<<

A Geoarchaeological Investigation of Storage and Surplus at Tel Tsaf,

by

Emily Marie Hubbard

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

© Copyright by Emily M. Hubbard 2015

A Geological Investigation of Storage and Surplus at Tel Tsaf, Israel

Emily Marie Hubbard

Doctor of Philosophy

Department of Anthropology University of Toronto

2015 Abstract

The period of the southern spans from c. 5800 - 3600 cal BC. Traditional hallmarks of this period include caches of copper items, non-domestic shrine sites and large off site cemeteries or burial caves. However, these features date to the later portion of the period

(Late Chalcolithic) and date to c. 4400 cal BC onwards. The focus of this dissertation is on the

Middle Chalcolithic period (c. 5040 – 4440 cal BC), specifically in relation to storage and access to resources at the village of Tel Tsaf, Israel.

Excavations at Tel Tsaf uncovered evidence of at least four walled courtyard complexes spanning two Chalcolithic phases. Each courtyard contained either rectangular broadroom houses or circular buildings, with no evidence of both structural forms in the same courtyard.

Each courtyard also contained multiple large, -lined circular silos. The research goals of this work address the implications of these varied architectural forms and the possible identification of surplus of both floral and faunal goods. Microscopic sediment analysis was used to investigate these questions, and the results are discussed in relation to the big picture debates of the Chalcolithic period.

ii

Micromorphology and bulk sediment samples were taken from Building Complexes I, II and IV during the 2006 and 2007 field seasons. Results demonstrate that domestic activities centred in the broadroom house and courtyard of Building Complex I, and animal penning activity was restricted to the circular buildings in Building Complex II. This distribution, combined with the numerous large silos, indicate there may have been some economic specialization at Tel Tsaf.

The relationship among storage, surplus and wealth is explored and put in context of public and private access to resources and potential feasting at the site.

It is established that the evidence is not significant enough to classify Tel Tsaf or the Middle

Chalcolithic societies as chiefdoms. The evidence of likely surplus and potential private ownership of animals without evidence of a central distribution system or regional hierarchy further differentiates Tel Tsaf from large Late Chalcolithic villages. However, the evidence from

Tel Tsaf demonstrates that it is distinct from the Early and Late Chalcolithic sites, demonstrating some similarity to both periods, as well as some evidence unique to the Middle Chalcolithic.

iii

Acknowledgments

I am indebted to a great number of people for their support during the creation of this dissertation. My sincere appreciation goes out to my advisor, Ted Banning, for his patience and guidance during this process. His insight and feedback was spot on, always identifying the areas I knew were weakest and offering suggestions to help put me back on track. I am also incredibly grateful to Lisa Maher, a late committee addition who was invaluable. Her readiness to have meetings at the drop of a hat, help me formulate ideas and structures, and provide prompt feedback were essential. Her endless optimism was also appreciated when my own waned! My final committee member, Michael Chazan, also provided excellent feedback and guidance. It was his encouragement in the early days of my PhD that set me on this specific path, and I am grateful it did. Many thanks to my external examiner, Arlene Rosen, for providing insightful feedback and constructive criticism during the defense process.

I owe a great deal of gratitude to Wendy Matthews and the Department of Archaeology at the University of Reading. The micromorphology analysis would not have been possible without Wendy’s willingness to take me on as a visiting student and provide me with a strong foundation. I would also like to acknowledge Yosef Garfinkel for allowing me to participate in the Tel Tsaf excavations and take samples from the site. His support of my analysis allowed me to collect extensive samples throughout the site.

Throughout the creation of this dissertation I received access to facilities at a number of institutions, both formally and informally. I would like to acknowledge the University of Bristol, the University of Leeds and the University of California, Berkeley. The Archaeology Research Facility at UC Berkeley has been especially welcoming, providing me with access to the facilities necessary to complete my research.

None of this would have been possible without the support of my many colleagues. At Toronto, talks and meetings with Jennifer Campbell, Emma Humphrey, Lesley Howse, Trevor Orchard, Terry Clark, Andrew Riddle, Kevin Gibbs and Adam Allentuck were essential in both academic and stress-management ways! At Reading, a huge thanks to Carol Mansfield, Kay Lakin and Alex Brown, all of whom not only shared advice and ideas, but spare rooms during my visits! A great field crew is also essential to good field seasons, and I am grateful to Chad Hill for two

iv years of encouragement during heat waves and early mornings. Discussion of data and ideas with Chad was essential in my dissertation! I was also a pleasure to have work with Kelly Sayers, Kit Atwood, Mirah Burgener and Jennifer Hood at Tel Tsaf.

A huge thanks to my family. My parents have always been supportive and I thank them for encouraging me to follow my dreams. I could not have made it this far without their example of hard work, kindness and perseverance. I am also lucky to have married into fantastic in-laws who have been patient and loving through this process. Last, and certainly not least, I am indebted to my patient, patient husband, Chris. He knew he was marrying my PhD dissertation when he got into this, but neither of us expected it to be with us for so long! Despite his career making this an adventure (immigrating twice!!), his support, love and encouragement have seen me through it all! We will both be pleased to hang the degree on the wall and move forward!

