The debitage assemblage also contains intentionally and unintentional produced by-products of reduction including core trimming spalls, core preparation tablets, and shatter. In terms of intentional by-products, core trimming spalls are by far the most common. The archaeological interpretation of these pieces being produced while removing undesirable ridges and protrusions from a core that inhibited further reduction

150

Table 6.2 – Cortex by Size Fraction for Flakes, Flake Fragments, and Truncations ______Fraction % with cortex n= with cortex % without Cortex n= without cortex 1-10mm 2.20 2 97.80 89 11-20mm 11.57 73 88.43 558 >20mm 26.61 116 73.39 320 ______

was confirmed during experimental flintknapping trials (this study). Detaching such a piece was often as simple as rubbing the core with a hammerstone. Other times a light blow was required to remove the ridge. Core preparation tablets where a striking platform that had become unusable was removed to refresh the core are much rarer, but could have extended the use life of simple core types. The rarity of these artifacts is not surprising due to the projected short use life of individual cores at Mvumu. Why go to the trouble of rejuvenating a core when a fresh one is only a stone‟s throw away? The relatively low amount of shatter (n = 417, 25.5% of debitage) is, however, unexpected.

As mentioned above, quartz is often brittle and prone to shattering (Cotterell and

Kamminga 1990; Tallavaara et al 2010). As such it should be expected that shatter will make up a significant portion of any quartz assemblage. For example, at Katanda 16 where quartz makes up 84.5% of the assemblage, shatter accounts for 54.8% of the debitage (Williams 2005). Once again, the preference for fine textured quartzes at

Mvumu seems the likeliest and simplest explanation for the relative lack of shatter in the study assemblage.

151

6.2.4 – Tools

The tool component of the study assemblage is consistent with a typical MSA assemblage (Delson et al 2000; Klein 2009) and includes a variety of scrapers, points, awls, and various other small tool types. Fine textured milky quartz is again the most common raw material observed accounting for 79.44% of the tools studied and outnumbering all other types combined by a ratio of 3.9:1. Coarser varieties of milky quartz form the second largest raw material category (coarse milky quartz 10.75%; very coarse milky quartz 4.67%) amongst tools followed by quartzite (4.21%) then rhyolite

(0.93%). Despite appearing in lower numbers than even the coarsest milky quartzes, the tool category provides the highest representation for non-quartz raw materials. The technological approach taken by Mvumu flintknappers when working these non-quartz materials is indistinguishable from the methods employed for quartz.

Scrapers are the most common type accounting for more than half (53.74%) of the total tool assemblage. The scraper category also contains the greatest diversity of subtypes. The majority are characterized by convex working edges. In some cases the convexity of the working edge extends around >50% of the scraper‟s periphery becoming semi-circular and concave, convergent, and denticulate examples are also present. Edge angles form a continuum from ~40˚ through ~90˚. Twenty four scrapers also possess unusual bevelled bases that are not often reported in other MSA sites (cf. Clark and

Kleindienst 2001 who reported bevel-based core scrapers at Kalambo Falls, Zambia).

The purpose of the bevelled base is unknown, but possibilities include thinning to facilitate hafting or simply creating a grip that would allow the tool to be used in hand

152

while creating sufficient clearance for the user‟s knuckles above the material being scraped. Blank choice (flake blank or core fragment) for the production of scrapers is for the most part indiscriminate; all types represented by a sample size of greater than two, with the exception of core scrapers which are by definition made on core fragments, were manufactured on both flake blanks and core fragments. That being said, core fragments were the preferred blank type for scrapers by a slight margin accounting for 64.35% of the scrapers studied (58.59% when core scrapers are excluded).

Awls make up the second largest category of tools (14.49%). The key difference between the majority of the awls reported here and those described at other MSA sites is the method by which they were produced: the truncation method. To produce the “drill bit” or working end of an awl, a blank is laid on an anvil so that the piece that is intended to be removed by the truncation blow is flat against the anvil and the opposing end that will be the body of the finished tool is supported slightly above the anvil‟s surface.

When a truncation blow is struck with the objective piece in this position it should result in a curved fracture that runs concave towards the center of the awl. Several attempts were required to reproduce this type of fracture through experimental knapping, but once the details of the technique were worked out it could be reproduced consistently by knappers of moderate ability. The best results were gained by bracing the raised end of the objective on a folded piece of leather so that the leather brace ran parallel to the desired fracture and delivering light blows with a small hammerstone until the fracture occurred (Figure 4.3). Delivering too heavy of a blow meant the fracture was poorly controlled and often caused the objective piece to shatter.

153

From a morphological standpoint, most truncation awls are very similar to awls produced by intersecting flake scars and burination reported at Twin Rivers Kopje and

Mumbwa Caves in Zambia (Barham 2000). However, in the case of the awls from

Mvumu, the morphological indicators of truncation (e.g. impact areas on the dorsal/ventral surfaces of flake blanks, demi-cones, etc) rather than intersecting flakes or burination are clear; there are no burins in the Mvumu study assemblage. Clark and

Haynes (1970:395, Fig 30) mention potentially similar tools in their description of the lithics from Mwanganda, Malawi noting four “beaked” tools formed by two opposing notches, but unfortunately the descriptions and illustrations in their paper are insufficient to confirm whether the notches on these tools were formed by truncation. However, tools from Clark‟s excavations in Malawi have been viewed by the author in Malawian museum collections and they do appear to be produced in a similar fashion to those from

Niassa.

At Mvumu these tools often take advantage of the original blank morphology such that only one truncation is required, but others use two or three converging truncations to form the working end or ends of the awl. Truncation was not the only method used for producing pointed tools at Mvumu. Rarer awl types use retouch or take advantage of the shape of topknot flakes to form their working ends. A single example also incorporates a retouched scraping edge along one of its sides creating a multifunctional tool. A single example of an awl with the working end created by fine retouch and three topknot awls are also present. These types are also unusual in light of what is known from other quartz based assemblages in the region (see regional

154

comparisons below). Flakes were the blank type of choice for awl production, but six examples of single truncation awls on core fragments are also present in the assemblage.

Points at Mvumu account for the third largest group of formal tool types (7.01%).

Tools in this category are amongst the most morphologically standardized artifacts recovered. All were made on fine milky quartz. Of the three major point subtypes corner-struck points are the most common (n=10). These points are likely the product of the discoidal reduction observed in the cores (Brooks et al 2006). This assessment was tentatively confirmed through experimental knapping (this study). After a brief period of experimentation the author was able to successfully reproduce the bases of these points with their distinctive corner platforms, but it proved difficult to reproduce the distal end of the point. Attempted replicas ended short before reaching the desired pointed tip

(Figure 5.7:D). This was likely the result of the flintknapping ability of the experimenter in working quartz rather than a problem with the technique; greater success was achieved with other easier to work materials such as chert (Figure 5.7:C). As such, with more practice it is reasonable to assume that they could also be reproduced on quartz.

Striking platforms on the archaeological corner-struck points are typically oriented to the right corner of the base (80%) suggesting a fairly consistent production methodology. All examples possess unifacial retouch to create the final form. In terms of size, most corner-struck points form a continuum between 20mm and 37mm in length.

However, a single example nearly doubles the maximum length of the second largest corner-struck point at 63.71mm. Technologically this large point is identical to the more common smaller examples giving no clues to the purpose of its greater size.

155

Corner-struck points are followed by Levallois points then base-struck points in frequency. As their name suggests, Levallois points are likely associated with the

Levallois cores. Also in keeping with their namesake production methodology, these points show few signs of retouch to their working edges following their detachment from a prepared Levallois core. There is however post detachment modification to the bases of these points. The bases of two examples have been modified to create a tang which defines the shouldered/tanged subtype. The single simple-based Levallois point also shows signs of attempted thinning of its base in the form of a single flake removal that was aimed at, but unsuccessful in, removing the striking platform. Base-struck points are not connectable to any specific reduction strategy. Blanks used are simply triangular flakes. Minor unifacial retouch is present on both examples to finish the edges of the points, but their bases are unmodified. There are no size outliers in either Levallois or base-stuck points.

The function of these points cannot be determined with certainty, but they are consistent with artifacts often interpreted as projectile points at other MSA sites (Brooks et al 2006). Levallois points, with their tangs and basal thinning, are the only point type at Mvumu that show signs of intentional modification to aid hafting. Corner-struck points, by virtue of the location of the striking platform on the side of their bases, have naturally thinned bases that may have facilitated hafting. Base-struck points have no basal thinning, but they are naturally thin enough that they could have been hafted without modification. This potential for hafting is one of the key criteria that set points apart from artifacts in this assemblage referred to as pointed pieces which, despite their

156

generally triangular pointed morphology, are too thick to have been effectively hafted as projectile or spear points.

Two other tool types, crescents and snapped retouched pieces, may suggest hafting was part of the lithic technology at Mvumu. Like points, crescents and snapped retouched pieces show a relatively high degree of morphological standardization and although Mvumu has no true mode 5 component (e.g. geometrics), these tool types may have been used in a similar fashion. Crescents at Mvumu resemble backed tools from

Twin Rivers Kopje, Zambia (Barham 2002a) and Mumba, Tanzania (Diez-Martín et al

2009) that are argued to have been hafted. Snapped retouched pieces would also almost certainly have had to be hafted to be usable due to their small size. The wedge shaped morphology of the snapped retouched pieces would have served hafting well by creating a tang that could be used to anchor the piece into a shaft. No analogous tools have been described from sites outside Niassa. Neither crescents nor snapped retouched pieces appear to be related to the microliths of the Howiesons Poort industry as there is no indication of a blade basis for their production or preferential selection of high quality exotic raw materials (Soriano et al 2007); they were made on simple flakes/flake fragments and the same local fine milky quartz preferred for other tools types at Mvumu.

The possibility that the hundreds of sharp flake fragments and pieces of shatter in the debitage assemblage from Mvumu could have been hafted and used as “microliths” without recognizable modification also remains open. Archaeological experiments and ethnographic accounts testify that such artifacts are suitable for the completion of many tasks (White 1967; White and Thomas 1972; Frison 1979; Andrefsky 1994) and many

157

pieces from the Mvumu assemblage could have been hafted as single pieces or in groups to form composite tools.

Remaining tool types include notches, chopper-like pebble tools, pointed pieces, and retouched flakes. Tools in these categories exhibit great morphological variability and appear very much as simple, expediently produced, unspecialized tools with working surfaces formed with unsystematic retouch aimed at producing a usable edge or point.

Notches are the one possible exception to the interpretation of these tools as unspecialized as their working edges are consistent with notches from other sites that are suggested to have been spokeshaves used to form wooden shafts (Wadley 2005).

However this potential use remains only speculation as no direct evidence is available to confirm it at this time.

Conspicuously missing from the Mvumu assemblage are large core tools such as core-axes and picks. Given the chronology of Mvumu, the lack of “archaic” core tools may not seem unusual, but such tools have been recovered from other sites in Niassa including nearby Mikuyu (Mercader et al 2008) and Ngalue (Mercader et al 2009a).

Core tool technology at Mikuyu is especially relevant give the extreme close proximity

(<1km) and the fact that the archaeological materials there were recovered from the same geological unit as those from Mvumu. Further examples of large core tools have also been collected during surface survey of the area around the Mvumu and Mikuyu excavations. Other aspects of lithic technology that have been recorded in Niassa but are absent from the assemblage studied here include large lanceolate bifacial points and grindstones. The bifacial points, superficially resembling crude Lupemban points, are

158

only known from surface finds so their affiliation to the Mvumu study assemblage and its chronology is unknown, but grindstones were found in the Middle Beds at Ngalue

(Mercader 2009; Mercader et al 2009a) confirming their existence within industries consistent with those from Mvumu.

6.3 – Behavioural implications of lithic technology at Mvumu

From a techno-typological point of view, the lithics from Mvumu fit well within a general MSA context. The assemblage contains all of the classic markers of the MSA including points and small scrapers, radially reduced cores, and an absence of large core tools. It is however difficult to fit into a more confined techno-typological classification within the MSA. Although core-axes, picks, and bifacial lanceolate points have been found in the Niassa region (Mercader et al 2008; Mercader et al 2009a), they are not present in the Mvumu study assemblage and therefore despite some similarities in the small tool components Mvumu is not considered to be an extension of the Sangoan or

Lupemban complexes.

Considered as a whole, the MSA lithic technology from Mvumu is characterized as an expedient and informal industry. These traits can be identified through a number of lines of evidence, many of which have been touched on above. The heavy reliance on local raw materials is one clear sign of the expedient approach. The near complete focus on local raw materials would have greatly reduced the amount of energy and planning required to maintain a lithic toolkit. Moreover, while the abundance of quartz at the

Mvumu locality is exceptionally high, quartz is also a relatively easily acquired resource

159

throughout the Lago and Sanga districts of Niassa and therefore employing a lithic technology suited to quartz would have allowed Niassan MSA populations to access a wealth of lithic resources throughout the greater region.

The easy availability of quartz may have been a contributing factor to another sign of expedience at Mvumu: the lack of concern for the conservation of raw materials during reduction. The indifference of the Mvumu flintknappers to economizing raw material usage can be seen in the high percentage of cores that retain patches of cortex

(62.88%) as well as in the low number of negative scars present of many examples.

Tools such as scrapers and awls made on large core fragments as opposed to flakes also show the low priority of using lithic materials to the fullest degree possible.

Informality is best seen in the tool component of the assemblage. Although formal tool types such as scrapers, awls and points are present in the assemblage, the manner in which they were manufactured and their overall morphology is distinctly informal. In most cases, the overall morphology of tools away from working edges is largely inconsistent. Also, although 87.38% of tools were finished with some sort of retouch, the amount of retouch used is minimal, confined to tool margins, and almost always unifacial. The indiscriminate choice of blank types further suggests that tool manufacture was an informal process with inconsistent reduction sequences for given end products aimed only at the end result and showing little regard for how it was achieved.

The one exception to this rule is the points which due to their presumed function as projectiles must conform to certain morphological and aerodynamic requirements to be effective (Shea 2006; Sisk and Shea 2009; Lombard and Phillipson 2010). A tendency

160

toward informality can also be seen in the high representation of simple reduction methods in the assemblage over more regular and deliberate prepared core strategies.

The informality seen in this assemblage could be a by-product of the expedient manner in which tools were produced. It could also be related to the unpredictable,

“poor” flaking characteristics of quartz (Andrefsky 1994), but given the fact that quartz and non-quartz raw materials are treated in the same manner and used to produce equally non-standardized tools these attributes alone cannot explain the phenomenon. Moreover, neither of these ideas helps explain the expediency of the industry. It is more likely that both the informality and expediency seen here are related to raw material economy at

Mvumu. As noted above, quartz is extremely abundant at Mvumu and such good availability often contributes to the expedient use of a raw material (Andrefsky 1994;

Tallavaara et al 2010). More formal techniques are typically related to attempts to maximize yield from raw materials (Andrefsky 1987) and with plentiful quartz on hand there was simply no need to expend energy on raw material conservation. This notion is well articulated by Andrefsky (1994) who based his conclusion on a review of ethnographic and archaeological evidence from Australia and the United States that showed a clear tendency for informal tool production when low quality raw materials are abundant. While Andrefsky‟s work is not directly related to the African context being discussed here, the theory is consistent with observed assemblages in Niassa and offers a possible explanation for the technological behaviour seen at Mvumu.

Alternatively, or perhaps concurrently, the informal expedient nature of this assemblage may be related to reduced resource procurement ranges (Parry and Kelly

161

1987; Henry 1989; Torrence 1989; Bamforth 1990; Andrefsky 1991; cf. Andrefsky

1994). To reduce the amount of energy required for the transportation of raw materials

Stone Age populations who relied on resources distributed over a wide area tend to show a preference for high quality raw materials and formal reduction strategies that allow them to extract the highest possible yield from a given amount of raw material. These groups also gravitate toward curating and renewing tools to extend their use life

(Andrefsky 1994). Groups focusing on resources from more constrained areas are less constrained by resource transportation costs and therefore they are often associated with lower quality raw materials and less economical reduction strategies when doing so is more convenient. Tools produced by such groups tend to have shorter use-lives and there is little evidence for curation (Andrefsky 1994). This is the case at Mvumu. Quartz‟s tendency to shatter unpredictably means that it typically contains less usable tool edge than higher quality stones (Tallavaara et al 2010), but this would have been of little concern to MSA people at Mvumu considering the abundance of quartz on the hillside.

Likewise, the lower efficiency of the informal production methods would have been mitigated by the constant easy availability of quartz in the Mvumu area.

Together, the expediency and informality of the Mvumu lithics creates a sense of crudeness in the assemblage. Crudeness in lithic repertoires is often interpreted as a measure of archaic behavioural capacity in the toolmakers. This notion is however, difficult to reconcile with the final MSA chronology of the Mvumu assemblage (~28-

33kya) and the current theories regarding the origins of modern human behaviour as summarized in chapter 2 that all argue behavioural modernity would have been in place

162

by this time. To understand this we must first acknowledge the fact that there is no a priori reason to assume that crude lithics represent archaic or less sophisticated behavioural capabilities. Simple, crude lithic technologies can often be well adapted to the needs of toolmakers that are complex and modern in every way. A good example of this can be seen in the “crude” lithic technology employed by early people in Australia which, like the lithics from Mvumu, is often quartz-based (e.g. Kamminga 1982; Hiscock

2003). Hiscock (2003) illuminates this point noting that although Aboriginal lithic technology is typically informal with little evidence of imposed form, there is abundant archaeological evidence for other “modern” behaviours in the form of mobile art, complex burials, and multi-component organic artifacts; there is no direct correlation between the crudeness/sophistication of lithic artifacts and the cognitive or behavioural capacities of the people who made them. While direct evidence for the types of behaviours cited by Hiscock is not present at Mvumu, we cannot assume that MSA people at Mvumu were not behaviourally modern simply because of crude looking stone tools without first ruling out other potential social and economic contributors to the technological nature of the assemblage. Moreover, although the overall appearance of the lithic technology from Mvumu is crude, many of the technological markers of modern human behaviour that were outlined in chapter 2 are present including prepared cores, specialized tools such as projectile points, and tools that would have required hafting.

163

6.4 – Comparisons with temporally and geographically related sites

Inter-site comparisons of MSA lithics can be a challenging endeavour. Such work is frequently complicated by a number of factors including a lack of available technological analyses, inconsistent classification schemes, specialized research goals that only focus on one small part of an assemblage, and poor chronological resolution at many sites (Basell 2008). Detailed quantitative data such as that presented here is also rarely produced and even when it is it is typically not published. But in spite of these difficulties, it is still useful to compare assemblages to whatever degree is possible to gauge variability and evaluate regional differences in stone tool technology both chronologically and geographically.

One of the most interesting aspects of the lithics described here is the late chronology of the assemblage. Geochronological analysis conducted at Mvumu has produced a date of ~28-33kya (Mercader et al, in prep). This places the assemblage in a very late phase of the MSA referred to as the final or terminal MSA which include MSA materials that are younger than ~42k years old (Wadley 2005). Final MSA sites are a rarity in African Stone Age research (Jacobs et al 2008), but a few similarly dated sites are known in South Africa including Sibudu Cave (Wadley and Jacobs 2004; Wadley

2005; Jacobs et al 2008), Rose Cottage Cave (Wadley 1993, 1997; Soriano et al 2007),

Strathalan Cave B (Opperman and Heydenrych 1990; Opperman 1996), Boomplaas

(Deacon 1995), and Florisbad (Kuman et al 1999). Unfortunately, many of these sites are insecurely dated and reported chronologies can only be considered minimum dates

(Wadley 2005). Sibudu Cave and Rose Cottage Cave are amongst the best dated of the

164

South African final MSA sites. Both sites have recently been re-dated by OSL which produced dates for the final MSA occupations of 38.6kya ± 1.9ky for Sibudu (Jacobs et al

2008) and 57kya ± 3ky - 33kya ± 2ky for the post-Howiesons Poort levels at Rose

Cottage (Soriano et al 2007). The chronology and assemblage integrity of the final MSA at Sibudu Cave is bolstered by a recent ESR date of 39kya ± 4ky and good stratigraphic security created by a hiatus beneath the final MSA running from ~47kya to 39kya (Jacobs et al 2008) paired with no evidence for an overlying LSA occupation that could be mixing with the final MSA (Wadley and Jacobs 2004).

These South African final MSA industries share some characteristics with the same period in Niassa, but several aspects set them apart as distinct. Technologically, both assemblages are flake-based; blades are rare (Wadley 2005; Soriano et al 2007).

The manner in which flakes were produced, however, was markedly different from flake production in Niassa. At Sibudu flakes were primarily produced from bipolar and

“minimal” cores (minimal cores are defined as cores from which only a few randomly placed flakes were removed; Wadley 2005:54). Prepared cores of all types, including

Levallois and blade, are rare. Prepared core types are also rare throughout most of the post-Howiesons Poort strata at Rose Cottage; they are absent from the LYN layer of the site which produced the only true final MSA date of 33kya ± 2ky (Soriano et al 2007).

Bipolar cores are also rare the final MSA at Rose Cottage; absent in the LYN layer (Clark

1997a; Soriano et al 2007). There is no mention of radial core reduction in the final MSA at Sibudu and it is insignificant at Rose Cottage representing only 4% of cores (Clark

1997a).

165

Local raw materials were used often at both Sibudu and Rose Cottage, but although it is present in both assemblages neither are dominated by quartz. Rose

Cottage‟s lithics are made on a chalcedony-like stone referred to locally as opaline with volcanic tuff used second most often and negligible amounts of hornfels, milky quartz and fine-grained quartzite making up the remainder (Clark 1997a; Soriano et al 2007).

The most common raw materials at Sibudu are dolerite and hornfels with small amounts of quartzite, quartz, and fine-grained sandstone also being used (Wadley 2005). Hornfels is the one possible exception to the preference of local raw materials. It may have been quarried from now buried outcrops near the site, but the closest currently accessible source is ~20km away (Wadley 2005). It is worth noting that if this remote source was indeed the origin of the hornfels used at Sibudu it was more distant from the site than the

Lunho River sources hypothesized for the rare high quality materials seen at Mvumu.

Scrapers and points are the most common tool categories at both Sibudu and Rose

Cottage Caves. This is typical throughout the MSA including at Mvumu. There are however differences in the techno-typological details of these tool categories that make them distinct. The best example of these differences is seen in the points in the assemblages. Both Sibudu and Rose Cottage contain both unifacial and bifacial points

(Wadley 2005; Soriano et al 2007; Jacobs et al 2008), but bifacial points are absent at

Mvumu. Sibudu also contains points that are referred to as hollow-based points – triangular bifacial points with concave thinned bases – that are seen as the fossil directeur of the final MSA at the site (Wadley 2005; Jacobs et al 2008). These points are absent from both Mvumu and Rose Cottage. Corner-struck points, the most abundant type at

166

Mvumu are absent from both South African sites which is not surprising given their lack of discoidal cores. Scrapers also share some common ground. Little detail on the overall morphology of scrapers is given in the discussions of the Sibudu and Rose Cottage assemblages, but the same general types present at Mvumu seem to be present including convex side/end and convergent examples (Wadley 2005; Soriano et al 2007).

Differences come in the technological basis for their manufacture; 64.3% of scrapers at

Mvumu were produced on core fragments, but there is no evidence for a similar practice in the flake-based assemblages from Sibudu or Rose Cottage. Like Mvumu, both of these sites contain small numbers of backed tools that are distinct from those seen in the

Howiesons Poort levels that underlie these South African final MSA assemblages

(Wadley 2005; Soriano et al 2007). Backed tools have been argued to represent symbolic, culturally mediated artifacts (Wurz 1999; Barham 2002) that may have even been used in gift exchange systems (Villa et al 2010), but despite the techno-typological similarities shared by these tools in eastern South Africa and Niassa it is unlikely that they represent any cultural connection between final MSA people in these areas.

Other sites in South Africa with similar flake-based final MSA assemblages include Umhlatuzana (Kaplan 1990) and Florisbad (Kuman et al 1999). These sites are however problematic. Umhlatuzana is very similar to Sibudu in terms of its technology including hollow-based points. However, questions have been raised regarding the stratigraphic integrity of the site and it may be intermixed with LSA materials from above

(Wadley 2005). The final MSA at Florisbad is reported as a flake-based informal industry, but it suffers from a very small sample size and poor preservation which makes

167

accurate definition and comparison difficult (Kuman et al 1999). The radiocarbon date of

19530 ± 650BP produced in the 1950‟s is also problematic and should only be considered a minimum date (Kuman et al 1999; Wadley 2005).

A number of other late and final MSA sites with greater technological differences are also present in South Africa. The key difference reported is a focus on blade production rather than flakes. This is evident at sites including Boomplaas (Deacon

1995), Strathalan Cave B (Opperman and Heydenrych 1990; Opperman 1996), Highlands

Rock Shelter (Deacon 1976; Volman 1984), and Klasies River (Singer and Wymer 1982;

Wurz 2002; Villa et al 2010). Older strata at Sibudu and Rose Cottage also contain blade-based Howiesons Poort industries, but such technology was not utilized in the final

MSA. The reason blade-based technologies were abandoned at some sites but remained in fashion at others is not well understood. Regardless of the reasons, blade-based technologies have not been found in Niassa at any site dating to between 105-33kya

(Mercader et al 2009a; Mercader et al, in prep) so any comparison with these industries in search of commonalities is largely moot.

Sub-Saharan final MSA sites outside of South Africa are rare, but some examples have been described. Unfortunately, many of these sites suffer from similar problems to those seen in South Africa including poorly controlled chronology, small sample sizes, publication in obscure unavailable reports, and inconsistent techno-typological approaches that make inter-site comparisons difficult. One such site is Mumba

Rockshelter near Lake Eyasi, Tanzania. Mumba has been the focus of intermittent archaeological research since the 1930‟s (Kohl-Larsen 1943; Roller 1954; Müller-Beck

168

1978; Mehlman 1979, 1987, 1989; Marks and Conard 2007; Dominguez-Rodrigo et al

2007; Dominguez-Roddrigo 2008; Prendergast et al 2007; Diez-Martín et al 2009).

Recent work has overcome the problems related to sample size and accessible publication, but some issues remain regarding the chronology of the final MSA at

Mumba and whether it should be considered MSA or reclassified as LSA (see below).

The materials associated with the final MSA come from a stratigraphic level referred to as Bed V. Eleven dates have been produced for Bed V, but unfortunately they do little to clarify the chronology of the unit. The majority of these dates postdate the 42kya threshold for the final MSA proposed by Wadley (2005), but two from the middle part of the package – 46600 ± 2050BP and 65686 ± 6049BP – suggest parts of Bed V may predate the final MSA boundary. Contrarily, some dates, including one from the lower part of the package at 20995 ± 680BP, suggest an extremely young age more in line with typical LSA chronologies.

A variety of raw materials were utilized in the manufacture of the Bed V lithics including several types of quartz, orthogneiss, metarhyolite, nephelinite, basalt, microgabbro, and chert (Diez-Martín et al 2009). Overall coarse-textured quartz is the most common. The bipolar technique accounts for ~50% of the reduction of these materials with lower quality stones being preferentially channelled toward the method

(Diez-Martín et al 2009). Non-bipolar reduction strategies include radial/discoidal,

Levallois, pyramidal, and single platform techniques (Diez-Martín et al 2009) and are more reminiscent of the reduction methods seen at Mvumu, but they are the minority and no mention is made of expedient freehand cores such as the simple cores that dominate

169

the Mvumu assemblage. The tool component of the Bed V assemblage is primarily composed of backed pieces made on both freehand and bipolar flakes with scrapers also being relatively common. Backed pieces include crescents resembling the rare examples of this type at Mvumu, but also a variety of other types creating a much more diverse

“microlith” assemblage than has been recorded in Niassa. Scraper types include types common at Mvumu such as side and end convex scrapers, convergent scrapers, denticulates, and notches (Diez-Martín et al 2009). One striking difference between the

Mvumu assemblage and Mumba Bed V is the scarcity of points at the latter. The re- excavation assemblage described by Diez-Martín et al (2009) does not contain a single point in its 7184 lithic sample. Likewise, no points are known from a much larger sample of 158101 lithics collected by Mehlman in 1981 (Diez-Martín et al 2009); although this assemblage has not been thoroughly studied (Mehlman 1989). The Mumba industry was originally interpreted as a transitional industry between the MSA and LSA (Mehlman

1987, 1989), but based on new chronological and techno-typological data from their recent re-excavation of the site Diez-Martín et al (2009) have reclassified the Bed V

Mumba Industry as purely LSA due to the high prevalence and morphological variation of microliths, reduction in the number of scrapers, and general decrease in the size of all detached products.

It should be noted that in general there is little consensus amongst archaeologists on what truly distinguishes the MSA from the LSA from a technological or any other standpoint (Sampson 1974; Phillipson 1993; McBrearty and Brooks 2000; Wadley

2005:58; Prenergast et al 2007; Diez-Martín et al 2009). Moreover, the date of the

170

transition is not consistent in the archaeological record. McBrearty and Brooks (2000) place the transition around 50kya. Some others prefer a later date around 40kya

(Beaumont 1978; Grün and Beaumont 2001). Both of these dates are problematic for the interpretation of the Mvumu assemblage as representing an MSA techno-typology. The only researchers that propose a MSA/LSA transition date that is compatible with the chronology seen at Mvumu are those working in South Africa who cite transition dates between 30-23kya (Wadley 1993, 1997; Mitchell 1994, 2002; Deacon 1995; Clark 1997a,

1997b), but as discussed above these industries are not techno-typologically related to known sites in Niassa (this study; Mercader et al 2008; Mercader et al 2009a).