v

Table of Contents

Acknowledgments ...... iv

Table of Contents ...... vi

List of Tables ...... x

List of Figures ...... xiii

List of Appendices ...... xix

Chapter 1 Introduction ...... 1

1.1 Research Objectives ...... 1

1.2 Summary of Chapters ...... 5

Chapter 2 Background Chapter ...... 7

2.1 Chronology ...... 7

2.2 Overview of Excavation History ...... 14

2.3 Early Chalcolithic (Wadi Rabah and variants) ...... 19

2.4 The Middle Chalcolithic ...... 20

2.5 The Late Chalcolithic Period ...... 22

2.5.1 Material Culture ...... 23

2.5.2 Burial Practices ...... 26

2.5.3 Architecture ...... 27

2.6 Beyond the ...... 28

2.7 Social Organization ...... 29

2.8 Tel Tsaf ...... 33

2.8.1 Regional Geology and Paleoclimate ...... 34

2.8.2 History of Excavation ...... 41

2.8.3 Area B ...... 41

2.8.4 Area C ...... 43

vi

2.8.4.1 Building Complex I ...... 47

2.8.4.2 Building Complex II ...... 50

2.8.4.3 Building Complex IV ...... 51

2.8.5 Summary of Excavations ...... 53

Chapter 3 Methods and Results ...... 55

3.1 Geoarchaeological Assessment and Site Formation ...... 55

3.2 Sediment pH ...... 55

3.3 Micromorphology ...... 57

3.3.1 Sampling Strategy ...... 58

3.3.2 Sample Extraction and Processing ...... 59

3.3.3 Sample Components ...... 61

3.3.4 Microfacies ...... 73

3.3.5 Building Complex I – Phases 3 and 4 ...... 79

3.3.5.1 Room 70 – Rectangular Room ...... 80

3.3.5.2 Courtyard ...... 84

3.3.5.3 Room 70 – Wall samples ...... 88

3.3.5.4 Silo 339 – Chalcolithic Burial Pit ...... 90

3.3.6 Building Complex II - Phase 3 ...... 92

3.3.6.1 Room 230 – Circular Structure ...... 93

3.3.6.2 Room 268 – Circular Structure ...... 97

3.3.6.3 Silo 272 ...... 98

3.3.6.4 Byzantine Test Pit (North East corner of excavation area) ...... 101

3.3.7 Building Complex IV – Phase 4 ...... 103

3.3.7.1 Room 662 ...... 103

3.3.7.2 Room 612 ...... 105

3.3.7.3 Southern Courtyard ...... 109 vii

3.3.7.4 Oven Feature, Courtyard ...... 111

3.4 Summary of Results ...... 113

3.5 Bulk Samples ...... 114

3.5.1 Room Interiors ...... 126

3.5.2 Courtyards ...... 129

3.5.3 Silos ...... 133

3.5.4 Roasting Pits ...... 135

3.5.5 Potential Animal Pens ...... 136

3.6 Chapter Summary ...... 139

Chapter 4 A Life History of Tel Tsaf ...... 147

4.1 Site Formation Processes ...... 147

4.2 Silos ...... 151

4.3 Livestock ...... 155

4.4 Cooking and Manufacturing ...... 159

4.5 Architectural changes between phases ...... 161

4.5.1 Building Complex I – Continuity ...... 161

4.5.2 Building Complex IV to Building Complex II - Why the change? ...... 164

4.6 Tel Tsaf during the Middle Chalcolithic ...... 165

Chapter 5 Storage, Surplus and Wealth ...... 167

5.1 Architecture and Archaeology ...... 167

5.1.1 Building Complex I ...... 171

5.1.2 Building Complex II ...... 174

5.1.3 Interpreting the Evidence ...... 177

5.2 Surplus and Wealth ...... 178

5.2.1 Tel Tsaf Silos Re-visited ...... 184

5.2.2 Animals at Tel Tsaf ...... 191 viii

5.3 Control of Resources – Public vs. Private ...... 192

5.4 Power, Conspicuous Consumption and Feasting ...... 195

Chapter 6 The Great Chiefdom Debate ...... 199

6.1 The Case for Chiefdoms in the Southern Levant ...... 199

6.1.1 Distinct Regional Groups ...... 201

6.1.2 Greater Population and Density ...... 201

6.1.3 Centres for Coordinating Social, Religious and Economic Activity ...... 202

6.1.3.1 Economic Centres ...... 202

6.1.3.2 Social and Ritual Centralization ...... 203

6.1.4 Defined Territorial Boundaries ...... 206

6.1.5 Greater Productivity ...... 207

6.2 All Hail the Chief at Tel Tsaf? ...... 208

6.3 Alternative Frameworks ...... 212

Chapter 7 Conclusion ...... 215

7.1 Clarification of Site-Formation Processes ...... 215

7.2 What are the implications of varied architectural forms during a single phase of occupation? ...... 216

7.3 With evidence of silos and animal pens at Tel Tsaf, is it possible to identify meaningful surplus? If so, what are its social and economic implications? ...... 217

7.4 How can the results of microscopic sediment analysis provide data that directly contributes to big-picture debates regarding the possibility of chiefdoms during the Chalcolithic? ...... 219

7.5 Tel Tsaf and the Middle Chalcolithic in Context ...... 220

References ...... 229

Appendices ...... 271

ix

List of Tables

Table 2.1: A general overview of current schemes regarding the classification of periods and cultures in and around the Chalcolithic in the Near East...... 10

Table 2.2: List of radiocarbon determinations used for the analysis...... 13

Table 2.3: Comparison of results for the calibrated boundaries between Early Chalcolithic (Wadi Rabah), Middle and Late Chalcolithic periods ...... 14

Table 2.4: Major cultural periods of the areas immediately surrounding the Southern Levant. . 28

Table 2.5: Renfrew’s key indicators of Chiefdoms within societies (1973)...... 31

Table 2.6: Overview of phases for Area C at Tel Tsaf as observed by Garfinkel et al. (2007a, b; 2009)...... 44

Table 3.1: Overview of Tel Tsaf occupation phases and Geoarchaeological Observations ...... 57

Table 3.2: Soil pH results from Tel Tsaf...... 57

Table 3.3 Main components found in the Tel Tsaf micromorphology samples...... 68

Table 3.4 Microfacies found in the Tel Tsaf micromorphology samples...... 78

Table 3.5: An overview of micromorphology samples from Building Complex I, Room 70 ..... 80

Table 3.6 An overview of micromorphology samples from the Building Complex I Courtyard...... 84

Table 3.7: An overview of micromorphology samples from Building Complex I wall surfaces...... 88

Table 3.8: An overview of micromorphology samples from the interior of Silo 339...... 90

Table 3.9: An overview of micromorphology samples from Building Complex II, Room 230.. 93

x

Table 3.10: An overview of micromorphology samples from Building Complex II, Room 268...... 97

Table 3.11: An overview of micromorphology samples from Building Complex II/IV, Silo 272...... 98

Table 3.12: An overview of micromorphology samples from the Byzantine Surface, north of Building Complex II...... 101

Table 3.13: An overview of micromorphology samples from Building Complex IV, Room 662...... 103

Table 3.14: An overview of micromorphology samples from Building Complex IV, Room 612...... 105

Table 3.15: An overview of micromorphology samples from the Courtyard south of Building Complex IV...... 109

Table 3.16: An overview of micromorphology samples from an Oven in the courtyard south of Building Complex IV...... 111

Table 3.17: Examples of densities of fecal spherulites and phytoliths from Tel Tsaf samples.. 118

Table 3.18: Bulk sample results by context...... 120

Table 3.19: Relative abundance of fecal spherulite and phytolith densities from all interior room contexts at Tel Tsaf...... 126

Table 3.20: Relative abundance of fecal spherulites and phytoliths from all courtyard contexts at Tel Tsaf...... 129

Table 3.21: Relative abundance of fecal spherulites and phytoliths from all silo contexts at Tel Tsaf...... 133

Table 3.22: Relative abundance of fecal spherulites and phytoliths from all roasting pit contexts at Tel Tsaf...... 135

xi

Table 3.23: Relative abundance of fecal spherulites and phytoliths from all animal pen contexts at Tel Tsaf...... 137

Table 3.24: An overview of the main contexts identified at Tel Tsaf and corresponding micromorphology and bulk sample evidence...... 145

Table 4.1: Number and dimensions of the silos from Phases 3 and 4 at Tel Tsaf...... 154

Table 4.2: Overview of silo area by building complex and phase...... 155

Table 5.1: Differences between Building Complex I and Building Complex II...... 176

Table 5.2: The number and dimensions of the silos from Phases 3 and 4 at Tel Tsaf...... 184

Table 5.3: Grain storage capacity in metric tons for all Phase 3 silos in Building Complexes I and II...... 186

Table 5.4: Modern grain densities per cubic metre...... 186

Table 5.5: Estimated silo volume and potential storage capacity for Building I, Phase 3 at Tel Tsaf...... 190

Table 5.6: A breakdown of roasting pits within Building I, Phase 3 with physical description and breakdown of faunal remains found within each (data from Ben-Shlomo, et al. 2009; Hill 2010) ...... 197

Table 6.1: Overview of the proposed evidence for chiefdom level organization in the Late Chalcolithic of the Southern Levant...... 200

Table 6.2: A summary of copper items found at Chalcolithic sites in the southern Levant...... 204

Table 6.3: Evidence for a chiefdom societal structure at Tel Tsaf based on the indicators outlined in Table 6.1 ...... 208

Table 7.1: An overview of the archaeological evidence from the Early Chalcolithic, Tel Tsaf and the Late Chalcolithic within the Southern Levant...... 221

xii

List of Figures

Figure 2.1: The Radiocarbon calibration curve shown as a 1σ band for the Chalcolithic period in the Southern Levant...... 12

Figure 2.2: Map of Chalcolithic sites in the southern Levant...... 15

Figure 2.3: An example of the distinctive Tsafian design on a sherd from Tel Tsaf...... 21

Figure 2.4: Examples of copper objects from the Chalcolithic...... 25

Figure 2.5: Clay ossuaries from Azor ...... 26

Figure 2.6: The location of Tel Tsaf...... 33

Figure 2.7: Map outlining the three hills of Tel Tsaf...... 33

Figure 2.8: Geological map of the Northern Valley around Tel Tsaf ...... 35

Figure 2.9: An exposed cross-section of the Lisan Formation, north of Beth Shean in the ...... 36

Figure 2.10: Area B well...... 42

Figure 2.11: Phase 3 at Tel Tsaf...... 45

Figure 2.12: Phase 4 at Tel Tsaf...... 46

Figure 2.13: Site plan of the area exposed during the 2005 excavation...... 47

Figure 2.14: Arial view of the eastern courtyard of Building Complex I...... 48

Figure 2.15: Roasting pit from Building I, Phase 3...... 48

Figure 2.16: Silo 339, Building I...... 49

Figure 2.17: Burial from Silo 339...... 50

Figure 2.18: Final excavation photo of Building Complex II, Phase 3...... 51

xiii

Figure 2.19: Final excavation photo of Building Complex II, Phase 4...... 52

Figure 2.20: The area directly east of Silo 272, Building IV ...... 53

Figure 2.21: Burn level located beneath the excavated boundary of room 612...... 53