Industries spanning the MSA/LSA are also known from several sites discovered in archaeological surveys along the Songwe River in the Mbeya region of southwestern

Tanzania (Willoughby 1996, 2001; Willoughby and Sipe 2002). The MSA here is characterized by a flake-based industry characterized by radial reduction and dominated by scrapers and points with rare burins, becs, bifaces, and backed pieces (Willoughby

1996, 2001; Willoughby and Sipe 2002). Rare Levallois reduction has also been reported

(Willoughby 2001). Some scrapers are reported to possess basal thinning that

Willoughby (1996) suggests may be related to hafting, but the description of this trait is insufficient to evaluate any similarity it may have to the bevel-based scrapers at Mvumu which could also be interpreted as being basally thinned. The LSA sees a shift toward a heavier reliance on backed and microlithic tools based on bipolar reduction (Willoughby

1996). Small single platform, pyramidal, and prismatic cores are also present in the LSA assemblages from Mbeya (Willoughby 2001). Quartz is the preferred raw material in

171

both the MSA and LSA, but chert, volcanics, and quartzite are also utilized (Willoughby

1996, 2001). Unfortunately, the situation in southwestern Tanzania is complicated by a lack of chronological control beyond conjecture based on typological assessments and the lack of sites that contain both MSA and LSA components in context. As such it is impossible to isolate the final MSA from the MSA and LSA in general and therefore no conclusions about the nature of the final MSA in the region can be drawn from this material. There do however seem to be noteworthy similarities between the MSA materials from Mbeya and Niassa.

Other industries in the south-central African archaeological record seem to be a much better techno-typological match for the Mvumu assemblage, but many predate it by a considerable margin or are simply not reliably dated at all. Two sites that show a strong techno-typological affinity to Mvumu are Twin Rivers Kopje and Mumbwa Cave in

Zambia, but assemblages from these sites are considerably older. Twin Rivers is the older of these two sites with the majority of its MSA component falling between 266-

170kya and possibly extending back to >400kya (Barham and Smart 1996; Barham

2000). Some younger dates ranging from ~101kya to ~13kya have also been produced from one excavation block at the site, but possible mixing of MSA and LSA sediments

(and therefore the archaeological materials) complicate the interpretation of these dates

(Barham 2000). The archaeological sequence at Mumbwa runs from >170kya right into the Holocene (Barham 2000), including the MSA/LSA transition (Barham 1995b), but unfortunately the resolution within this broad period is not overly detailed. The three earliest levels merely possess dates of >170kya. Other levels are attributed to huge

172

brackets of time including an archaeologically poor unit dated to 170-130kya. The main

MSA level spans 120-107kya and is followed by a hiatus from ~100-40kya at which point a small late MSA assemblage is present. The chronology of the MSA/LSA transition in the level above the late MSA is undated.

While these sites may not be directly comparable to Mvumu temporally they do share many technological elements. Assemblages at both sites are highly dependent on local vein quartz (Barham 2000; Clark and Brown 2001). Other raw materials include different varieties of quartz, chalcedony, chert, quartzite, limonite, dolerite, and a few other rare stones (Barham 2000). While local materials were preferred, there is some evidence at Mumbwa that, unless closer sources that are no longer visible were available during the MSA, materials were occasionally procured from up to 200km away (Barham

2000). Cores were reduced by similar means to those used at Mvumu including prepared methods like radial/discoidal and Levallois and simple informal approaches (Barham

2000; Clark and Brown 2001). Throughout much of the sequence, especially the younger portions, the simple informal strategies dominate; prepared types become less common with time. At Mumbwa the tendency toward simple reduction is seen in the prevalence of expedient “chunk” cores and single-/multi-platform types (Barham 2000). This preference peaks around 40kya when the record shows a “technological shift” away from radial reduction towards more informal approaches (Barham 2000). Twin Rivers also contains many chunk and single-/multi-platform cores, but it shares a connection with

Mumba, Tanzania in that bipolar reduction is the most prominent reduction method

(Barham 2000; cf. Clark and Brown 2001). Bipolar is present at Mumbwa, but never

173

with the frequency seen at Twin Rivers. The main difference between core technology at these Zambian sites and at Mvumu is the presence of blade and bladelet cores. Both

Mumbwa and Twin Rivers have a blade component from early in their sequences that is downsized to become bladelet technology with time (Barham 2000). These blade technologies never form a large part of the MSA at Mumbwa or Twin Rivers, but its presence is significant. No such parallel exists at Mvumu or any of the other known sites in Niassa (this study; Mercader et al 2008; Mercader et al 2009a).

The tool components of the Twin Rivers and Mumbwa assemblages are slightly different than at Mvumu in terms of composition, but the same general typology is present. At the Zambian sites awls and borers are the most common tool category

(Barham 2000). Some awls and boring tools are manufactured by striking one or more burin blows to form the working end of the tool, but many other have working ends formed by two intersecting flake scars that create a broad point (Barham 2000). This technique is not seen at Mvumu, but it is more reminiscent of the truncation method employed at Mvumu than burination-based methods and produces a similar broad working end. At Mumbwa, awls/borers are closely followed in frequency by scrapers, but at Twin Rivers backed tools form the second most numerous category (Barham 2000,

2002). Backed tools here are represented by 11 distinct subtypes and are amongst the oldest backed tools known in the MSA dating to as much as 300kya (Barham 2000,

2002). Backed tools are also present at Mumbwa, but are much rarer (Barham 2000).

Scrapers round out the top three tool categories at Twin Rivers and add support to the notion of informality seen in the core technology by being made on a variety of blank

174

types. Points are the third most common type at Mumbwa (Barham 2000). Points at both sites are often bifacial and include lanceolates that appear as small, crude Lupemban points (Barham 2000). Unifacial points are rare but present at both sites (Barham 2000:

Clark and Brown 2001). Some points, again at both sites, possess thinned, tanged bases that suggest they were hafted (Barham 2000). The backed tools reported above also presuppose hafting (Barham 2000), providing multiple lines of evidence for its presence in the technological behaviour of even early MSA peoples in Zambia. Rare heavy duty tools such as core-axes are also present (Barham 2000; Clark and Brown 2001).

Overall the assemblages from Twin Rivers Kopje and Mumbwa Cave are argued to be a close match for the Lower Lupemban (Barham 2000; Clark and Brown 2001).

The Lupemban at Twin Rivers is the only well dated example and occurs between 270-

170kya (Barham 2000). Similar assemblages are known from throughout the greater region at sites including Kalambo Falls, Zambia (Clark 1969, 1974, 2001), Redcliff Cave,

Zimbabwe (Brain and Cooke 1967; Cooke 1978, Cruz-Uribe 1983), and even as far away as northeastern Angola (Clark 1963). Unfortunately, these sites are not well dated so it is unclear how they relate temporally to Twin Rivers and Mumbwa or to Mvumu and the other known sites in Niassa. However, although Mvumu lacks the bifacial lanceolate points present at these sites that are diagnostic of the Lupemban, many other aspects of the technology are quite similar despite the implied temporal disconnect.

Other South-Central Central African assemblages from throughout the MSA also show affinities to the lithic technology from Mvumu. Undated lithics from Bambata

Cave in Zimbabwe show similarities to the industry from Mvumu in the heavy use of

175

milky quartz as a raw material, similar awls from a morphological and presumably technological point of view, and corner-struck points (Armstrong 1931). Corner-struck points are also common in MSA sites in Botswana such as ≠Gi, Rhino Cave, and White

Paintings Rockshelter (Robbins et al 2000; Brooks et al 2006; Phillipson 2007a).

Technological affinities can also be found at sites in the tropical forests of D.R. Congo such as Katanda 9 (Yellen 1996) and Katanda 16 (Williams 2005). Lithic technology from these sites dates to ~174-90kya (Brooks et al 1995) and is primarily made on quartz with some quartzite and small amounts of chert and sandstone (Yellen 1996; Williams

2005). Radial core reduction is common at both sites, but at Katanda 16 formless cores are most abundant (Yellen 1996; Williams 2005). Like at Mvumu, retouched pieces do not conform to tightly defined morphological categories (Yellen 1996). Scrapers are the most common tool type with rare heavy bifaces, points, modified flakes, and grindstones also occurring (Yellen 1996; Williams 2005). The points in the Katanda area bear no resemblance to the Lupemban lanceolates from Zambia (Yellen 1996).

Not surprisingly, the techno-typologically closest assemblages to Mvumu come from other sites in the Niassa basin. Several sites in the Karonga district in northern

Malawi on the west side of Lake Niassa provided the first look at lithic technology in the area, but unfortunately little information about them is available and they are not well dated (cf. Kaufulu and Stern 1987; Kaufulu 1990; Betzler and Ring 1995). Work here began in the 1960‟s and consisted of surface collections and excavations, but the available reports provide only basic information on the lithic technology that was encountered (Clark 1966; Clark and Haynes 1970; Kaufulu and Stern 1987). Based on

176

314 artifacts, Clark and Haynes‟ (1970) report on the excavations conducted at

Mwanganda is the most complete description of the MSA lithic technology in Malawi.

Quartz and quartzite are the most commonly used raw materials and they were used in roughly even quantities (Clark and Haynes 1970). Reduction was conducted by breaking cobbles then striking flakes from one or two simple platforms or using a radial approach.

No Levallois reduction or blade cores are reported. The tool assemblage is primarily composed of small light duty tools, but rare heavy duty tools including core scrapers and core-axes are also present. Like at Mvumu, scrapers are the most abundant small tool type and show a variety of working edge morphologies including convex, concave, straight, and denticulate. Becs or “beaked” tools formed by the intersection of two notches and reminiscent of the double truncation awls at Mvumu are also present at

Mwanganda. The description given by Clark and Haynes (1970) of these lithics is supported by personal observations made by the author in museum collections in

Mangoche and Blantyre, Malawi. No mention is made of points in the report from

Mwanganda, but Clark and Haynes (1970) do report faunal remains, including those of an elephant, that have led to the site being interpreted as a kill and butchery site centered on the animal. As such, even without points Mwanganda suggests that hunting was a part of the behavioural repertoire of MSA people in the Niassa Rift region.

Lithic technology from sites on the Mozambican side of Lake Niassa is known from larger assemblages than those in Malawi and is securely dated. Mvumu (this study;

Mercader et al, in prep) has received the most comprehensive lithic analysis, but lithics from the highland cave site of Ngalue (Mercader et al 2009a) ~70km inland from the lake

177

shore and the nearby open air site Mikuyu (Mercader et al 2008) have also been studied by PAC. MSA lithics from Ngalue come from a unit referred to as the Middle Beds which is bracketed by a basal U-series date of 105 ± 13kya and two dates by the same method of 55 ± 5kya and 55 ± 14kya at the level‟s upper boundary (Mercader et al

2009a). These Middle Bed lithics conform closely to the techno-typological classification created for Mvumu. The assemblage is almost entirely composed of quartz with rare pieces of quartzite and rhyolite making up the remainder. Discoidal reduction is the most common reduction method, but significant amounts of simple reduction are also present. Levallois and bipolar cores are scarce and blade technology is absent.

Tools at Ngalue show only minimal retouch and are dominated by scrapers. Points include the corner-struck, Levallois, and base-struck varieties seen at Mvumu and awls also consist of the same types reported at Mvumu. Two crescents and a small (<20mm) thumbnail scraper are the only evidence of microlithic tools at Ngalue. Heavy duty tools are limited to a faceted quartz core tool and a rhyolite core-axe/grinder. A ground piece of rhyolite interpreted as a grindstone was also present.

The published portion of the assemblage from Mikuyu is too small to make proper comparisons (n=33, Mercader et al 2008), but additional data from unpublished reports (Bennett, n.d.; Raja 2008) which discuss larger samples of the artifacts confirms the assemblage‟s affinity to Mvumu. Raw material selection essentially mirrors that seen at Mvumu. Cores were reduced by simple and prepared techniques in essentially even proportions with no evidence of blade and bipolar core technologies. The tool component of the assemblage contains techno-typological features common at Mvumu;

178

the majority of awls had their working ends formed by truncation, point techno-types include corner-struck, Levallois, and base-struck varieties, and artifacts resembling the snapped retouched pieces described at Mvumu were also recorded. Distinct from

Mvumu are a single lanceolate point was found on the surface near Mikuyu providing the only evidence of a Lupemban-like component resembling that reported in Zambia by

Barham (2000) and a core-axe also from the surface. Mikuyu remains undated, but its assemblage is derived from the same geological unit as Mvumu‟s lithics so it is reasonable to assume a similar chronology.

179

Chapter Seven: Conclusions

This thesis represents the largest and most detailed technological analysis of lithics from the Niassa basin conducted to date. The artifacts from Mvumu‟s main excavation trench were analyzed from a primarily quantitative perspective to evaluate the technological behaviour of the people who produced them. From a techno-typological standpoint the Mvumu study assemblage easily fits within the MSA. In broad terms, lithic technological behaviour at Mvumu during the final MSA is defined by the following traits:

1) The nearly exclusive use of quartz as a raw material with a preference for

fine grained quartzes

2) The nearly exclusive use of local raw materials (<15km procurement

radius)

3) The use of simple and prepared core technologies

4) A tool assemblage dominated by scrapers, awls, and points

5) An expedient approach to reduction with a lack of concern for raw

material conservation

6) Little morphological standardization of finish products

7) Very limited use of bipolar technology

8) A lack of Mode 4 and Mode 5 technology

9) The presence of all phases of reduction from core initiation to tool

manufacture

180

10) The persistence of “crude” lithics into the final MSA that lack readily

recognizable stylistic attributes

This techno-typological profile is consistent with other known assemblages from both

Mozambique‟s Niassa province and Malawi‟s Karonga district on the west side of Lake

Niassa. While the lithics from Niassa basin suggest a local character, it is difficult to recognize this as a formal local style because no truly diagnostic geographically constrained artifact types have been identified within the basin. However, they do fit into an overarching MSA technological complex that is found throughout the south-central

African woodlands and forests. The majority of sites known throughout this region show a preference for quartz in their lithic industries. Moreover, quartz is often reduced by expedient informal methods, but most assemblages also contain radial and Levallois cores similar to those seen in the Niassa basin. The tool components of assemblages found throughout the south-central African woodlands are also fairly consistent with the tools from Mvumu.

Mvumu‟s occupation at ~28-33kya makes it one of the youngest MSA lithic assemblages that has been studied anywhere in Africa. This demonstrates that the MSA persisted quite late in Niassa and provides evidence that the transition from MSA to LSA lithic technologies happened at different times in different parts of Africa. This date, combined with older dates from other Niassan sites that contain very similar lithic industries, suggests that MSA lithic technology in this region was characterized by long term continuity. It also warns that the “crudeness” seen in the Mvumu study assemblage,

181

and indeed throughout the Niassa basin on both sides of the lake, should not be misconstrued as an indicator of greater antiquity compared to more finely crafted lithic industries; despite its crude and archaic appearance lithic technology at Mvumu belongs to the final MSA.

Finally, by ~28-33kya all of the theories regarding the origins of modern behaviour would expect fully expressed modernity at Mvumu. However, many of the technological indicators of modern behaviour that relate to lithic technology are absent or poorly represented in the Mvumu study assemblage. For example, although there is a wide diversity of tool types represented at Mvumu, there is no evidence suggesting

Modes 4 and 5 were part of the Niassan technological repertoire and little evidence suggesting that any attempts at morphological standardization were made. Moreover, there is no indication of temporal variability or that raw material procurement networks extended beyond the immediate area at Mvumu (or other Niassan sites studied by PAC).

While this certainly does not mean that final MSA people in Niassa were any less

“modern” than their neighbours in other parts of Africa, it does highlight some of the shortcomings of the trait list approach in identifying modern behaviour in the archaeological record (Henshilwood and Marean 2003). Clearly there is still more to learn about concrete elements of MSA behavioural variability before we concern ourselves with abstract, poorly defined concepts such as modernity (Basell 2008; Shea

2011).

182

References

Adamowicz L (1987) Projecto "CIPRIANA", 1981-1985. Trabalhos de Arqueologia e Antropologia 3:47-144

Adamowicz L (1988) Contribuição para o registro computarizado das estações arqueológicas em Moçambique (Idade da Pedra Superior e Idade do Ferro). Trabalhos de Arqueologia e Antropologia 5:85-107

Adamowicz L (1990) The ecomonic and social context of early farming adaptation to highlands of northern Mozambique (Nampula, Cabo Delgado and Niassa). In: Sinclair P, Rakotoarisoa J (eds) Urban Origins in Eastern Africa: Proceedings of the 1989 Madagascar Workshop. Swedish Central Board of National Antiquities, Stockholm, pp 179-184

Afonso R, Marques J (1998) Recursos minerais da República de Moçambique, 2 edn. Direccção Nacional de Geologia, Maputo

Afonso R, Marques J, Ferrara M (1998) A evolução geológica de Moçambique: Uma síntese. Lisbon

Aiello L (1993) The fossil evidence for modern human origins in Africa; a revised view. American Anthropologist 95:73-96

Ambrose S (1998) Chronology of the Later Stone Age and food production in East Africa. Journal of Archaeological Science 25:377-392

Ambrose SH (2001) Paleolithic Technology and Human Evolution. Science 291:1748- 1753

Andrefsky W (1987) Diffusion and innovation from the perspective of wedge shaped cores in Alaska and . In: Johnson JK, Morrow CA (eds) The Organization of Core Technology. Westview Press, London, pp 13-44

Andrefsky W (1991) Inferring trends in prehistoric settlement behavior from lithic production technology in the Southern Plains. North American Archaeologist 12:129-144

Andrefsky W (1994) Raw-Material Availability and the Organization of Technology. American Antiquity 59(1):21-34

Andrefsky W (2005) Lithics: Macroscopic Approaches to Analysis, 2nd ed. Cambridge University Press, Cambridge

183

Armstrong AL (1931) Rhodesian archaeological expedition (1929): excavation in Bambata cave and researches on prehistoric sites in southern Rhodesia. Journal of the Royal Anthropological Institute of Britain and Ireland 61:239-276

Backwell L, d'Errico F (2000) A new functional interpretation of the early hominid bone tools. Journal of Human Evolution 38:A4-A5

Backwell L, d'Errico F (2001) Evidence of termite foraging by Swartkrans early hominids. Proceedings of the National Academy of Sciences 98(4):1358-1363

Backwell L, d'Errico F (2008) Early hominid bone tools from , South Africa. Journal of Archaeological Science 35:2880-2894

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

Bamforth DB (1990) Settlement, raw material, and lithic procurement in the Central Mohave Desert. Journal of Anthropological Archaeology 9:70-104

Barber RJ (ed) (1981) Quartz Technology in Prehistoric New England. Institute for Conservation Archaeology, Cambridge

Barham L (1987) The bipolar technique in Southern Africa: a replication experiment. The South African Archaeological Bulletin 42:45-50

Barham L (1995a) Making the most of a radial core: The topknot flake. Lithics 14:9-14

Barham L (1995b) The Project, Zambia, 1993-94. Nyame Akuma 43:66- 72

Barham L (1998) Possible early pigment use in South-Central Africa. Current Anthropology 39:703-710

Barham L (2000) The Middle Stone Age of Zambia, South Central Africa. Western Academic & Specialist Press, Bristol

Barham L (2002a) Backed Tools in Middle Pleistocene Central Africa and their Evolutionary Significance. Journal of Human Evolution 43:585-603

Barham L (2002b) Systematic pigment use in the Middle Pleistocene of South-Central Africa. Current Anthropology 43:181-190

184

Barham L, Smart P (1996) Early date for the Middle Stone Age of central Zambia. Journal of Human Evolution 30:287-290

Barker G, Barton H, Beavitt P, Bird M, Daly P, Doherty C, Gilbertson D, Hunt C, Krigbaum J, Lewis H, Manser J, McClaren S, Paz V, Piper P, Pyatt B, Rabett R, Reynolds TEG, Rose J, Rushworth G, Stephens M (2002) Prehistoric foragers and farmers in South-east Asia: renewed investigation at Niah Cave, Sarawak. Proceedings of the Prehistoric Society 68:147-164

Barker G, Barton H, Bird M, Daly P, Datan I, Dykes A, Farr L, Gilbertson D, Harrisson B, Hunt C, Higham T, Kealhofer L, Krigbaum J, Lewis H, McLaren S, Paz V, Pike A, Piper P, Pyatt P, Rabett R, Reynolds T, Rose J, Rushworth G, Stephens M, Stringer C, Thompson G, Turney C (2007) The „human revolution‟ in lowland tropical Southeast Asia: the antiquity and behavior of anatomically modern humans at Niah Cave (Sarawak, Borneo). Journal of Human Evolution 52:243-261

Barradas L (1949) Panorama of prehistory of Mozambique. South African Journal of Science 45:67-68

Barradas L (1962) O Quaternário do Antigo lago Lunho e da margem Portuguesa do lago Niassa. Direcção Nacional de Geologia, Maputo

Bar-Yosef O (1992) Middle Palaeolithic chronology and the transition to the Upper Palaeolithic in southwest Asia. In: Bräuer G, Smith F (eds) Continuity or replacement. A. A. Balkema, Rotterdam, pp 261-272

Bar-Yosef O (1998) On the nature of transitions: The Middle to Upper Palaeolithic and Neolithic Revolution. Cambridge Archaeological Journal 8(2):141-163

Bar-Yosef O (2002) The Upper Paleolithic Revolution. Annual Review of Anthropology 31:363-393

Bar-Yosef O, Kuhn S (1999) The big deal about blades: Laminar technologies and human evolution. American Anthropologist 101(2):322-338

Basell L (2008) Middle Stone Age (MSA) site distribution in eastern Africa and their relationship to Quaternary environmental change, refugia and the evolution of Homo sapiens. Quaternary Science Reviews 27:2484-2498

Bayle des Hermens R (1975) Recherches préhistoriques en République Centrafricaine. Labethno, Paris

Beaumont PB (1978) Border Cave. Unpublished M.A. thesis, University of Cape Town

185

Bennett T (n.d.) Preliminary report on Middle Stone Age lithics from Mikuyu, Mozambique. Unpublished report housed in the Tropical Archaeology Laboratory, University of Calgary, Calgary

Bennett T, Simpson S, Raja M, Mercader J (2010) Middle Stone Age lithic technology in Niassa, Mozambique. Paper presented at the Canadian Archaeological Association 2010 Annual Meeting, Calgary, AB

Bernal K, Brooks A (2010) Ex Africa semper aliquid novi: Implications of the early emergence of Levallois technology in Africa. Paleoanthropological Society Annual Meeting, St. Louis

Bettencourt Dias M (1948) Contribução para o estudos da Pré-história do Sul do Save. Boletim da Sociedade de Estudos da Colónia de Moçambique 13:19

Betzler C, Ring U (1995) Sedimentology of the Malawi Rift: facies and stratigraphy of the Chiwondo Beds, northern Malawi. Journal of Human Evolution 28:23-35

Bloesch U, Mbago F (2006) Selous-Niassa Wildlife Corridor: Vegetation study - Biodiversity, conservation values, and management strategies. The United Republic of Tanzania Ministry of Natural Resources and Tourism - Wildlife Division.

Boëda E (1995) Levallois: a volumetric construction. In: Dibble H, Bar-Yosef O (eds) The definition and Interpretation of Levallois technology. Prehistoric Press, Madison, pp 41-68

Bond G (1948) Rhodesian Stone Age man and his raw materials. South African Archaeological Bulletin 3(11):55-60

Bordes F (1961) Typologie du Paléolithique Ancien et Moyen. Delmas, Bordeaux

Borges A (1944) Estação Pré-histórica de Mangulane. Boletim da Sociedade de Estudos da Colónia de Moçambique

Boriskovsky PI, Soloview VV (1978) New data on the Stone Age of Guinée. West African Journal of Archaeology 8:51-74

Brain CK, Cooke CK (1967) A preliminary account of the Redcliff Stone Age Cave site in Rhodesia. South African Archaeological Bulletin 21:171-182

186

Bräuer G (1984) A craniological approach to the origin of anatomically modern Homo sapiens in Africaand implications for the appearance of modern Europeans. In: Smith F, Spencer F (eds) The Origins of Modern Humans. Alan R. Liss, New York, pp 327-410

Bräuer G (1989) The evolution of modern humans: A comparison of the African and non- African evidence. In: Mellars P, Stringer C (eds) The Human Revolution. Princeton University Press, Princeton, pp 123-154

Bräuer G (1992) Africa‟s place in the evolution of Homo sapiens. In: Bräuer G, Smith F (eds) Continuity or Replacement: Controversies in Homo sapiens Evolution. Balkema, Rotterdam, pp 83-98

Bräuer G (2001a) The „out of Africa‟ model and the question of regional continuity. In: Tobias P, Raath M, Moggi-Cecchi J, Doyle G (eds) Humanity from African naissance to coming millennia. Firenze University Press, Florence, pp 183-189

Bräuer G (2001b) The KRM-ER 3884 hominid and the emergence of modern anatomy in Africa. In: Tobias P, Raath M, Moggi-Cecchi J, Doyle G (eds) Humanity from the African naissance to coming millennia. Firenze University Press, Florence, pp 191-197

Bräuer G, Mehlman MJ (1988) Hominid molars form a Middle Stone Age level at the Mumba Rock Shelter, Tanzania. American Journal of Physical Anthropology 75:69-76

Brehm D (2007) Preliminary faunal analysis of anthropogenic bone modifications at Nankambe, Mozambique (Middle Stone Age). Unpublished report. Tropical Archaeology Laboratory, University of Calgary, Calgary

Breuil H (1944) Il parte das primeras impresses duma viagem de estudos arqueológicos no sul de Moçambique. Moçambique Documentário Trismestral 40:39-48

Bromage T, Schrenk F, Zonneveld F (1995) Paleoanthropology of the Malawi Rift: An early hominid mandible from the Chiwondo Beds, Northern Malawi. Journal of Human Evolution 28:71-108

Brooks A, Helgren D, Cramer J, Franklin A, Hornyak W, Keating J, Klein R, Rink W, Schwarcz H, Smith J, Stewart K, Todd N, Verniers J, Yellen J (1995) Dating and context of three Middle Stone Age sites with bone points in the Upper Semliki Valley, Zaire. Science 268:548-553

Brooks A, Nevell L, Yellen J, Hartman G (2006) Projectile technologies of the African MSA. In: Hovers E, Kuhn S (eds) Transitions Before the Transition. Springer, New York, pp 233-255

187

Brumm A, Moore MW, van den Bergh GD, Kurniawan I, Morwood MJ, Aziz F (2010) Stone technology at the Middle Pleistocene site of Mata Menge, Flores, . Journal of Archaeological Science 37:451-473

Cahen D (1978) New excavations at Gombe (ex-Kalina) Point, Kinshasa, Zaire. Antiquity 52:51-56

Campbell MC, Tishkoff SA (2008) African genetic diversity: Implications for human demographic history, modern human origins, and complex disease mapping. Annual Review of Genomics and Human Genetics 9:403-433

Campbell MC, Tishkoff SA (2010) The evolution of human genetic and phenotypic variation in Africa. Current Biology 20:R166-R173

Caramelli D, Laluez-Fox C, Vernesi C, Lari M, Casoli A, Mallegni F, Chiarelli B, Dupanloup I, Bertranpetit J, Barbujani G, Bertorelle G (2003) Evidence for a genetic discontinuity between Neandertals and 24,000 year old anatomically modern humans. Proceedings of the National Academy of Sciences 100(11):6493-6597

Carbonell E, Bermúdez de Castro J, Parés J, Pérez-González A, Cuenca-Bescós G, Ollé A, Mosquera M, Hughet R, Van der Made J, Rosas A, Sala R, Vallverdú J, García N, Granger D, Martinón-Torres M, Rodríguez X, Stock G, Vergès J, Allué E, Burjachs F, Cáceres I, Canals A, Benito A, Díez C, Lozano M, Mateos A, Navazo M, Rodrígues, J., Rosell J, Arsuaga JL (2008) The first hominin of Europe. Nature 452:465-469

Catuneau O, Wopfner H, Eriksson PG, Cairncross B, Rubidge BS, Smith RMH, Hancox PJ (2005) The Karoo basins of south-central Africa. Journal of African Earth Sciences 43:211-253

Chapola L, Kaphwiyo C (1992) The Malawi rift: geology, tectonics and seismicity. Tectonophysics 209:159-164

Chase P, Dibble H (1987) Middle Paleolithic symbolism: A review of current evidence and interpretations. Journal of Anthropological Archaeology 6:263-296

Chazan M (1995) Variability in the Middle Paleolithic of , Libya. Paléorient 21(1):65-73

Chazan M (1997) Redefining Levallois. Journal of Human Evolution 33:719-735

Chazan M, Ron H, Matmon A, Porat N, Goldberg P, Yates R, Avery M, Summer A, Horwitz L (2008) Radiometric dating of the Earlier Stone Age sequence in Excavation I

188

at , South Africa: preliminary results. Journal of Human Evolution 55:1-11

Clark AMB (1997a) The final Middle Stone Age at Rose Cottage Cave: A distinct industry in the Basutolian ecozone. South African Journal of Science 93:449-459

Clark AMB (1997b) The MSA/LSA transition in southern Africa: New technological evidence from Rose Cottage Cave. South African Archaeological Bulletin 52(113-121)

Clark G (1977) World Prehistory: A New Outline. Cambridge University Press, Cambridge

Clark JD (1962) The Kalambo Falls prehistoric site: an interim report. In: Mortelmans G, Nenquin J (eds) Actes du IV Concres Panafricain de Prehistoire et de l'Etude du Quaternaire, Series in -8 edn. Musee Royal de l'Afrique Centrale, Tervuren, pp 195-202

Clark JD (1963) Prehistoric Cultures of Northeast Angola and their Significance in Tropical Africa. Companhia de Diamantes de Angola (DIAMANG), Lisbon

Clark JD (1964a) The influence of environment inducing culture change at the Kalambo Falls prehistoric site. South African Archaeological Bulletin 19:93-101

Clark JD (1964b) The Sangoan Culture of Equatoria: the implications of its stone equipment. Miscellanea en Homenage al Abate Henri Breuil. Instituto de Prehistoria y Arquelogia, Monografias, Diputacion Provincial de Barcelona, pp 309-325

Clark JD (1965) The later Pleistocene Cultures of Africa. Science 150:833-847

Clark JD (1966) Initial investigation of the Archeology of Karonga District, Malawi. American Anthropologist 68:67-89