Figure 3.1: The extent of the mud-brick spill is not clear, nor are the boundaries of Rooms 612 and 662...... 54

Figure 3.2: Room 70, Building Complex I...... 55

Figure 3.3: Plans of Area C, Phase 3 and 4 at Tel Tsaf ...... 56

Figure 3.4: Building Complex II, Room 230...... 60

Figure 3.5: Building Complex II, Room 230...... 60

Figure 3.6: Regional map of Tel Tsaf with location of micromorphology samples...... 62

Figure 3.7: Water-laid, calcareous clays from the Lisan Formation near Tel Tsaf...... 62

Figure 3.8 Building Complex I, Phases 4 and 3...... 80

Figure 3.9 Building Complex I, Room 70, baulk highlighted by white lines ...... 81

Figure 3.10: Room 70, Building Complex I...... 82

Figure 3.11: Room 70, Building Complex I...... 83

Figure 3.12 Building Complex I, Courtyard...... 85

Figure 3.13 Building Complex I, Courtyard...... 86

Figure 3.14: Building Complex I, Courtyard...... 87

Figure 3.15: Building Complex I, Room 70...... 88

Figure 3.16: Building Complex I, Room 70. Flatbed scan of micromorphology slide of a fired mud-brick...... 89

xiv

Figure 3.17: Building Complex I, Room 70. Flatbed scan of a wall sample with mud rendered surface between a) fill and b) mud-brick layers...... 90

Figure 3.18: Building Complex I, Silo 339. Northern wall of the Chalcolithic burial pit...... 91

Figure 3.19: Building Complex I, Silo 339. An example of unworked clay incorporated into the mud-brick used in construction of the silo...... 92

Figure 3.20: Building Complex II, Phase 3. The micromorphology samples marked are those that were processed for analysis and are discussed below...... 93

Figure 3.21: Room 230, Building Complex II. The baulk (highlighted in white) transects the southern portion of the interior of the building, bisecting a later pit...... 94

Figure 3.22: Building Complex II, Building 230. Idealized section drawing reconstructed from field notes...... 95

Figure 3.23: Building Complex II, Building 230. Laminate phytolith and fecal spherulite matrix (MF 3a)...... 96

Figure 3.24: Room 268, Building II. A small baulk was left along the southeast wall at the beginning of the 2007 season...... 97

Figure 3.25: Building Complex II, Room 268. The same field of view is presented in both images (MF 3b)...... 98

Figure 3.26: Building Complex II and IV, Silo 272. The silo was sectioned exposing alternating layers of hard mud-brick and softer fill material...... 99

Figure 3.27: Building Complex II and IV, Silo 272. A piece of hard-packed floor material from the silo has visible white phytoliths embedded in it...... 100

Figure 3.28: Building Complex II and IV, Silo 272. The location of the micromorphology samples are circled...... 101

Figure 3.29: Byzantine Surface, located north of Building Complex II (circled). The test pit is indicated by the square...... 102

xv

Figure 3.30: Byzantine Surface...... 102

Figure 3.31: Building IV, Phase 4. Circular room 662 and Rectangular room 612...... 103

Figure 3.32: Building Complex IV, Room 662. Example of a courtyard floor fragment with some characteristic features indicated (MF 1b)...... 105

Figure 3.33: Building Complex IV, Room 612. Wall area where sample TS07-39 was extracted: ...... 107

Figure 3.34: Building Complex IV, Room 612. Amorphous organic material with ash infill. 107

Figure 3.35: Building Complex IV, Room 612...... 108

Figure 3.36: Building Complex IV, Southern Courtyard. Possible mud-brick...... 110

Figure 3.37: Building Complex IV, Southern Courtyard. Some building fragments from this area had a clear, dark clay coating on the lower exterior surfaces...... 111

Figure 3.38: Building III, Courtyard. This burnt level was interpreted as an oven feature and is located in the south-east portion of the courtyard...... 112

Figure 3.39: Building Complex III, Oven Feature...... 113

Figure 3.40: Phytoliths from Tel Tsaf bulk sample ...... 115

Figure 3.41: Example of fecal spherulites preserved at Tel Tsaf...... 116

Figure 3.42: Location of bulk sample collection...... 119

Figure 3.43: Cumulative frequency graph illustrating fecal spherulite relative abundance by context...... 120

Figure 3.44: Cumulative frequency graph illustrating trends in phytolith relative abundance by context...... 121

Figure 3.45: Cumulative frequency graphs of fecal spherulites and phytoliths by context...... 122

xvi

Figure 3.46: The mean grain-size distribution from each context at Tel Tsaf and the overall mean distribution of all samples...... 124

Figure 3.47: Mean grain-size distribution from all contexts at Tel Tsaf with Standard Deviation (1 σ) for each context and grain-size...... 125

Figure 3.48: Grain-size distribution of bulk samples from room interiors at Tel Tsaf...... 127

Figure 3.49: Grain-size distribution of bulk samples from courtyards at Tel Tsaf...... 132

Figure 3.50: Grain-size distribution of bulk samples from silos at Tel Tsaf...... 134

Figure 3.51: Grain-size distribution of bulk samples from roasting pits at Tel Tsaf...... 136

Figure 3.52: Grain-size distribution of bulk samples from animal pens at Tel Tsaf...... 138

Figure 3.53: Building Complex I noting all bulk samples and relative abundance of fecal spherulites and phytoliths are plotted and microfacies present in each context...... 140

Figure 3.54: Building Complex II noting all bulk samples and relative abundance of fecal spherulites and phytoliths are plotted and microfacies present in each context...... 142

Figure 3.55: Building Complex IV. All bulk samples and relative abundance of fecal spherulites and phytoliths are plotted and microfacies present in each context are noted ...... 143

Figure 4.1: Formation processes evident in two micromorphology samples from Building Complex I...... 148

Figure 4.2: Formation processes evident in three micromorphology samples from Building Complex II...... 149

Figure 4.3: Formation processes evident in three micromorphology samples from Building Complex IV...... 150

Figure 4.4: Phases 4 and 3, Building Complex 1...... 162

Figure 4.5: Room 70, Building Complex I. The indicate the Phase 4 and Phase 3 walls...... 163

xvii

Figure 4.6: Phases 4 and 3, Building Complexes IV and II...... 164

Figure 5.1: Hillier and Hansen’s (1984) eight elementary syntaxes...... 171

Figure 5.2: Building Complex I. The Room 70 entrance is circled and potential north or west courtyard entrances are indicated by arrows...... 172

Figure 5.3: Building Complex II. The entrances to rooms 230 and 263 are circled and the potential east and south courtyard entrances are indicated with arrows...... 175

Figure 5.4: The relationship among storage, surplus and wealth...... 181

Figure 6.1: Diagram of Silo 339……………………………………………………...………..210

Figure 6.2: Regional cultural entities as suggested by Gilead (2011)...... 225

xviii

List of Appendices

Appendix I: Micromorphology Sample Analysis Forms ...... 271

Appendix II: Bulk Sample Grain Size Analysis Data………………………………………….326

xix 1

Chapter 1 Introduction

In the southern Levant, the Chalcolithic Period (c. 5200-3600 calBC) can be viewed as the transition between egalitarian and hierarchical, urban societies in the region. It has tantalizing hints of emerging social hierarchy and religion at both the site and regional level. However, the archaeological evidence required to truly understand how society and the economy were organized remains elusive. Archaeologists have found prestige items made of copper, but do not know who them made or why. Formal cemeteries and tombs have been discovered, but surprisingly little is known about the lives of the people who are buried within them. There are shrines, but the archaeology in the vicinity of these structures is limited and shed little light on their use or relationships to other sites. Large-scale excavation of settlements is limited to a few larger village sites (e.g., Teleilat Ghassul, Shiqmim, Gilat), with the rest of our data coming from small-scale excavation of special-purpose sites, such as shrines and tombs. Further complicating the situation, debate rages regarding which theoretical framework is appropriate when discussing Chalcolithic society. One site that is beginning to contribute to these debates is Tel Tsaf.

Excavations at the Chalcolithic site of Tel Tsaf in Israel began between 2004 and 2007, under the direction of Yosef Garfinkel (Hebrew University), with Danny Rosenberg (University of ) renewing work there in 2013. The original project research goals included providing data regarding site layout and structure in order to contribute to the limited settlement data mentioned above and to better elucidate phasing of the poorly understood Middle Chalcolithic period. This dissertation will focus on the results of micromorphological and bulk-sediment analysis that the author conducted while participating in the first phase of research at the site, with the aim of contributing to our understanding of domestic life at Tel Tsaf and evaluating the data in light of current debates within the archaeological community regarding social organization in the Chalcolithic period.

1.1 Research Objectives

The research outlined in this dissertation uses micro-scale geoarchaeological techniques to explore the use of space and its potential social and economic implications at the Chalcolithic site of Tel Tsaf, Israel. Small-scale analysis of occupational deposits at the site has provided rich

2 datasets regarding the elusive function of several structures that were not obtainable otherwise. The results help to address the primary research questions of this dissertation:

1. What are the implications of varied architectural forms during a single phase of occupation?

2. With evidence of silos and animal pens at Tel Tsaf, is it possible to identify meaningful surplus? If so, what are its social and economic implications?

3. How can the results of microscopic sediment analysis provide data that directly contributes to big picture debates regarding the possibility of chiefdoms during the Chalcolithic?

These research questions complement current discourses regarding to the Chalcolithic period in the southern Levant. Scholars have spent the past few decades discussing how to classify the Chalcolithic period, with the most prominent argument advocating a chiefdom-level society (Alon and Levy 1980; Burton and Levy 2011a; Dawson, et al. 2003; Levy 1986a, 1995, 2006; Levy and Shalev 1989). This position, however, has not been unanimously accepted and has often met opposition (Bourke 2002; Gilead 1988, 2011; Joffe and Dessel 1995a). Proponents of chiefdom-level organization in the southern Levantine Chalcolithic contend that a number of features, including an increase in population and population density, regional diversity in material culture, competition and warfare, agricultural intensification, craft specialization, deployment of public labour and the maintenance of communal ritual or ceremonial centres, provide key evidence (Alon and Levy 1980; Burton and Levy 2011a; Levy 1986a, 2006; Levy and Shalev 1989). Of these features, agricultural intensification, craft specialization and communal ceremonial centres are most relevant to the research questions above, as they are often framed within the context of surplus food production evidence at domestic sites.

Childe (1950), following a Malthusian perspective, suggested that population density can be directly related to a population’s access to resources, technique of exploitation and ability to transport and store food. He posits that a population can only increase if it has the means and ability to feed itself first. Alternatively, it has been suggested that other factors, perhaps medical, biological or political, can lead to population increase, which in turn forces populations to intensify agricultural activities (Boserup 1965, 1981). Regardless of which approach one

3 chooses to follow, if a group of any size is not able to meet its basic subsistence needs, it is not sustainable. When maintenance of a population is a priority, every member of a group will focus on food-procuring activities such as hunting or farming. Only once yields increase and storage is improved, is it possible to create a surplus of food. Once a group is able to sustain itself, regardless of the reason for population increase, members can be freed to explore other endeavors, creating a ‘social surplus’ (Childe 1950). This means that there is enough food to support not only those who procured it, but those who are engaged in non-subsistence tasks such as craft production or communal building projects. A surplus of agricultural products influences a change in organization because it can provide one individual or group with more than another, potentially creating prestige or power imbalances. It has been suggested that the degree of stratification in a society is directly tied to the amount of surplus being produced (Sahlins 1958). Therefore it is imperative to identify accurately what was being stored and on what scale, as well as who benefited from the stored goods, in order to engage meaningfully in discussions of the social and economic implications on a larger scale. This dissertation attempts to do so by providing a detailed analysis and discussion of the micro-scale domestic evidence from Tel Tsaf.

Excavations at Tel Tsaf have been limited to a single period, with good architectural preservation from at least two Middle Chalcolithic phases. The excavations there uncovered parts of five incomplete courtyard complexes. All the complexes consist of courtyard walls that enclose buildings and silos. The shapes and sizes of the internal buildings vary among complexes and phases, raising interesting questions regarding how each complex was used. The varied architecture, combined with the site’s chronological placement, makes Tel Tsaf especially important in discussions regarding social and economic organization during the Chalcolithic period.

In order to obtain the data required to interrogate the use of space, micromorphology and micro- bulk sample analysis were chosen methods of analysis. Micromorphology is a powerful tool for assessing how space was constructed and used. It can provide details regarding floor and brick construction, and it can highlight maintenance or repair of these features. Microscopic analysis provides evidence of the presence and distribution of plants and animals at a site by identifying and mapping indicators such as phytoliths and fecal spherulites. These types of data enrich our understanding of how people choose to structure and use the build environment in which they lived. Building layouts can also provide social and economic insights if one looks at, for

4 example, whether access to rooms and storage was open or restricted (Banning 2010; Hillier and Hanson 1984). In order to understand why access may be restricted, it helps to understand how each space within a building was used by its occupants. Exploring these questions allows greater insight into how households, neighbourhoods and communities functioned and interacted. Combing both micro and macro-scale data allows researchers to present a larger data set to such discussions.

This dissertation explores issues stemming from the data obtained through the geoarchaeological analysis of occupational deposits at Tel Tsaf. The data challenges existing interpretations that all small buildings at pre-urban sites are homes, contributing to discussions about how prehistoric societies organized themselves and their lives. Providing evidence that allows such conversations is essential, especially in relation to reconstructing social and economic organization in pre-urban society.

The use of secondary products, as animal products besides meat, and their contribution to changing economic structures at Tel Tsaf also receives attention here. Faunal evidence from the site indicates that cattle were likely used for plowing fields, which may have increased crop yields and contributed to the need for large silos. Pigs represent a significant portion, by Number of Identified Specimen (NISP), of the animal population at Tel Tsaf and their potential contributions to the economic system will also be explored, especially in relation to the presence of animal pens on site.

The question of when storage becomes surplus is discussed in detail. This is a complex issue and one that deserves more attention in the archaeological literature. It is unlikely that surplus can exist without some kind of storage facilities, but having silos and bins does not necessarily mean its contents are surplus. The presence of large storage silos and animal pens makes Tel Tsaf an ideal site for addressing this complex issue.

All of these issues then play into the larger question of how society was organized during the Chalcolithic period. Tel Tsaf is slightly earlier than the Late Chalcolithic sites that are central to the debate about chiefdom-level organization in the region (Blackham 2002; Bourke 2001; Gilead 1988, 2011; Levy 1986a, 1998, 2006; Philip 2011; Rowan and Golden 2009). The data

5 obtained through this research contributes significantly to the debate concerning how and when social and economic structures began to change in the southern Levant.

1.2 Summary of Chapters

A detailed overview of the Chalcolithic period in the southern Levant appears in Chapter 2. The chapter begins by providing chronological context, including a discussion of debates regarding classification of late prehistoric periods in the region. This is followed by an overview of the history of excavation in the southern Levant, placing the current discussion in its regional and historical context. An archaeological overview of all the relevant Chalcolithic phases (Early, Middle and Late) is provided. This is followed by a discussion of the current theoretical discourse on Chalcolithic society, with a focus on chiefdoms. The origin of the term and its specific application to Late Chalcolithic societies is outlined, providing a basis for further discussion in later chapters. The chapter concludes with a detailed overview of the architecture and material culture of the study site, Tel Tsaf.

Methods and results of this research are outlined and presented in Chapter 3. Chapter 3 begins with an overview of site-formation processes at Tel Tsaf with specific reference to how anthropogenic features relate to each other. This is followed by an overview of the types of features most common in the Tel Tsaf samples and their distribution across the site. A more detailed discussion of the results from each complex, including types of features and their distribution, follows. The results from the bulk-sample analysis focus on the relative abundance of phytoliths and fecal spherulites across complexes and phases.