Clark JD (1969) The Kalambo Falls Prehistoric Site, Vol. 1: the Geology Palaeoecology, and Detailed Stratigraphy of the Excavations. Cambridge University Press, Cambridge

Clark JD (1970) The Prehistory of Africa. Praeger, New York

Clark JD (1972) Human behavioral differences in Southern Africa during the . American Anthropologist 73:1211-1236

Clark JD (1974) The Kalambo Falls Prehistoric Site, Vol II: The Later Pleistocene Cultures. Cambridge University Press, Cambridge

189

Clark JD (1975) A comparison of the late Acheulian industries of Africa and the Middle East. In: Butzer KW, Isaac G (eds) After the Australopithecines: Stratigraphy, Ecology, and Culture Change in the Midddle Pleistocene. The Hague, Mouton, pp 605-660

Clark JD (1982) The transition from Lower to Middle Palaeolithic in the African Continent. In: Ronen A (ed) The Transition from Lower to Middle Palaeolithic and the Origin of Modern Man International Symposium to commemorate the 50th anniversary of excavations in the Mount Carmel Caves by DAE Garrod University of Haifa October 1980. BAR International Series, 151

Clark JD (1988) The Middle Stone Age of East Africa and the Beginnings of Regional Identity. Journal of World Prehistory 2(3):235-305

Clark JD (2001a) The Chipeta industry of the Sangoan industrial complex: The archaeology of the Mkamba Member, Ochreous Sands Bed. In: Clark JD (ed) Kalambo Falls Prehistoric Site, Vol III The Earlier Cultures: Middle and Earlier Stone Age. Cambridge University Press, Cambridge, pp 234-285

Clark JD (ed) (2001b) Kalambo Falls Prehistoric Site, Vol. III. The Earlier Cultures: Middle and Earlier Stone Age. Cambridge University Press, Cambridge

Clark JD, Haynes CV (1970) An elephant butchery site at Mwanganda's village, Korongo, Malawi, and its relevance for Paleolithic archaeology. World Archaeology 1:390-411

Clark JD, Haynes CV, Mawby JE, Gautier A (1970) Interim report on palaeoanthropological investigations in the Lake Malawi Rift. Quaternaria 13:305-354

Clark JD, Williamson KD, Michaels JW, Marean CA (1984) A Middle Stone Age occupation site at Porc Epic cave, Dire Dawa (east-central Ethiopia). The African Archaeological Review 2:37-71

Clark JD, Brown K (2001) The Twin Rivers Kopje, Zambia: stratigraphy, fauna, and artefact assemblages from the 1954 and 1956 excavations. Journal of Archaeological Science 28:305-330

Clark JD, Kleindienst M (2001) The Stone Age Cultural Sequence: Terminology, Typology, and Raw Materials. In: Clark JD (ed) Kalambo Falls Prehistoric Site, Vol III. Cambridge University Press, Cambridge, pp 34-65

Clark JD, Beyene T, WoldeGabriel G, Hart WK, Renne PR, Gilbert H, Defleur A, Suwa G, Katoh S, Ludwig KR, Boisserie J-R, Asfaw B, White T (2003) Stratigraphic,

190

chronological, and behavioural contexts of Pleistocene Homo sapiens from Middle Awash, Ethiopia. Nature 423(6941):747-752

Cohen A, Stone J, Beuning K, Park L, Reinthal P, Dettman D, Scholz C, Johnson T, King J, Talbot M, Brown E, Ivory S (2007) Ecological consequences of early late Pleistocene megadroughts in tropical Africa. Proceedings of the National Academy of Sciences of the United States of America 104(42):16422-16427 Cole GH (1967) The Later Acheulian and Sangoan of southern Uganda. In: Bishop WW, Clark J (eds) Background to Evolution in Africa. University of Chicago Press, Chicago, pp 481-527

Conard N, Bolus M (2003) Radiocarbon dating and the appearance of modern humans and timing of cultural innovations in Europe: New results and new challenges. Journal of Human Evolution 44:332-371

Conkey M (1990) Experimenting with style in archaeology: Some historical and theoretical issues. In: Conkey M, Hastorf C (eds) The Uses of Style in Archaeology. Cambridge University Press, Cambridge, pp 5-17

Cooke CK (1962) The Sangoan industries of southern Rhodesia. South African Archaeological Bulletin 17(68):212-230

Cooke CK (1963) Report on excavations at Pomongwe and Tshangula Caves. South African Archaeological Bulletin 18:73-151

Cooke CK (1966) Re-appraisal of the industry hitherto named the Proto-Stillbay. Arnoldia 2(22):1-14

Coon C, Bricker H, Johnson F, Lamberg-Karlovsky C (1968) Yengema Cave Report. The University Museum, University of Pennsylvania, Philadelphia

Cornelissen E (1997) Central African transitional cultures. In: Vogel JC (ed) Encyclopedia of precolonial Africa. Altamira Press, Walnut Creek, pp 312-320

Cornelissen E (2002) Human responses to changing environments in Central Africa between 40,000 and 12,000 B.P. Journal of World Prehistory 16(3):197-235

Cornelissen E (2003) On microlithic quartz industries at the end of the Pleistocene in Central Africa: The evidence from (NW Cameroon). African Archaeological Review 20(1):1-24

Correia A (1934) As ruinas de Zimbabwe e a arqueologia da Moçambique. Trabalhos do I Congresso Nacional de Anthropologia Colonial 2:498-500

191

Cotterell B, Kamminga J (1990) Mechanics of Pre-Industrial Technology. Cambridge University Press, Cambridge

Cruz e Silva T (1976) A preliminary report on an Early Iron Age site: Matola IV/68. In: Morais J (ed) Iron Age Research in Moçambique, Report 2, p 33

Cruz e Silva T (1980) First indications of Early Iron Age in southern Moçambique: Matola IV 1/. In: Leakey R, Ogot BA (eds) VIIIth Panafrican Congress of Prehistory and Quaternary Studies, Nairobi, pp 349-350

Cruz-Uribe K (1983) The mammalian fauna from Redcliff Cave, Zimbabwe. South African Archaeological Bulletin 38:7-16

Dart RA (1960) The bone tool-manufacturing ability of Australopithecus prometheus. American Anthropologist 62(1):134-143

Day MH (1969) Omo human skeletal remains. Nature 222:1135-1138

Day MH, Twist MHC, Ward S (1991) Les vestiges post-craˆniens d‟Omo I (Kibish). L‟Anthropologie 95:595-610 de Freitas de Senna-Martinez JC (1968a) Paleollítico na Catembe. Boletim do Ceda 2:23 de Freitas de Senna-Martinez JC (1968b) Revisão das indústrias não-roladas do Movene IIA e IIB. Boletim do Ceda 2:25-26

Deacon H (1995) Two late Pleistocene - Holocene archaeological depositories from the Southern Cape, South Africa. South African Archaeological Bulletin 50:121-131

Debénath A (1994) L'Atérian du nord de l'Afrique du Sahara. Sahara 6:21-30

Debénath A, Dibble HL (1994) Handbook of Paleolithic Typology, vol. 1: Lower and Middle Paleolithic of Europe. University of Pennsylvania Press, Philadelphia

Delson E, Tattersall I, Van Couvering J, Brooks A (eds) (2000) Encyclopedia of Human Evolution and Prehistory, 2 edn. Garland Publishing, Inc., New York d'Errico F, Henshilwood C, Vanhaeren M, van Niekerk K (2005) Nassarius kraussianus shell beads from Blombos Cave: evidence for symbolic behaviour in the Middle Stone Age. Journal of Human Evolution 48(1):3-24

192

d'Errico F, Henshilwood CS (2007) Additional evidence for bone technology in the southern African Middle Stone Age. Journal of Human Evolution 52(2):142-163 d‟Errico F, Vanhaeren M, Wadley L (2008) Possible shell beads from the Middle Stone Age layers of Sibudu Cave, South Africa. Journal of Archaeological Science 35:2675– 2685

Derricourt R (1975) Some coastal shell middens in southern Moçambique. Azania 7:135- 139

Détroit F, Dizon E, Falguères C, Hameau S, Ronquillo W, Sémah F (2004) Upper Pleistocene Homo sapiens from Tabon Cave (Palawan, The Philippines): description and dating of new discoveries. Comptes Rendus Palevol 3:705-712

Dickson FP (1977) Quartz Flaking. In: Wright RVS (ed) Stone Tools as Cultural Markers: Change, Evolution, and Complexity. Australian Institute of Aboriginal Studies, Canberra, pp 97-103

Diez-Martín F, Domínguez-Rodrigo M, Sánchez P, Mabulla A, Tarriño A, Barba R, Prendergast M, de Luque L (2009) The Middle to Later Stone Age technological transition in East Africa. New data from Mumba Rockshelter Bed V (Tanzania) and their implications for the origin of modern human behavior. Journal of African Archaeology 7(2):1-27

Dixey F (1927) The Tertiary and post-Tertiary lacustrine sediments of the Nyassan Rift- valley. Quarterly Journal of the Geological Society of London 83:432-437

Dizon E, Détroit F, Sémah F, Hameau S, Ronquillo W, Cabanis E (2002) Notes on the morphology and age of the Tabon Cave fossil Homo sapiens. Current Anthropology 43:660-666

Domínguez-Rodrigo M, Díez-Martín F, Mabulla A, Luque L, Alcalá L, Tarriño A, Antonio López-Sáez J, Barba R, Bushozi P (2007) The archaeology of the Middle Pleistocene deposits of Lake Eyasi, Tanzania. Journal of Arfrican Archaeology 5(1):47- 78

Domínguez-Rodrigo M, Mabulla A, Luque L, Thompson JW, Rink J, Bushozi P, Díez- Martin F, Alcala L (2008) A new archaic Homo sapiens fossil from Lake Eyasi, Tanzania. Journal of Human Evolution 54:899-903

Duarte RT (1993) Northern Mozambique in the Swahili World, an Archaeological approach. Studies in African Archaeology 4 Repro HSC, Uppsala

193

Dunai T (2010) Cosmogenic Nuclides: Principles, Concepts and Applications in the Earth Surface Sciences. Cambridge University Press, Cambridge

Ekblom A (2004) Changing Landscapes: An Environmental History of Chibuene, Southern Mozambique. African and Comparative Archaeology, Department of Archaeology and Ancient History, Uppsala University, Uppsala

Eren M, Greenspan A, Sampson G (2008) Are Upper Paleolithic blade cores more productive than Middle Paleolithic discoidal cores? A replication experiment. Journal of Human Evolution 55:952-961

FAO (1998) World Reference base for soil resources. FAO, Rome

Finney BP, Scholz CA, Johnson TC, Trumbore S, J. S (1996) Late Quaternary lake-level changes of Lake Malawi. In: Johnson TC, Odada EO (eds) The Limnology, Climatology, and Paleoclimatology of the East African Lakes. Gordon & Breach, Toronto, pp 495-508

Foley R, Lahr MM (1997) More 3 technologies and the evolution of modern humans. Cambridge Archaeological Journal 7(1):3-36

Foley R, Lahr MM (2003) On stony ground: Lithic technology, human evolution, and the emergence of culture. Evolutionary Anthropology 12:109-122

Fox R (1970) The Tabon Cave. Monograph of the National Museum No. 1, Manila

Frison GC (1979) Observations on the use of stone tools: Dulling of working edges on some chipped stone tools in bison butchering. In: Hayden B (ed) Lithic Use-Wear Analysis. Academic Press, New York, pp 259-268

Frost P (1996) The ecology of Miombo woodlands. In: Campbell B (ed) The Miombo in transition: Woodlands and welfare in Africa. CIFOR, Bogor, pp 11-57

Gama M (1990) O Povo Yao. Instituto de Investigação Cientifica e Tropical, Lisbon

Garcea E (2004) Crossing deserts and avoiding seas: Aterian north African-European Relations. Journal of Anthopological Research 60(1):27-53

Garcea E, Giraudi C (2004) Late Quaternary human settlement patterning in the Jebel Gharbi. Journal of Human Evolution 51:411-421

Garlake P (1976) An investigation of Manekeni, Mozambique. Azania 11:25-47

194

Gibbon R, Granger D, Kuman K, Partridge T (2009) Early Acheulean technology in the Rieputs formation, South Africa, dated with cosmogenic nuclides. Journal of Human Evolution 56:152-160

Gifford D, Damrosch D, Pryor J, Thunen R (1985) The third dimension in site structure: an experiment in trampling and vertical dispersal. American Antiquity 50:803-818

Gonder MK, Mortensen HM, Reed FA, de Sousa A, Tishkoff SA (2007) Whole-mtDNA Genome Sequence Analysis of Ancient African Lineages. Molecular Biology and Evolution 24(3):757-768

Goodwin A (1928) An introduction to the Middle Stone Age in South Africa. The South African Journal of Science 25:410-418

Goodwin A, Van Riet Lowe C (1929) The Stone Age cultures of South Africa. AMS Press, New York

Gosse J, Phillips F (2001) Terrestrial in situ cosmogenic nuclides: theory and applications. Quaternary Science Reviews(20):1475-1560

Granger DE, Muzikar P (2001) Dating sediment burial with cosmogenic nuclides: Theory, techniques, and limitations. Earth and Planetary Science Letters 188(1-2):269- 281

Green RE, Krause J, Briggs AW, Maricic T, Stenzel U, Kircher M, Patterson N, Li H, Zhai W, Fritz MH-Y, Hansen NF, Durand EY, Malaspinas A-S, Jensen JD, Marques- Bonet T, Alkan C, Prüfer K, Meyer M, Burbano HA, Good JM, Schultz R, Aximu-Petri A, Butthof A, Höber B, Höffner B, Siegemund M, Weihmann A, Nusbaum C, Eric S. Lander ES, Russ C, Novod N, Affourtit J, Egholm M, Verna C, Rudan P, Brajkovic D, Kucan Ž, Gušic I, Doronichev VB, Golovanova LV, Lalueza-Fox C, de la Rasilla M, Fortea J, Rosas A, Schmitz RW, Johnson PLF, Eichler EE, Falush D, Birney E, Mullikin JC, Slatkin M, Nielsen R, Kelso J, Lachmann M, Reich D, Svante Pääbo S (2010) A Draft Sequence of the Neandertal Genome. Science 328:710-722

Gresham TH, Brandt SA (1996) Variability in the MSA of the Horn of Africa. In: Pwiti G, Soper R (eds) Aspects of African Archaeology. University of Zimbabwe Publications, Harare

Grine F (2000) Middle Stone Age human fossils from Die Kelders Cave 1, Western Cape Province, South Africa. Journal of Human Evolution 38:129-145

Grün R (2006) The chronology of Nankambe Cave, Mozambique. Unpublished report. Tropical Archaeology Laboratory, University of Calgary, Calgary

195

Grün R, Stringer C (1991) Electron spin resonance dating and the evolution of modern humans. Archaeometry 33(2):153-199

Grün R, Beaumont PB (2001) Border Cave revisited: A revised ESR chronology. Journal of Human Evolution 40:467-482

Harrisson T (1975) Tampan: ‟s palaeolithic reconsidered. Modern Quaternary Research in SE Asia 1:53-69