Having provided a summary of the geoarchaeological findings in Chapter 3, Chapter 4 discusses potential implications for the use of specific areas at Tel Tsaf. This section incorporates other lines of evidence, such as faunal and ethnographic data, to present a more detailed discussion of what life may have been like at Tel Tsaf during the Chalcolithic. It addresses questions such as which animals were kept on site, how animals and their products were being used by the inhabitants and why some building complexes changed in form or function between occupation phases while others remained consistent. The chapter concludes with a summary of how space may have been used by the inhabitants of Tel Tsaf.

6

Chapter 5 builds upon the practical assertions of the previous chapter and begins to address some of the larger social and theoretical issues that arise, most focused on the evidence from Phase 3. It begins with a more in-depth discussion of architectural variation between building complexes, drawing briefly on space syntax and access analysis to highlight social and functional differences. There is further discussion of the most likely factors influencing these differences. Next, the idea of surplus is introduced, related to both grain storage in the prominent silos and animal storage as highlighted by the presence of pens. Potential explanations for the extensive storage capacities on site are discussed and evaluated. Having established the likely presence of food surplus at Tel Tsaf, the idea of control of resources and the concept of public vs. private ownership is explored. Furthering this discussion, the concepts of power, conspicuous consumption and feasting are addressed within the context of the evidence from Tel Tsaf.

Having presented and evaluated potential social and economic structures at Tel Tsaf, Chapter 6 aims to place the site within the wider Chalcolithic context. Revisiting the discussion of chiefdoms in Chapter 2, this chapter presents and evaluates the current debates regarding social organisation of Chalcolithic societies. The evidence from Tel Tsaf differs from what has been seen at other sites so it is important to place the findings within these wider theoretical discussions. This chapter provides an overview of how the evidence fits into the larger regional and chronological discussion.

Chapter 7 provides a summary of the research and outlines the conclusions and contributions of this study. It also outlines questions that have arisen from this work and avenues of further research that can contribute to the scholarly debate surrounding the Chalcolithic.

7

Chapter 2 Background Chapter

The Chalcolithic was an important period of social and political change, yet it is often difficult to know how widespread they were in the Near East. There have been excavations at a few special- purpose sites in the region but limited comprehensive settlement excavations. Of the few sites that have been excavated on a large-scale, most provide few secure radiocarbon dates (with the exception of Teleilat Ghassul and Shiqmim), and often represent atypical sites such as burials, specialized craft-production or religious sites (Levy 1995; Rowan and Golden 2009). Attempts to summarize the data are often constrained by modern geo-political borders (Bourke 2001; Levy 1995), making it difficult to gain a view of the Levantine Chalcolithic as a whole. Furthermore, both excavation data and relevant literature have traditionally focused on the Late Chalcolithic (), so publications purporting to discuss the Chalcolithic as a whole often omit data from earlier phases (Levy 1995). Excavation projects have provided excellent, high-resolution data that helped shape our understanding of specific sites, but our understanding of Chalcolithic society across the southern Levant is still somewhat limited. Despite these hurdles, the Chalcolithic of the southern Levant has much to offer regarding our understanding of Near Eastern prehistory.

This chapter provides an overview of the period and current debates that involve it. It aims to highlight these issues while providing the reader with the necessary information to place discussion of Tel Tsaf into its broader archaeological context. Summary of the issues related to chronology and a brief excavation history are followed by an overview of material culture from the period. Current debates regarding social and economic organization of the period are outlined. Finally, an overview of the architecture and material culture of Tel Tsaf is provided.

2.1 Chronology

The chronological classification of the southern Levant is strongly rooted in the Three Age System, introduced and discussed by Christian Jürgensen Thomsen (1848) and Jens Jacob Asmussen Worsaae (1849). Adopting a culture-history approach, archaeologists in the southern Levant applied this framework to delineate changes in technology and settlement type within the region by inserting a Chalcolithic period between the and Bronze Ages. Classification

8 of sites within this system has traditionally relied upon material-culture analysis, with archaeologists creating culture histories of artifact classes based on changes in style and technology. The Chalcolithic was first defined not only on the perceived introduction of copper production but also the change from plain to painted pottery (Lloyd and Safar 1945). J. Perrot (1955, 1963, 1968) on the basis of his excavations in the Beersheva Valley, presented the Ghassulian as a new population in the area that brought copper smelting with them. Noting a limited distribution, de Vaux (1970) downplayed the role of copper in defining the Chalcolithic and instead suggested defining the period by villages of farmers, potters and metalworkers who were just beginning to use copper.

Within this ‘stage’ system, the Chalcolithic period has always been the least defined in terms of dating and identifying features. De Vaux’s observation that “great confusion in fact prevails in the application of the terms ‘Neolithic’ and ‘Chalcolithic’ by Palestinian archaeologists” (1970: 520) still holds true today. Agricultural villages in the Neolithic and urbanization in the Early drew most attention from early excavators, with the period between the two classified as Chalcolithic or Proto-Urban (Mellaart 1966). Within the archaeological literature for the southern Levant, the Chalcolithic was almost exclusively defined by the excavations at Teleilat Ghassul (Hennessy 1969; Koeppel, Mallon, et al. 1940; R. Mallon, et al. 1934; North 1961) and the Beersheva Valley (Perrot 1955) until the 1980’s, when our understanding of the and Chalcolithic improved through increased excavation of 5th- and 4th- millennium sites (Gilead 1988; Levy 1986b, 1987). A wider variety of sites and improved radiocarbon dating helped to define the Chalcolithic period more effectively. Levy moved beyond dates and pottery typology to suggest that formal temples and burial grounds combined with settlement hierarchies within and among sites defined the period and marked very distinct social and economic changes (Levy 1986a).

As excavations continued and more material culture was available for study, it became apparent that treating the entire Chalcolithic as a single period was inappropriate. Excavations of Late Neolithic and Early Chalcolithic sites highlighted greater variability in settlement and house size, as well as other aspects of material culture across the southern Levant. The ceramic finds at sites such as Wadi Rabah prompted Kaplan to propose more nuanced divisions within the Neolithic- Chalcolithic transition, highlighting pre-Ghassulian Chalcolithic material (Kaplan 1958b, 1959) Even the site of Teleilat Ghassul was not consistently “Ghassulian” throughout its Chalcolithic

9 occupation sequence. In an attempt to date and define the period better, Joffe and Dessel (1995b) compiled and analysed known radiocarbon dates from the Chalcolithic period. They were unable to define the beginning of the period clearly, but suggested that Wadi Rabah and Qatifian complexes should be considered Early Chalcolithic (table 2.1). They suggested the term Developed Chalcolithic to describe the Beersheva, Ghassulian and Golan sites, dating the phase to 4500-3700 cal BC. They further suggested that the cache dated to the Terminal Chalcolithic, dating 3700-3500 cal BC. While their terminologies were not adopted, the study did help to constrain the time period, and better separate the Late Chalcolithic from the Early Bronze age.

Based on excavations at Munhata (1992a) and the results of his PhD thesis, Y. Garfinkel published a comprehensive synthesis of prehistoric ceramics from the Levant (1999). This volume summarized historical research, and used typology and vessel morphology to create a framework for ceramic analysis and chronology for the region. This analysis formally subdivided the Chalcolithic into three phases: Early, Middle and Late. Garfinkel’s attempt to create a universal chronology met with criticism of both his method and interpretation (Banning 2001a; Blackham 2002; Gilead 2009; Lovell 2001b), but it became a starting point for discussions to refine chronological and material culture data for the Late Neolithic through the Early Bronze Age. An abundance of new data and dates appeared in the early 2000’s (Blackham 2002; Bourke, et al. 2001; Burton and Levy 2001; Garfinkel and Miller 2002; Kafafi 2001; Kerner 2001; Lovell 2001a). These studies still relied almost exclusively on ceramic data, and were inconsistent in how they made use of radiocarbon data (Banning 2002). These studies and their critiques highlight the need for more data and a greater consensus in how chronology and culture of the Levant should be discussed.

Many recent discussions have acknowledged the lack of a consistent terminology to describe the cultural and chronological phasing of the Late Neolithic and Chalcolithic (6th and 5th millennia calBC) in the southern Levant (Anfinset , et al. 2011; Banning 2007a, 2007b; Banning, et al. 2011; Bourke 2007; Burton and Levy 2011a; Garfinkel, et al. 2007; Gilead 2011; Kafafi 2011; Kerner 2010; Lovell, et al. 2007; Philip 2011; Roux, et al. 2011; Rowan and Lovell 2011; Shugar and Gohm 2011). A high degree of regional variation in material culture and settlement structure has become increasingly apparent and has led many researchers to favour the use of cultural phases, such as the Wadi Rabah and Besorian, over chronological terminology during the

10

Neolithic/Chalcolithic transition. Within a cultural framework, some have even argued that the Chalcolithic should only be used to described the Ghassulian culture dating to 4500–3900 BC with all earlier sites, Tel Tsaf included, being classified as either Intermediate Neolithic/Chalcolithic or Late Neolithic (Anfinset , et al. 2011; Gilead 2011). The extent of disagreement over chronological and cultural classification of the Late Neolithic and Chalcolithic in the southern Levant is summarized in the table below (table 2.1).

Bourke Bourke Gilead Garfinkel Banning (South) (North) Lovell Levy

3500 Late Late Late Chalcolithic Chalcolithic Chalcolithic Ghassulian Late Ghassulian Ghassulian Chalcolithic 4000 Late Late Ghassulian Pella Chalcolithic Chalcolithic Middle Ghassulian Ghassulian Chalcolithic

Early Chalcolithic Early 4500 Neolithic- Early Chalcolithic Chalcolithic Middle Chalcolithic Pottery Transition calBC Chalcolithic Middle Late Neolithic Wadi Rabah Qatifian Chalcolithic Neolithic "Ziqlab Neo" variants, 5000 Late Qatifian WadiNeolithic Rabah, Yarkmoukian,Lodian, Late Timnian, Early Neolithic Late Lodian Chalcolithic Wadi Rabah Neolithic 5500 Wadi Rabah

6000

Table 2.1: Chronological classifications of the Chalcolithic. A general overview of current schemes regarding the classification of periods and cultures in and around the Chalcolithic in the Near East. (Gilead 2011; Garfinkel 1999; Garfinkel et al. 2007; Banning 2007a: 141; Bourke 2007; Lovell 2001, 2007; Levy 2007)

The disparity between cultural phases and chronological periods is further exacerbated by inconsistency in the interpretation of radiocarbon data. The radiocarbon calibration curve for this portion of prehistory is in itself problematic, as demonstrated below (Banning 2007a; Campbell 2007; Manning 2007a). Refinement of the calibration curve is continuous as new environmental

11 data is integrated. Many dates from Chalcolithic sites were calibrated based on the IntCal98 curve, which was estimated at a 10-year resolution, resulting in a more ‘ragged’ curve (Manning 2007a, 2007b). As more detailed dendrochronological data becomes available, the calibration curves are updated, creating smoother curves. To demonstrate the effect of new data on archaeological estimates of ages, we can look at the IntCal04 calibration curve, as it has been clearly laid out and discussed (Manning 2007a, 2007b). The model is estimated at five-year intervals creating a smoothing effect on the curve. This new, smoother curve results in the ability to distinguish between possible dates more discretely due to the higher resolution, but in some cases the smoother plateaux actually present a larger possible calendar BP date range for BC dates during the 4th millennium (figure 2.1). This has further been refined with the release of the IntCal09 curve and IntCal2013. Combining these issues with few secure dates and varying approaches to the data as whole, it is difficult for the academic community to come to consensus on the dating of periods (Banning 2007a; Gilead 2007; Lovell, et al. 2007).

While this can make meaningful discussion of chronology difficult, the academic community has begun to address these problems and continues to move towards a more universal understanding of the Chalcolithic (Anfinset , et al. 2011; Banning 2007a, 2007b; Banning, et al. 2011; Bourke 2007; Burton and Levy 2011b; Gilead 2011; Kafafi 2011; Lovell, et al. 2007; Philip 2011; Roux, et al. 2011; Rowan and Lovell 2011; Shugar and Gohm 2011). Most recently, Rowan and Lovell (2011) have suggested moving beyond dates and ceramics to rethink theoretical approaches to the data. By approaching the period with clearer research objectives and hypotheses, they hope that the choices made during classification may be better understood and contribute to a chronological definition of the Chalcolithic.

12

Figure 2.1: The Radiocarbon calibration curve shown as a 1σ band for the Chalcolithic period in the Southern Levant (Manning 2007a). Insets A and B demonstrate areas of significant variation between the IntCal98 (Stuiver, et al. 1998) and IntCal04 (Reimer, et al. 2004) curves. Inset A demonstrates that IntCal04 allows researchers to distinguish dates between 5300-5200 BC from those between 5200-5060 BC, something that was not possible with the IntCal98 curve. It also highlights the longer calendar date range possible for some BP years as the result of the smoother IntCal04 curve. For example when a date of 6200 BP is plotted, possible calendar dates range between 5160 and 5060 BP (after Manning 2007a).

It is beyond the scope of this dissertation to attempt to resolve the dating and terminology issues addressed above. However, it is necessary to provide a clear chronological framework for the discussion contained within. It could be argued that only two Chalcolithic sub-phases, Early and Late, are appropriate given the current literature and data. If we were to follow this model, Tel Tsaf would be Early Chalcolithic and the Wadi Rabah would fall within the Late Neolithic. However, if one chooses to place the Wadi Rabah chronologically as Early Chalcolithic, there is justification for the use of the three sub-phases. For the purpose of this dissertation, I will use this approach and continue to refer to Tel Tsaf as a Middle Chalcolithic site. This remains consistent with much of the current published data about the site (Ben-Shlomo, et al. 2009; Garfinkel and Ben Shlomo 2007; Garfinkel, et al. 2007) and works towards creating consistency in the literature.

13

In order to assess the probable start and end dates, as well as probable duration for the Middle Chalcolithic period, I used the Bayesian probability analysis software BCal (http://bcal.sheffield.ac.uk; (Buck, et al. 1999), following Banning (2007b). I entered accepted Wadi Rabah dates from Banning (2007b), Middle Chalcolithic dates from Tel Tsaf and Abu Hamid II (which Banning suggests is probably later than the Wadi Rabah phase Abu Hamid I)(Banning 2007b; Garfinkel, et al. 2009) and Late Chalcolithic dates from modern excavations at Shiqmim (Burton and Levy 2011b) and Teleilat Ghassul (Bourke, et al. 2004)(table 2.2). The model was run four times using the IntCal2009 calibration curve (Reimer, et al. 2009), with different random-number seeds each run, in order to evaluate the reliability of the results (table 2.3)