Hawkins A, Kleindienst M (2002) Lithic raw material usages during the Middle Stone Age at Dakhleh Oasis, Egypt. Geoarchaeology 17(6):601-624

Henry DO (1989) Correlations between reduction strategies and settlement patterns. In: Henry DO, Odell GH (eds) Alternative Approaches to Lithic Analysis. Westview Press, Boulder, pp 139-212

Henshilwood C, Sealy J (1997) Bone Artefacts from the Middle Stone Age at Blombos Cave, Southern Cape, South Africa. Current Anthropology 38:890-895

Henshilwood C, D'Errico F, Marean CW, Milo RG, Yates RJ (2001a) An Early Bone Tool Industry from the Middle Stone Age, Blombos Cave, South Africa: Implications for the Origins of Modern Human Behavior, Symbolism, and Language. Journal of Human Evolution 41:631-678

Henshilwood C, Sealy J, Yates R, Cruz-Uribe, Goldberg P, Grine F, Klein R, Poggenpoel C, Niekerk V, Watts I (2001b) Blombos Cave, Southern Cape, South Africa: preliminary report on the 1992-1999 excavations of the Middle Stone Age levels. Journal of Archaeological Science 28:421-448

Henshilwood C, d'Errico F, Vanhaeren M, Van Niekerk K, Jacobs Z (2004) Middle Stone Age Shell Beads from South Africa. Science 304:304

Henshilwood C, d'Errico F, Yates R, Jacobs Z, Tribolo C, Duller G, Mercier N, Sealy J, Valladas H, Watts I, Wintle A (2002) Emergence of Modern Human Behavior: Middle Stone Age Engravings from South Africa. Science 295:1278-1280

Henshilwood C, Marean CW (2003) The Origin of Modern Human Behavior: A Review and Critique of Models and Test Implications. Current Anthropology 44:627-651

Hiscock P (1982) A technical analysis of quartz assemblages from the South Coast. In: Bowder S (ed) Coastal Archaeology in Eastern Australia. Australian National University Press, Canberra, pp 32-45

196

Hiscock P (1996) Mobility and technology in the Kakadu Coastal Wetlands. Indo-Pacific Prehistory Assocciation Bulletin 15:151-157

Hiscock P (2003) Early Australian Implement Variation: A Reduction Model. Journal of Archaeological Science 30:239-249

Howells W (1976) Explaining modern man: Evolutionists versus migrationists. Journal of Human Evolution 5(5):477-495

Ingman M, Kaessmann H, S. P, Gyllenstein U (2000) Mitochondrial genome variation and the African origin of modern humans. Nature 408(6813):708-713

Inizan ML, Roche H, Tixier J (1992) Technology of Knapped Stone. CREP, Nanterre

Inizan ML, Reduron-Ballinger M, Roche H, Tixier J (1999) Technology and Terminology of Knapped Stone. CREP, Nanterre

Isaac G, Isaac B (1977) Olorgesailie: archeological studies of a Middle Pleistocene lake basin in Kenya. University of Chicago Press, Chicago

Instituto Nacional de Investigação Agronómica (1995) Legenda da Carta Nacional de Solos. Instituto Nacional de Investigação Agronómica, Maputo

Jacobs Z, Wintle A, Duller G, Roberts R, Wadley L (2008) New ages for the post- Howiesons Poort, late and final Middle Stone Age at Sibudu, South Africa. Journal of Archaeological Science 35:1790-1807

Johnson C, McBrearty S (2010) 500,000 year old blades from the Kapthurin Formation, Kenya. Journal of Human Evolution 58:193-200

Juwayeyi Y, Betzler C (1995) Archaeology of the Malawi Rift: the search continues for Early Stone Age occurrences in the Chiwondo Beds, northern Malawi. Journal of Human Evolution 28(1):115-116

Kalb JE, Jaegar M, Jolly CJ, Kana B (1982) Preliminary geology, paleontology and paleoecology of a Sangoan site at Andalee, Middle Awash Valley, Ethiopia. Journal of Archaeological Science 9:349-363

Kamminga J (1982) Over the Edge: Functional Analysis of Australian Stone Tools. Occational Papers in Anthropology 12

197

Kaplan J (1990) The Umhlatuzana Rock Shelter sequence: 100,000 years of Stone Age history. Natal Museum Journal of Humanities 2:1-94

Kaufulu Z (1990) Sedimentary environments at the Mwanganda site, Malawi. Geoarchaeology 5(1):15-27

Kaufulu Z, Stern N (1987) The first stone artefacts to be found in situ within the Plio- Pleistocene Chiwondo beds in northern Malawi. Journal of Human Evolution 16:729-740

Klatzow S (1994) Roosfontein, a contact site in the Eastern Orange Free State. South African Archaeological Bulletin 49(159):9-15

Klein RG (1995) Anatomy, behavior, and modern human origins. Journal of World Prehistory 9:167-198

Klein RG (1999) The Human Career, 2 edn. University of Chicago Press, Chicago

Klein RG (2000) Archaeology and the evolution of modern human behavior. Evolutionary Anthropology 9:17-36

Klein RG (2001) Southern Africa and Modern Human Origins. Journal of Anthopological Research 57(1):1-16

Klein RG (2009) The Human Career, 3 edn. University of Chicago Press, Chicago Klein R, Cruz-Uribe K (1996) Exploitation of large bovids and seals at Middle and Later Stone Age sites in South Africa. Journal of Human Evolution 31:315-334

Klein RG, Scott K (1986) Re-analysis of faunal assemblages from the Haua Fteah and other Late Quaternary archaeological sites in Cyrenaican Libya. Journal of Archaeological Science 13:515-542

Kleindienst M (2000) On the Nile corridor and the Out of Africa model. Current Anthropology 41:107-109

Knight J (1991) Vein Quartz. Lithics 12:37-56

Knowles FHS, Barnes AS (1937) Manufacture of Gun-Flints. Antiquity 11:201-207

Knutsson K (1988) Making and using stone tools: The analysis of the lithic assemblages from Middle Neolithic sites within the Västerbotten, northern Sweden. Societas Archaeologica Uppsaliensis, Uppsala

198

Knutsson K, Lindgren C (2009) Making sense of quartz: Presentation and results of an experimental analysis applied to quartz from a number of sites in Södertöm. In: Akerlund A, Olsson E, Gustavsson P, Miller U (eds). Swedish National Heritage Board, Hägersten, pp 5-36

Kohl-Larsen L (1943) Auf den Spuren des Vormenschen. Strecker und Schröder, Stuttgart

Kooyman BP (2000) Understanding Stone Tools and Archaeological Sites. University of Calgary Press, Calgary

Kuhn S (1995) Mousterian Lithic Technology. Princeton University Press, Princeton

Kuhn S (2007) Cores, tools, and the priorities of lithic analysis. In: McPherron SJP (ed) Tools Verus Cores. Cambridge Scholars Publishing, Newcastle, pp 267-276

Kuman K (2001) An Acheulean factory site with prepared core technology near Taung, South Africa. South African Archaeological Bulletin 56(173/174):8-22

Kuman K (1998) The earliest South African industries. In: M. P, Korisettar R (eds) Early Human Behaviour in Global Context: The Rise and Diversity of the Lower Palaeolithic Record. Routledge, New York, pp 151-162

Kuman K, Baron JL, Gibbon R (2005) Earlier Stone Age archaeology of the Vhembe- Dongola National park (South Africa) and vicinity. Quaternary International 129:23-32

Kuman K, Inbar M, Clarke RJ (1999) Palaeoenvironments and cultural sequence of the Florisbad Middle Stone Age Hominid Site, South Africa. Journal of Archaeological Science 26:1409-1425

Lächelt S (2004) Geology and mineral resources of Mozambique. Direcção Nacional de Geologia, Maputo

Lahr MM, Foley R (1994) Multiple dispersals and modern human origins. Evolutionary Anthropology 3(2):48-60

Lahr MM, Foley R (1998) Towards a theory of modern human origins: Geography, demography, and diversity in recent human evolution. Yearbook of Physical Anthropology 41:137-176

Lanfranchi R, Schwartz D (1990) Évolution des paysages de la Sangha (R.P. du Congo) au Pleistocène supérieur. Bilan des observations archéologiques, géomorphologiques,

199

pédologiques et paléobiologiques. In: Lanfranchi R, Schwartz D (eds) Paysages Quaternaires du l'Afrique centrale Atlantique. Orstom, Paris, pp 248-259

Lanfranchi R (1996) Une industrie M.S.A. de stone-line en forêt dense: le site de Mokeko (Congo). In: Pwiti G, Soper R (eds) Aspects of African Archaeology. University of Zimbabwe, Harare, pp 166-175

Leakey MD (1971) Olduvai Gorge Vol. 3: Excavations in Beds I and II, 1960 - 1963. Cambridge University Press, Cambridge

Leakey MD, Tobias PV, Martyn JE, Leakey REF (1969) An Acheulian industry with prepared core technique and the discovery of a contemporary hominid mandible at Lake Baringo, Kenya. Proceedings of the Prehistoric Society 35:48-76

Lebatard A-E, Bourlès DL, Duringer P, Jolivet M, Braucher R, Carcaillet J, Arnaud N, Monié P, Lihoreau F, Likius A, Mackaye HT, Vignaud P, Brunet M (2008) Cosmogenic nuclide dating of Sahelanthropus tchadensis and Australopithecus bahrelghazali: Mio- Pliocene hominids from Chad PNAS 105(9):3226-3231

Lioubine V, Guede F (2000) The Paleolithic of Republic Côte d'Ivoire (West Africa). In Russian with an English summary. Russian Academy of Sciences, St. Petersburg

Liu H, Prugnolle F, Manica A, Balloux F (2006) A geographically explicit genetic model of worldwide human-settlement history. American Journal of Human Genetics 79:230- 237

Lombard M (2005) Evidence of Hunting and Hafting during the Middle Stone Age at Sibidu Cave, KwaZulu-Natal, South Africa: a Multianalytical Approach. Journal of Human Evolution 48:279-300

Lombard M (2006a) First impressions of the functions and hafting technology of Still Bay pointed artefacts from Sibudu Cave. Southern African Humanities 18(1):27-41

Lombard M (2007) The gripping nature of ochre: The association of ochre with Howiesons Poort adhesives and Later Stone Age mastics from South Africa. Journal of Human Evolution 53:406-419

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

200

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, Phillipson L (2010) Indications of bow and stone-tipped arrow use 64000 years ago in KwaZulu-Natal, South Africa. Antiquity 84:635-648

Macamo S (2006) Privileged Places in South Central Mozambique: The Archaeology of Manyikeni, Niamara, Songo, and Degue-Mufa. African and Comparative Archaeology, Department of Archaeology and Ancient History, Uppsala University, Uppsala

Macamo S, Saetersdal T (2004) Archaeology and cultural heritage management in Mozambique - Some experiences made and some fresh challenges. In: Oestigaard T, Anfinset N, Saetersdal T (eds) Combining the Past and Present: Archaeological Perspectives on Society. BAR International, Oxford

Macaulay V, Hill C, Achilli A, Rengo C, Clarke D, Meehan W, Blackburn J, Semino O, Scozzari R, Cruciani F, Taha A, Shaari NK, Raja JM, Ismail P, Zainuddin Z, Goodwin W, Bulbeck D, Bandelt H-J, Oppenheimer S, Torroni A, Richards M (2005) Single, rapid coastal settlement of Asia revealed by analysis of complete mitochondrial genomes. Science 308:1034-1036

MacCalman HR, Viereck A (1967) Peperkorrel, a factory site of Lupemban affinities from central south West Africa. The South African Archaeological Bulletin 22:41-50

Mackay A, Welz A (2008) Engraved ochre from a Middle Stone Age context at Klein Kliphuis in the Western Cape of South Africa. Journal of Archaeological Science 35:1521-1532

Madiquida H (2006) The Iron Age Communities in the Zambezi River Basin: Excavations in Mozambique. SAfA, Calgary, Canada

Madiquida H (2007) The Iron-Using Communities of the Cape Delgado Coast from AD 1000. Studies in Global Archaeology 8, Uppsala Universitet

Marean C, Bar-Mathews M, Bernatchez J, Fisher E, Goldberg P, Herries A, Jacobs Z, Jerardino A, Karkanas P, Minichillo T, Nilssen P, Thompson E, Watts I, Williams H (2007) Early human use of marine resources and pigment in South Africa during the Middle Pleistocene. Nature 449:905-909

Marks A, Conard N (2007) Technology vs. typology: The case for and against a transition from the MSA to the LSA at , Tanzania. In: Aubry T, Almeida F, Araújo AC, Tiffagom M (eds) Space and Time: Which Diachronies, Which Synchronies,

201

Which Scales? Typology vs Technology. BAR International Series 1831, Archaeopress, Oxford

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 D (ed) Cultural beginnings Approaches to understanding early hominid life-ways in the African savanna. Romisch-Germanisches Zentralmuseum, Bonn, pp 159-176

McBrearty S (1992) Sangoan Technology and Habitat at Simbi. Nyame Akuma 38:34-40

McBrearty S, Bishop L, Kingston J (1996) Variability in traces of Middle Pleistocene hominid behavior in the Kapthurin Formation, Barino, Kenya. Journal of Human Evolution 30:563-580

McBrearty S, Bishop L, Plummer T, Dewar R, Conard N (1998) Tools underfoot: human trampling as an agent of the lithic artifact edge modification. American Antiquity 63:108- 129

McBrearty S, Brooks A (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 (2006) From Acheulean to Middle Stone Age in the Kapthurin Formation, Kenya. In: Hovers E, Kuhn S (eds) Transitions Before the Transition. Springer, New York, pp 257-277

McDougall I, Brown FH, Fleagle JG (2005) Stratigraphic placement and age of modern humans from Kibish, Ethiopia. Nature 433(7027):733-736

McDougall I, Brown FH, Fleagle JG (2008) Sapropels and the age of hominins Omo I and II, Kibish, Ethiopia. journal of Human Evolution 55:409-420

Medeiros E (1997) História de Cabo Delgado e do Niassa (c. 1836-1929). Central Impressora, Maputo

Mehlman M (1979) Mumba-Höhle revisited: the relevance of a forgotten excavation to some current issues in East African prehistory. World Archaeology 11:80-94

Mehlman M (1987) Provenience, Age and Associations of Archaic Homo sapiens Crania from Lake Eyasi, Tanzania. Journal of Archaeological Science 14:133-162

202

Mehlman M (1989) Late Quaternary archaeological sequences in northern Tanzania. Unpublished Ph.D dissertation, University of Illinois, Urbana

Mellars P (2005) The Impossible Coincidence. A Single-Species Model for the Origins of Modern Human Behavior in Europe. Evolutionary Anthropology 14:12-27

Mellars P (1996) The Neanderthal legacy. An archaeological perspective from Western Europe. Princeton University Press, Princeton

Meneses P (1988) Idade da Pedra em Moçambique. Trabalhos de Arqueologia e Antropologia 5:3-49

Meneses P (1992) O mapa arqueologico de Moçambique (Idade da Pedra). Leba 7:221- 234