EARLY CHALCOLITHIC (WADI RABAH) MIDDLE CHALCOLITHIC LATE CHALCOLITHIC Site (and level) Lab No. Mean Error Source Site (and level) Lab No. Mean Error Source Site (and level) Lab Number Mean Error Source Haber and Lovell et Burton and Hagoshrim 4 6505 120 Abu Hamid II Ly-6248 5685 65 Shiqmim Strat 1 Beta-161863 5780 100 Dayan 2004 al. 2007 Levy 2011 Gopher and Megadim Pts-43648 6310 70 Abu Hamid II Ly-6249 5580 120 Ibid. Shiqmim Strat 1 RT-859E 5390 180 Ibid. Gophna 1993 Megadim Pta-4339 6270 50 Ibid. Abu Hamid II Ly-6251 5545 75 Ibid. Shiqmim Strat 1 RT-859D 5370 180 Ibid. Tel Dover TO-8310 6840 210 Banning 2007 Abu Hamid II Ly-6253 5860 70 Ibid. Shiqmim Strat 1 Beta-161868 5460 90 Ibid. Garfinkel et Tel Hanan TO-11714 6750 130 Ibid. Tel Tsaf RTT 5479 6150 55 Shiqmim Strat 1 Beta-161865 5440 80 Ibid. al. 2009 Ard Tlaili basal K-1432 6870 130 Kirkbride 1969 Tel Tsaf RTT 5478 6110 75 Ibid. Shiqmim Strat 1 RT-554A 5250 140 Ibid. Ard Tlaili basal K-1433 6850 130 Ibid. Tel Tsaf RTT 5477 6085 50 Ibid. Shiqmim Strat 1 RT-1341 5370 40 Ibid. Ard Tlaili basal K-1434 6790 130 Ibid. Tel Tsaf RT 6040 6040 65 Ibid. Shiqmim Strat 1 Beta-161871 5300 80 Ibid. Ard Tlaili middle K-1431 6660 140 Ibid. Tel Tsaf RTT 5476 5955 50 Ibid. Shiqmim Strat 1 RT-859C 5080 180 Ibid. Lovell et al. Abu Hamid I Ly-6174 6170 50 Tel Tsaf RTT 5475 5850 45 Ibid. Shiqmim Strat 1 Beta-161875 5280 70 Ibid. 2007 Abu Hamid I Ly-6254 6230 55 Ibid. Tel Tsaf RT 5044 5845 35 Ibid. Shiqmim Strat 1 Beta-161866 5270 70 Ibid. Abu Hamid I Ly-6255 6215 70 Ibid. Tel Tsaf RTT 5040 5840 55 Ibid. Shiqmim Strat 1 Beta-161874 5270 70 Ibid. Abu Hamid I Ly-6259 6170 60 Ibid. Shiqmim Strat 1 Beta-161870 5270 70 Ibid. Basatin TO-12151 6710 70 Banning 2007 Shiqmim Strat 1 Beta-161876 5300 50 Ibid. Basatin TO-12738 6650 140 Ibid. Shiqmim Strat 1 Beta-161867 5130 70 Ibid. Basatin TO-13091 6550 60 Ibid. Shiqmim Strat 1 Beta-161869 5250 50 Ibid. Basatin TO-13092 6680 60 Ibid. Shiqmim Strat 1 Beta-161864 5220 70 Ibid. Basatin TO-13094 6400 80 Ibid. Shiqmim Strat 1 Beta-161872 5220 70 Ibid. Banning et al. Bourke et Tabaqat al-Buma TO-2114 6590 70 Teleilat Ghassul OZD 030 5552 163 1994 al. 2001 Tabaqat al-Buma TO-2115 6630 80 Ibid. Teleilat Ghassul OZD 029 5524 88 Ibid. Tabaqat al-Buma TO-3408 6190 70 Ibid. Teleilat Ghassul OZD 033 5454 58 Ibid. Tabaqat al-Buma TO-3409 6900 70 Ibid. Teleilat Ghassul OZD 034 5342 71 Ibid. Tabaqat al-Buma TO-3410 6350 70 Ibid. Teleilat Ghassul OZD 031 5605 80 Ibid. Tabaqat al-Buma TO-3411 6670 60 Ibid. Teleilat Ghassul OZD 028 5581 67 Ibid. Tabaqat al-Buma TO-3412 6380 70 Ibid. Teleilat Ghassul OZD 032 5577 71 Ibid. Tabaqat al-Buma TO-4277 6490 70 Ibid.

Table 2.2: List of radiocarbon determinations used for the analysis. Further radiocarbon dates have been released for some sites, but were excluded from this analysis if the error range was too large.

14

First Run Second Run Third Run Fourth Run Early/Middle Chalcolithic 5101 - 5028 5098 - 5024 5095 - 5022 5102 - 5025 Boundary Middle/Late Chalcolithic 4469 - 4399 4471 - 4399 4468 - 4400 4466 - 4399 Boundary Duration of Middle 571-679 years 576 - 679 years 573 - 681 years 568 - 678 years Chalcolithic

Table 2.3: Comparison of results for the calibrated boundaries between Early Chalcolithic (Wadi Rabah), Middle and Late Chalcolithic periods and probable duration of the Middle Chalcolithic period. The data presented are the 68% confidence intervals. The dates were calibrated using the IntCal09 calibration curve.

The data above does not conclusively prove the start or end dates for the Middle Chalcolithic, but it does correlate well with Garfinkel’s proposed date for the period. The results also provide a good indication of the likely occupation dates of the site of Tel Tsaf itself. Based on the above data, I propose the Tel Tsaf occupation should be dated between approximately 5040 and 4440 cal BC, with a duration of 600 years. Until more Middle Chalcolithic sites are excavated and dated, the Tel Tsaf occupation dates also represent the known duration of the period itself.

Having discussed the chronological context for the southern Levantine Chalcolithic, established a framework for this dissertation and presented probable start and end dates and duration for the Middle Chalcolithic period, we will now move to an overview of the Chalcolithic period.

2.2 Overview of Excavation History

A first glance at figure 2.2, showing known Chalcolithic sites within the Southern Levant, suggests that there is a wealth of sites to inform archaeologists about the lives of Chalcolithic people. Closer inspection of the map highlights the number of survey, test pit and salvage excavations, as well as many special-purpose sites, demonstrating that a wealth of sites does not actually directly translate to abundance of information. Special-purpose sites, such as burials, copper hoards and shrines provide information about highly specialized aspects of society, but do not often shed light on everyday activities. Survey, test pit and salvage excavations are excellent for identifying the location of Chalcolithic sites, but they often only yield a limited number of artifacts, sometimes without context. They may identify limited evidence of architecture, but contribute little to the larger picture of site structure and organization. This section is intended to establish an understanding of where our information for the Chalcolithic comes from, and to highlight why there are still so many questions surrounding the nature of Chalcolithic society in the southern Levant.

15

Figure 2.2: Map of Chalcolithic sites in the southern Levant. Sites are classified by period and type. In cases where evidence of multiple periods or uses are present, the best defined phase or use was mapped. Well-defined Late Chalcolithic sites are characterised by good horizontal exposures of architectural features combined with well published reports. Minimally defined sites are those which have been identified, but have limited horizontal exposure or are poorly published. The Early Chalcolithic site of Hagoshrim is not on the map, but is located north of Tel Turmus, east of the Jordan River. (after Rowan and Golden 2009)

16

The first Chalcolithic-focused excavations were conducted by the Pontifical Biblical Institute at the site of Teleilat Ghassul in the Jordan Valley (Blackham 1999; Koeppel, Mallon, et al. 1940; R. Mallon, et al. 1934; North 1961). The extensive excavations revealed architecture and material culture that has come to define the Late Chalcolithic. To this day, it remains the most complete settlement excavation from the period and was the focus of renewed excavations by Hennessy in the 1960’s (Hennessy 1969; 1982; 1989) and again by Bourke in the late 1990’s (Bourke 1997a, 2002; Bourke, et al. 1994, 2000). Pre-Ghassulian levels at the site have also contributed to the debate regarding classification of the Late Neolithic, Early and Middle Chalcolithic, providing good ceramic sequences and some architecture (Bourke 1997b; Gilead 2007; 2001a, 2001c).

During the 1950’s and 60’s, excavations were undertaken in the at Bir es-Safadi and Tell Abu Matar (Perrot 1955, 1959, 1968). These assemblages presented similar features to those at Teleilat Ghassul and highlighted the potential wealth of data to be found in the Beersheva Valley. Renewed survey and excavation were undertaken beginning in the late 1970’s, running for decades and enriching the understanding of the Chalcolithic in the region. The bulk of the evidence from this region comes from the two sites mentioned above, two other large sites of Shiqmim and Horvat Beter, as well as many smaller sites discovered during survey and explored with sondages (Alon and Levy 1980; Levy 1987; Levy and Alon 1979, 1982, 1985a, 1985b, 1985b, 1987b; Levy, et al. 1991, 1994, 2006a). In addition, the Wadi Gaza/Nahal Besor area revealed two further sites, Grar and Gilat that have become key to our understanding of the Beersheva Valley (Gilead and Alon 1988; Gilead 1989; Gilead, et al. 1995; Levy 2006). The first evidence of the Wadi Rabah culture was also uncovered during excavations by Kaplan (1958, 1958b, 1969b) at the sites of Wadi Rabah, Tel Batashi, HaBashan St and ‘En Jarba. These revealed new pottery styles and decoration that Kaplan classified as belonging to a new cultural phase between the Yarmoukian and Ghassulian ceramic sequences.