Meneses P (1996) Some aspects regarding the Acheulean in southern Mozambique. In: Pwiti G, Soper R (eds) Aspects of African Archaeology: Papers from the Tenth Congress of the Pan-African Association for Prehistory and Related Studies. University of Zimbabwe Press, Harare, pp 129-138

Meneses P (1999) New methodological approaches to the study of the Acheulean from Southern Mozambique. PhD dissertation, Department of Anthropology. Rutgers, New Brunswick

Meneses P (2004) O Achaulense no Sul de Moçambique: Novas Abordagens Metodológicas. PROMÉDIA, Maputo

Mercader J (2009) Mozambican grass seed consumption during the Middle Stone Age. Science 326(5960):1680-1683

Mercader J (2010) Public archaeology (or community service) in Mozambique. Chacmool Conference, University of Calgary

Mercader J, Brooks A (2001) Across Forests and Savannas: Later Stone Age Assemblages from Ituri and Semliki, Northeast Democratic Republic of Congo. Journal of Anthropological Research 57(2):197-217

Mercader J, Martí R (1999) Archaeology in the tropical forest of Banyang-Mbo, SW Cameroon. Nyame Akuma 52:17-24

Mercader J, Martí R, Martínez J, Brooks A (2002) The Nature of 'Stone-Lines' in the African Quaternary Record: Archaeological Resolution at the Rainforest Site of Mosumu, Equatorial Guinea. Quaternary International 89:71-96

203

Mercader J, Martí R (2003) The hunter-gatherer occupation of Atlantic central Africa: New evidence from Equatorial Guinea Cameroon. In: Mercader J (ed) Under the Canopy: the Archaeology of Tropical Rainforests. Rutgers University Press, pp 64-92

Mercader J, Fogelman A (2006) Between east and west: First Middle Stone Age in northern Mozambique. Paleoanthropology Society Annual Meeting, San Juan, Puerto Rico

Mercader J, Bennett T, Raja M (2008) Middle Stone Age Starch Acquisition in the Niassa Rift, Mozambique. Quaternary Research 70:283-300

Mercader J, Asmerom Y, Bennett T, Raja M, Skinner A (2009a) Initial excavation and dating of Ngalue Cave: a Middle Stone Age site along the Niassa Rift, Mozambique. Journal of Human Evolution 57(1):63-74

Mercader J, Bennett T, Esselmont C, Simpson S, Walde D (2009b) Phytoliths in woody plants from the Miombo woodlands of Mozambique. Annals of Botany 104(1):91-113

Mercader J, Astudillo F, Barkworth M, Bennett T, Esselmont C, Kinyanjui R, Laskin Grossman D, Simpson S, Walde D (2010) Poaceae phytoliths from the Niassa Rift, Mozambique. Journal of Archaeological Science 37:1953-1967

Mercader J, Bennett T, Esselmont C, Simpson S, Walde D (2011) Soil Phytoliths from Miombo Woodlands in Mozambique Quaternary Research 75:138-150

Mercader J, Bennett T, Gosse J, Raja M (in prep) Mvumu: A Final Middle Stone Age site from Niassa, Mozambique

Mijares AS, Détroit F, Piper P, Grün R, Bellwood P, Aubert M, Champion G, Cuevas N, De Leon A, Dizon E (2010) New evidence for a 67,000-year-old human presence at Callao Cave, Luzon, Philippines. Journal of Human Evolution 59:123-132

Miller GH, Beaument PB, Deacon HJ, Brooks AS, Hare PE, Jull AJT (1999) Earliest modern humans in southern Africa dated by isoleucine epimerization in ostrich eggshell. Quaternary Scinence Reviews 18:1537-1548

Milo R (1998) Evidence for hominid predation at Klasies River Mouth, South Africa, and its implications for the behaviour of early modern humans. Journal of Archaeological Science 25:99-133

204

Mitchell P (1994) Understanding the MSA/LSA transitions: the pre-20,000 BP assemblages from new excavations at Segonghong rock shelter, . South African Field Archaeology 3:15-25

Mitchell P (2002) The Archaeology of Southern Africa. Cambridge University Press, Cambridge

Mithen S (1996) The prehistory of the mind: A search for the origins of art, religion, and science. Thames and Hudson, London

Moore MW, Sutikna T, Jatmiko, Morwood M, Brumm A (2009) Continuities in stone flaking technology at , Flores, Indonesia. Journal of Human Evolution 57:503- 526

Morais J (1984) Mozambican Archaeology: Past and Present. African Archaeological Review 2:113-128

Morais J (1988) The Early Farming Communities of Southern Mozambique. Eduardo Mondlane University, Maputo

Müller-Beck H (ed) (1978) Die Archäologischen und Anthropologischen Ergebnisse der Kohl-Larsen-Expeditionen in Nord-Tanzania 1933-1939. Tübingen Monographien zur Urgeschichte 4 (2), Tübingen

Nenquin J (1967) Contribution to the Study of the Prehistoric Cultures of Rwanda and Burundi, Tervuren

Nespoulet R, El Hajraoui MA, Amani F, Ben Ncer A, Debénath A, El Idrissi A, Lacombe J-P, Michel P, Oujaa A, Stoetzel E (2008) Palaeolithic and Neolithic occupations in the Témara Region (Rabat, Morocco): Recent data on hominin contexts and behavior. African Archaeological Review 25:21-39

Ngunga A (ed) (2009) Lexicografia e Descrição de Línguas Bantu. Imprensa Universitária, Maputo

Ngunga A, Nhongo N, Moisés L, Langa J, H. C, Mucavele J (2010) Educação Bilingue na Província de Gaza: Avaliação de um Modelo de Ensino. Imprensa Universitária, Maputo

Norwegian Geological Survey (2004) Map of Metangula, Macaloge, Chiconono at 1:250,000

205

Nygaard SE, Talbot R (1984) Stone Age archeology and environment on the southern Accra plains, Ghana. Norwegian Archaeological Review 17(1):19-38

Omar LL, António A (20004) As Dinastias Mataaka (Séc. XVIII-XX). ARPAC - Instituto de Investigação Sócio-Cultural, Maputo

Omi G (1977) Prehistoric sites and implements in Cameroon. An annex to the Interim Report of the Tropical African Geomorphology Research Project 1975 / 76., Nagoya, Japan

Opperman H (1996) Strathalan cave B, North-Easterns Cape Province, South Africa: Evidence for human behaviour 29,000-26,000 years ago. Quaternary International 33:45- 53

Opperman H, Heydenrych B (1990) A 22000 year-old Middle Stone Age camp site with plant food remains from the North-Eastern Cape. South African Archaeological Bulletin 45:93-99

Parry WJ, Kelly RL (1987) Expedient core technology and sedentism. In: Johnson JK, Morrow CA (eds) The Organization of Core Technology. Westview Press, Boulder, pp 285-304

Patterson LW (1983) Criteria for determining the attributes of man-made lithics. Journal of Field Archaeology 10(3):297-307

Pearson OM (2008) Statistical and biological definitions of „„anatomically modern‟‟ humans: Suggestions for a unified approach to modern morphology. Evolutionary Anthropology 17:38-48

Pearson OM, Royer DF, Grine FE, Fleagle JG (2008) A description of the Omo I postcranial skeleton, including newly discovered fossils. Journal of Human Evolution 55:421-437

Petraglia M, Korisettar R, Boivin N, Clarkson C, Ditchfield P, Jones S, Koshy J, Lahr MM, Oppenheimer C, Pyle D, Roberts R, Schwenninger J-L, Arnold L, White K (2007) Middle Paleolithic assemblages from the Indian Subcontinent before and after the Toba super-eruption. Science 317:114-116

Phillipson DW (1976) The Prehistory of Eastern Zambia. Memoir 6. British Institute in Eastern Africa, Nairobi

Phillipson DW (1993) African Archaeology, 2nd edn. Cambridge University Press, Cambridge

206

Phillipson L (2007a) Reassessment of selected Middle Stone Age artefacts from Rhino Cave and from White Paintings Rock Shelter, Hills, Botswana. South African Archaeological Bulletin 185:19-30

Phillipson L (2007b) A Replication Study of Lithic Point and Tool Production. South African Archaeological Bulletin 62(185):31-43

Pleurdeau DW (2005) Human technical behavior in the African Middle Stone Age: the lithic assemblage of Porc-Epic Cave (Dire Dawa, Ethiopia). African Archaeological Review 22(4):177-197

Pommeret Y (1966) Civilisations prehistoriques au Gabon, Tome 2; Vallee du Moyen Ogooue: presentation de l'industrie lithiques de traditions sangoenne, lupembienne et neolithique. Centre culturel Francaise Saint Exupery, Libreville

Powell A, Shennan S, Thomas MG (2009) Late Pleistocene demography and the appearance of modern human behavior. Science 324:1298-1301

Prendergast M, Luque L, Domínguez-Rodrigo M, Diez-Martín F, Mabulla A, Barba R (2007) New excavations at Mumba Rockshelter. Journal of African Archaeology 5(2):217-243

Raja M (2008) Estação arqueológica de Mikuyu: Evidências da ocupação humana na margem centro oriental do Lago Niassa na Idade da Pedra Média. Departmento de História. Universidade Edduardo Monddlane, Maputo

Reed FA, Tishkoff SA (2006) African human diversity, origins and migrations. Current Opinion in Genetics and Development 16:597-605

Reid I, Frostick E (1986) Slope processes, sediment derivation and landform evolution in a rift valley basin, Northern Kenya. Geological Society Special Publication 25:99-111

Relethford J, Jorde L (1999) Genetic evidence for larger African population size during recent human evolution. American Journal of Physical Anthropology 108:251-260

Renethford J (1999) Models, predictions and the fossil record of modern human origins. Evolutionary Anthropology 8:7-10

Renfrew C, Bahn P (2008) Archaeology: Theories, Methods and Practice. Thames and Hudson, London

207

Rightmire GP, Deacon HJ (1991) Comparative studies of Late Pleistocene human remains from Klasies River Mouth, South Africa. Journal of Human Evolution 20(2):131-156

Robbins L, Murphy M, Stewart K, Campbell A, Brooks G (1994) Barbed bone points, paleoenvironment and the antiquity of fish exploitation in the Kalahari desert, Botswana. Journal of Field Archaeology 21:257-264

Robbins L, Murphy M, Brook G, Ivester A, Campbell A, Klein R, Milo R, Brooks A, Kokis J, Downey W (2000) Archaeology, palaeoenvironment, and chronology of the Tsodilo Hills White Paintings Rock Shelter, Northwest Kalahari Desert, Botswana. Journal of Archaeological Science 27:1085-1113

Robert A, Soriano S, Rasse M, Stokes S, Huysecom E (2003) First Chrono-Cultural Reference Framework for the West African Paleolithic: New Data From Ounjougou, Dogon Country, Mali. Journal of African Archaeology 1(2):151-169

Rodrigues MC (2006a) O primeiro síto com vestígios de utilização do ferro e cerâmica "tradicional" da Early Iron Age localizado em Moçambique - província de Zambézia. Revista Portuguesa de Arqueologia 9(2):415-449

Rodrigues MC (2006b) A primeira cerâmica "tradicional recente" proveniente de Tete (província de Tete, Moçambique. Revista Portuguesa de Arqueologia 9(1):197-223

Rodrigues MC (2007) Artefactos líticos das “estações” dos terraços fluviais da margem sul do rio Zambeze – provincial de Tete, Moçambique: uma primeira abordagem. revista Portuguesa de Arqueologia 10(1):401-482

Roller O (1954) Die Steingeräte der Mumba-Höhle. Ein Beitrag zur Kenntnis der Vorgeschichte Ostafrikas. Unpublished Ph.D. dissertation, University of Tübingen

Rose JI (2004) New evidence for the expansion of an Upper Pleistocene population out of East Africa, from the site of Station One, Northern Sudan. Cambridge Archaeological Journal 14(2):205-216

Rots V, Van Peer P (2006) Early evidence of complexity in lithic economy: Core-axe production, hafting, and use at Late Middle Pleistocene site 8-B-11, Sai Island (Sudan). Journal of Archaeological Science 33:360-371

Rougier H, Milota S, Rodrigo R, Gherase M, Sarcina L, Moldovan O, Zilhao J, Constantin S, Franciscus RG, Zollikofer CPE, Ponce de Leon M, Trinkaus E (2007) Pestera cu Oase 2 and the cranial morphology of early modern Europeans. Proceedings of the National Academy of Sciences 104(4):1165-1170

208

Russell A, Jamniczky H (2006) Faunal identifications from Nankambe Cave, Mozambique. Unpublished report. Tropical Archaeology Laboratory, University of Calgary, Calgary

Sackett JR (1977) The meaning of style in archaeology: A general model. American Antiquity 42(3):369-380

Sackett JR (1982) Approaches to style in lithic archaeology. Journal of Anthopological Archaeology 1:59-112

Sackett JR (1990) Style and ethnicity in archaeology: A case for isochrestism. In: Conkey M, Hastorf C (eds) The Uses of Style in Archaeology. Cambridge University Press, Cambridge, pp 32-43

Saetersdal T (2003) Archaeological research and cultural heritage management in Mozambique: Report on rock art survey, Angónia district 2003. Direcção nacional do patrimóio cultural, Maputo

Salman G, Abdula E (1995) Development of the Mozambique and Rovuma sedimentary basins, offshore Mozambique. Sedimentary Geology 96(Special Monograph)

Sampson CG (1974) The Stone Age Archaeology of Southern Africa. Academic Press, New York

Santos Júnior JR (1941) On the prehistory of Moçambique. Moçambique Documentário Trismestral 28:23-88

Schick K, Toth N (1993) Making silent stones speak. Human evolution and the dawn of technology. Simon and Schuster, New York

Scholz CA, Cohen AS, King J, Johnson TC, Lyons RP, Talbot MR (2006) Scientific Drilling on Lake Malawi: A Long, High-Resolution Record of Climatic and Limnological Change from Southern Tropical East Africa. 2006 Philadelphia Annual Meeting

Seong C (2004) Quartzite and vein quartz as lithic raw materials reconsidered: A view from the Korean Paleolithic. Asian Perspectives 43(1):73-91

Serra C (ed) (2000) História de Moçambique. Imprensa Universitária, Maputo

Serre D, Langaney A, Chech M, Teschler-Nicola M, Paunovic M, Mennecler P, Hofreiter M, Possnert G, Pääbo S (2004) No evidence of Neandertal mtDNA contribution to early modern humans. Public Library of Science Biology 2(3):313-317

209

Shea J (2003) Neandertals, competition, and the Origins of modern human behavior in the Levant. Evolutionary Anthropology 12:173-187

Shea J (2006) The origins of projectile point technology: evidence from Africa, the Levant, and Europe. Journal of Archaeological Science 33:823-846

Shea 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 55:488-485

Shea J (2009) The impact of projectile weaponry on Late Pleistocene hominin evolution. In: Hublin J-J, Richards MP (eds) The Evolution of Hominin Diets: Integrating Approaches to the Study of Palaeolithic Subsistence Springer, pp 189-199

Shea J (2011) Homo sapiens Is as Homo sapiens Was. Current Anthropology 52(1):1-35

Shen G, Gao X, Gao B, Granger D (2009) Age of Zhoukoudian Homo erectus determined with 26Al/10Be burial dating. Nature 458:198-200

Sheppard PJ, Kleindienst MR (1996) Technological change in the Earlier and Middle Stone Age of Kalambo Falls (Zambia). African Archaeological Review 13(3):171-195

Siame LL, Bourlès DL, Brown ET (eds) (2006) In Situ Produced Cosmogenic Nuclides and Quantification of Geological Processes. Geological Society of America Special Paper 415. Geological Society of America Boulder

Simões M (1951) A préhistoria de Moçambique. Boletim da Sociedade de Estudos da Colónia de Moçambique 68:115-152

Simões M (1958) Sequência da Evolução da Indústria da Pedra em Moçambique a Sul do Save. Boletim da Sociedade de Estudos da Colónia de Moçambique (n.3):149-156

Simpson S, Bennett T, Eddie J, Murchie T, Oakden N, Swan W, Mercader J (2010) Usewear analysis of experimental quartz tools and their archaeological counterparts from Mozambique. Canadian Archaeological Association 2010 Annual Meeting, Calgary, AB

Sinclair P (1982) Chibuene: An early trading site in southern Mozambique. Paideuma 28:149-164

Sinclair P (1985) An Archaeological reconnaissance of Northern Mozambique. Part I: Nampula province. Working papers in African Studies 14

210

Sinclair P (1986) An Archaeological reconnaissance of Northern Mozambique. Part II: Cabo Delgado province. Working papers in African Studies 14

Sinclair P, Nydorf N-G, Wickman-Nydorf G (1987) Excavations at the University Campus Maputo, Mozambique 1984-85. Central Board of Antiquities, Stockholm

Sinclair P, Morais J, Adamowicz L, Duarte R (1993) A perspective on archaeological research in Mozambique. In: Shaw T, Sinclair P, Okpoko A (eds) The Archaeology of Africa: Food, Metals, and Towns. Routledge, London, pp 409-431

Singer R, Wymer J (1982) The Middle Stone Age at Klasies River Mouth in South Africa. University of Chicago Press, Chicago

Sisk ML, Shea J (2009) Experimental use and quantitative performance analysis of triangular flakes (Levallois points) used as arrowheads. Journal of Archaeological Science 36:2039-2047

Skinner A, Blackwell B, Long R, Seronie-Vivien MR, Tillier A-M, Blickstein J (2007) New ESR dates for a new bone-bearing layer at Pradayrol, Lot, . Paleoanthropology Society Annual Meeting, Philadelphia, PA

Smolla G, Korfmann M (n.d.) Unpublished report on Mozambican archaeology (1970).

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

Steele TE, Klein RG (2009) Late Pleistocene subsistence strategies and resource intensification in Africa. In: Hublin J-J, Richards MP (eds) The Evolution of Hominin Diets: Integrating Approaches to the Study of Palaeolithic Subsistence Springer, pp 113- 126

Stringer C (2001) Modern Human Origins – Distinguishing the Models. African Archaeological Review 18(2):67-75

Stringer C (2003) Out of Ethiopia. Nature 423(6941):692-694

Stringer C (2007) The origin and dispersal of Homo sapiens: our current state of knowledge. In: Mellars P, Boyle K, Bar-Yosef O, Stringer C (eds) Rethinking the Human Revolution McDonald Institute for Archaeological Research, Cambridge, pp 15-20

Stringer C, Andrews P (1988a) Genetic and fossil evidence for the origin of modern humans. Science 239(4845):1263-1268

211

Stringer C, Andrews P (1988b) Modern human origins. Science 241(4867):773-774

Stringer C, Gamble C (1993) In search of Neanderthals: Solving the puzzles of human origins. Thames and Hudson, London

Tallavaara M, Manninen MA, Hertell E, Rankama T (2010) How flakes shatter: a critical evaluation of quartz fracture analysis. Journal of Archaeological Science 37:2442-2448

Templeton A (2002) Out of Africa again and again. Nature 416(6876):45-51

Thackeray AI (1989) Changing fashions in the Middle Stone Age: the stone artefact sequence from Klasies River Main Site, South Africa. The African Archaeological Review 7:33-57

Thackeray AI (1992) The Middle Stone Age of the Limpopo River. Journal of World Prehistory 6:385-440

Thackeray A (2000) Middle Stone Age artefacts from the 1993 and 1995 excavations of Die Kelders Cave 1, South Africa. Journal of Human Evolution 38:147-168

Thompson J (2010) The Malawi earlier-Middle Stone Age project: Preliminary results from 2009 fieldwork. Paleoanthropological Society Annual Meeting, St. Louis

Thorne A, Wolpoff M (1992) The multiregional evolution of modern humans. Scientific American 266:76-83

Tishkoff SA, Reed FA, Friedlaender FR, Ehret C, Ranciaro A, Froment A, Hirbo jB, Awomoyi AA, Bodo J-M, Doumbo O, Ibrahim M, Juma AT, Kotze MJ, Lema G, Moore JH, Mortensen H, Nyambo TB, Omar SA, Powell K, Pretorius GS, Smith MW, Thera MA, Wambebe C, L. W, Williams SM (2009) The Genetic Structure and History of Africans and African Americans. Science 22:1035-1044

Torrence R (1989) Time, Energy and Stone Tools. Cambridge University Press, New York

Trapani J (2008) Quaternary fossil fish from the Kibish Formation, Omo Valley, Ethiopia. Journal of Human Evolution 55:521-530

Trigger BG (2006) A History of Archaeological Thought, 2 edn. Cambridge University Pess, Cambridge

212

Tryon CA (2006) "Early" Middle Stone Age lithic technology of the Kapthurin Formation (Kenya). Current Anthropology 47(2):367-375

Tryon CA, 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(4):199-229

Tryon C, Roach N, Logan M (2008) The Middle Stone Age of the northern Kenya Rift: age and context of new archaeological sites from the Kapedo Tuffs. Journal of Human Evolution 55:652-664

Valladas H, Reyss J, Joron J, Valladas G, Bar-Yosef O (1988) Thermoluminescence dating of Mousterian 'proto-Cro-Magnon' remains from Israel and the origin of modern man. Nature 331(6157):614-616

Van Moorsel H (1970) Recherches préhistoriques au pays de l'entre-fleuves Lukenie- Kasai. Etudes d'Histoire Africaine 1:7-36

Van Noten F, Haesaerts P, Cahen D (1972) Un habitat lupembien a Masango, Burundi. Rapport préliminaire. Africa Tervuren 18:78-85

Van Noten F (1982) The Archaeology of Central Africa. Akademische Druck unds, Verlagsanstalt Graz,

Van Peer P (1991) Interassemblage variability and Levallois styles: The case of the Northern African Middle Palaeolithic. Journal of Anthropological Archaeology 10:107- 151

Van Peer P (1992) The Levallois reduction strategy. Prehistory Press, Madison

Van Peer P (1998) The Nile corridor and the Out of Africa model: An examination of the archaeological record. Current Anthropology 39:S115-S140

Van Peer P, Fullagar R, Stokes S, Bailey R, Moeyersons J, Steenhoudt F, Geerts A, Vanderbeken M, 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, Fullagar R, Stokes S, Bailey R, Moeyersons J, Steenhoudt F, Geerts A, Vanderbeken M, 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

213

van Riet Lowe C (1943) A contribution to the prehistory of Moçambique. Moçambique Documentário Trismestral 36:1-5 van Riet Lowe C (1945) The evolution of the Levallois technique in South Africa. Man 45:49-59

Vermeersch P, Paulissen E, Stokes S, Charlier C, Van Peer P, Stringer C, Lindsay W (1998) Middle Palaeolithic burial of a modern human at Taramsa Hill, Egypt. Antiquity 72:475-484

Verniers J, Paulis R, Frasca S, Jourdan P (1978) Relatório da brigada geológica de cartografía da bacia carbonífera de Maniamba (DINAGECA). DINAGECA, Maputo

Vigilant L, Stoneking M, Harpending H, Hawkes K, Wilson A (1991) African populations and the evolution of human mitochondrial DNA. Science 253(5027):1503- 1507

Villa P, Delagnes A, Wadley L (2005) A late Middle Stone Age artifact assemblage from Sibudu (Kwazulu-Natal): comparisons with the European Middle Palaeolithic. 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(1):89-122

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

Volman TP (1984) Early prehistory of southern Africa. In: Klein R (ed) Southern African Prehistory and Palaeoenvironments, pp 169-219

Wadley L (1992) Rose Cottage Cave: The Later Stone Age levels with European and Iron Age artefacts. South African Archaeological Bulletin 47(155):8-12

Wadley L (1993) The Pleistocene Later Stone Age south of the Limpopo River. Journal of World Prehistory 7:243-296

Wadley L (1997) Rose Cottage Cave: archaeological work 1987-1997. South African Journal of Science 93:439-444

Wadley L (2001) What is cultural modernity? A general view and a South African perspective from Rose Cottage Cave. Cambridge Archaeological Journal 11(2):201-221

214

Wadley L, Jacobs Z (2004) Sibudu Cave, KwaZulu-Natal: Background to the excavations of Middle Stone Age and Iron Age occupations. South African Journal of Science 100:145-151

Wadley L, Williamson B, Lombard M (2004) Ochre in hafting in Middle Stone Age southern Africa: a practical role. Antiquity 78(301):661-675

Wadley L (2005) A typological study of the Final Middle Stone Age stone tools from Sibudu Cave, Kwazulu-Natal. The South African Archaeological Bulletin 60(182):51-63

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

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 Science 106(19):9590-9594

Watson E, Forster P, Richards M, Bandelt H-J (1997) Mitochondrial footprints of human expansion in Africa. American Journal of Human Genetics 61:691-704

Watts I (2002) Ochre in the Middle Stone Age of Southern Africa: Ritualised display or hide preservative? The South African Archaeological Bulletin 57(175):1-14

Wayland EG (1934) Rifts, rivers, rains, and early man in Uganda. The Journal of the Royal Anthropological Institute of Great Britain and Ireland 64:333-352

Wayland EG (1935) The M horizon: A result of climatic oscillation in the second pluvial period. Bulletin of the Geological Survey of Uganda 2:68-76

Wayland EG, Smith R (1923) Some primitive stone tools from Uganda. Geological Survey of Uganda Occasional Paper 1

Weidenreich F (1937) The relation of Sinanthropus pekinensis to Pithecanthropus, Javanthropus, and Rhodesian man. Journal of the Royal Anthropological Institure 67:51- 65

Wells LH (1943) Relatório sobre objectos encontrados em restos de cozinha perto da foz do Limpopo. Documentário Trimestral 36:13-23

Wendorf F, Schild R (1974) A Middle Stone Age sequence from the Central Rift Valley, Ethiopia. Instytut Historii Materialnej, Warsaw

215

Wendorf F, Schild R (1992) The Middle Paleolithic of North Africa: A status report. In: Klees F, Kuper R (eds) New Light on the Northeast African Past. Heinrich - Barth - Institut, Köln, pp 40-78

Wendorf F, Schild R (1993) Probable functions of Mousterian points and convergent sidescrapers in the Middle Stone Age of Ethiopia. Quaternaria Nova 3:39-51

Wendorf F, Close AE, Schild R (1987) Recent Work on the Middle Palaeolithic of the Eastern Sahara. The African Archaeological Review 5:49-63

Wendt WE (1976) Art mobilier from the , southwest Africa: Africa's oldest dated works of art. South African Archaeological Bulletin 31:5-11

White JP (1967) Ethno-Archaeology in New Guinea. Mankind 6:409-414

White JP, Thomas DH (1972) What mean these stones? Ethno-taxonomic models and archaeological interpretations in the New Guinea highlands. In: Clarke D (ed) Models in Archaeology. Methuen, London, pp 275-308

White F (ed) (1983) The Vegetation of Africa: A descriptive memoir to accompany the Unesco/Aetfat/Unso vegetation map of Africa. Unesco, Paris

White T, Asfaw B, DeGusta D, Gilbert H, Richards G, Suwa G, Howell FC (2003) Pleistocene Homo sapiens from Middle Awash, Ethiopia. Nature 423:742-747

Whittaker JC (1994) Flintknapping: Making and Understanding Stone Tools. University of Texas Press, Austin

Wieschoff HA (1941) The Zimbabwe-Monomotapa Culture. General Studies in Anthropology, Wisconsin, pp 1-115

Wilkins JR (2008) Prepared core reduction strategies at Kudu Koppie and the modern human behaviour debate. Unpublished MA Thesis. Department of Archaeology. University of Calgary, Calgary

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

Williams H (2005) Technology and behaviour in the Middle Stone Age: an analysis of the lithic assemblage from Katanda 16, Democratic Republic of Congo (DRC). Unpublished MA Thesis. Department of Anthropology. George Washington University, Washington D.C.

216

Willoughby P (1996) Middle Stone Age technology and adaptation in Southwestern Tanzania. In: Pwiti G, Soper R (eds) Aspects of African Archaeology. University of Zimbabwe Press, Harare, pp 177-189

Willoughby P (2001) Middle and Later Stone Age technology from the lake Rukwa, Southwestern Tanzania. South African Archaeological Bulletin 56:34-45

Willoughby P, Sipe C (2002) Stone Age Prehistory of the Songwe River valley, Lake Rukwa basin, southwestern Tanzania. African Archaeological Review 19(4):203-221 Wolpoff M (1996) Interpretations of multiregional evolution. Science 274:704-707

Wolpoff M (1989) Multiregional evolution: The fossil alternative to Eden. In: Mellars P, Stringer C (eds) The Emergence of Modern Humans. Cambridge University Press, Cambridge, pp 97-141

Wolpoff M, Spuhler J, Smith F, Radovčić J, Pope G, Frayer D, Eckhardt R, Clark G (1988) Modern human origins. Science 241:772-774

Wolpoff M, Relethford J (1997) Population size and structure, and models of human evolution. Abstracts Cold Springs Harbor Meeting on Human Evolution, Cold Springs Harbor, New York, p 89

Wolpoff M, Hawks J, Caspari R (2000) Multiregional, not multiple origins. American Journal of Physical Anthropology 112:129-136

Wurz S (1999) The Howiesons Poort backed artefacts from Klasies River: An argument for symbolic behaviour. South African Archaeological Bulletin 54:38-50

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 (1996) The evolution of tools and symbolic behaviour. In: Lock A, Peters CR (eds) Handbook of Human Symbolic Evolution. Clarendon Press, Oxford

Yellen J (1996) Behavioural and taphonomic patterning at Katanda 9: a Middle Stone Age site, Kivu province, Zaire. Journal of Archaeological Science 23:915-932

Yellen J (1998) Barbed bone points: Tradition and continuity in Saharan and sub-Saharan Africa. African Archaeological Review 15:173-198

217

Yellen J, Brooks A, Cornelissen E, Mehlman M, Stewart K (1995) A Middle Stone Age worked bone industry from Katanda, Upper Semliki Valley, Zaire. Science 268:553-556

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 Middle Awash Vally, Ethiopia. PaleoAnthropology 10:25- 100

Zimba B, Alpers E, Issacman A (eds) (2005) Slave Routes and Oral Tradition in Southeastern Africa. Filsom, Maputo