Beginning in the 1970’s, Claire Epstein conducted extensive survey and excavation of at least 25 small sites in the Golan region (Epstein 1977, 1978b, 1985, 1988, 1998). Her excavations uncovered multiple broadroom houses as well as many figurines, ceramics and lithics. The material showed many similarities to the Ghassulian material culture, but there were marked differences in settlement size, layout and ceramic forms. There is also evidence that the region may have been a centre for olive oil production (Epstein 1993). The sites discovered were very

17 limited in size and are often referred to in the literature collectively as the Golan sites, rather than by individual names.

The Jordan Valley and highlands north of the have also received a fair amount of attention from scholars. The region includes multi-phase sites with significant Chalcolithic components, such as Pella (Bourke, et al. 2003; McNicoll, et al. 1992; McNicoll and Smith 1982), Tell ash-Shuna North (Baird and Philip 1992, 1994; Gustavson-Gaube 1985, 1986; Mellaart 1962; Philip and Baird 1993), Munhata (Garfinkel 1992a; Gopher 1989; Perrot 1966, 1967), Beth Shan (Fitzgerald 1934, 1935) and Abu Hamid (Dollfus and Kafafi 1993; Lovell, et al. 2004; Lovell, et al. 1997). Overall, within the region excavations revealed some architecture, but the majority of the data comes from artifactual remains. Long-term programs, such as the Wadi Ziqlab Project, have also identified a number of Late Neolithic and Chalcolithic sites (Banning 1983, 1992, 1995, 1996, 1997; Banning and Najjar 1999; Banning, et al. 1987, 1989, 1992, 1998, 2002, 2005; Blackham, et al. 1998; Gibbs, et al. 2006, 2010; Kadowaki, et al. 2008). Such projects have highlighted the diversity in site types during these periods, with extensive survey locating numerous small sites as well as larger settlements.

In the Chalcolithic literature, much focus has been placed on special-purpose sites. In general, burial sites, cave caches and ritual buildings have been better preserved, excavated and published. Salvage excavations in the Galilee region of Israel uncovered prolific burial finds at the site of Peqi’in (Gal, et al. 1996, 1997). The discovery of two well-preserved burial chambers at Peqi’in containing ossuaries, chalices and figurines gained much public attention in the late 1990’s. The ossuaries demonstrated many similarities to those uncovered from numerous cave sites along the Mediterranean Coast. The cave sites near Azor contained numerous anthropomorphic ossuaries (Ben-Tor 1975). In both the Galilee and the coastal caves, nearby substantial settlements have yet to be identified, leaving many questions regarding who actually used these burial sites. The Galilee Prehistory Project has been designed in part to address this question and current excavations at the site of Marj Rabbah should contribute greatly to our knowledge of the Chalcolithic inhabitants of the region. In addition to burial caves, there have been a few cemeteries associated with habitation sites such as Shiqmim and Gilat (Alon and Levy 1989; Levy and Alon 1982, 1985a, 1985b).

18

Two Chalcolithic ritual sites, ‘En Gedi and Gilat, are located in Israel. ‘En Gedi is a large compound located high on a mountaintop with no close proximity to any known Chalcolithic village sites. Its narrow range of decorated pottery forms, the arrangement of features within the compound and the presence of possible feasting pits suggest that the site was used for some sort of public or ritual purpose rather than domestic use (Gilead 1990; Goren and Gopher 1995; Levy 1998; D. Ussishkin 1980). Gilat, located in the northern Negev, is a settlement site with a large multi-room building. There is greater evidence of domestic goods at Gilat than at ‘En Gedi, but the size and layout of the building combined with a significant number of special-use pottery forms and figurines at Gilat suggest it also served a ritual purpose. (Alon and Levy 1989; Levy 2006b; Commenge 2005; Commenge et al. 2006a, 2006b).

Copper objects were present but rare during the Chalcolithic period and the majority of known copper has been found in caches rather than in villages or hamlets. Shiqmim and Abu Matar present evidence of local smelting of copper and some finished products have been uncovered from these sites (Golden 2009; Golden et al. 2001; Shugar 2001; Shugar and Gohm 2010). The copper used in this production has been sourced to Wadi Faynan (Hauptman 2007; Perrot 1955; Shalev and Northover 1987; Shugar 2001)(figure 2.2). There was also copper available near Timna in the Sinai, but there is no compelling evidence this was exploited during the Chalcolithic period. Objects produced include axes, adzes, chisels and awls that were most frequently produced in open molds and pounded into their final form. The lost-wax technique was also employed to produce objects such as standards, mace heads, vessels and more elaborate objects (Levy and Shalev 1989; Shalev 1999). The largest collection of copper items was uncovered at Nahal Mishmar, also known as the Cave of the Treasure, in the early 1960’s (Ussishkin 1971). Another significant cache of copper items was found Nahal Qanah, which also contained gold ingots (Gopher and Tsuk 1991).

This overview highlights the skew of data towards special-purpose or ritual sites with more limited understanding of habitation sites. Regional variation is common and to be expected, but with a concentration of data in a few regions, especially Teleilat Ghassul and the northern Negev, and limited evidence tying ritual activities to daily routine, it can be difficult to assess a Chalcolithic culture as a whole. That being stated, I will now attempt an overview based on the information that we do have.

19

2.3 Early Chalcolithic (Wadi Rabah and variants)

The Wadi Rabah cultural phase is usually considered to bridge the chronological gap between the Pottery (or Late) Neolithic and the Chalcolithic. Its placement within this sequence is contentious and its classification as a culture, rather than a period, further complicates this. Regardless of how one chooses to label it, it is important to understand this cultural phase in order to fit the findings from Tel Tsaf into their regional and cultural context.

Kaplan (1958) first identified the Wadi Rabah complex during his excavations at the site of the same name. Located on the left bank of Wadi Rabah, a tributary of the Yarkon River, the site lies just east of Tel Aviv. Kaplan’s discovery of rectangular buildings there fit well with classic Ghassulian architecture, but the pottery was distinct from both Yarmoukian and Ghassulian examples (Kaplan 1958a). Kaplan based his classification of the site on the basis of its differences from previously known Yarmoukian and Ghassulian pottery forms. Diagnostic features included bow-rim jars and red and black burnished wares. Decoration is often incised or impressed (Garfinkel 1999; Rowan and Golden 2009).

Gopher and Gophna (1993: 326-327) suggest that at least 35 known sites exhibit Wadi Rabah (or related) features. They highlight the difficulty of defining the Wadi Rabah culture across the wider region, with suggested variants, including the Qatifian in the Negev. They do not provide any firm dates for the culture, but suggest that it does not overlap with the Yarmoukian or the Ghassulian (Gopher and Gophna 1993: 342). Architecturally, it is defined by rectangular buildings with stone foundations and a mudbrick superstructure (Gopher and Gophna 1993; Gopher 1998). The buildings are either single- or multi-celled. Gopher (1998) mentions a number of rounded silos associated with rectangular buildings, but does not specify at which sites these are found. Domesticated livestock dominate faunal assemblages with sheep and goat most common (Gopher 1998).

Excavations at Hagoshrim have revealed Yarmoukian, IX and Wadi Rabah levels, allowing researchers to document changes in subsistence patterns at the site and to contribute to the chronology of these cultures (Haber and Dayan 2004). The most significant change is a notable increase in domestic pig consumption during the Wadi Rabah phase.

20

Other sites with significant Early Chalcolithic occupation include Munhata (Garfinkel 1992a, 1995; Gopher 1989), ‘En Jarba (Kaplan 1969a), Neve Yam (Galili, Rosen, et al. 2009; Galili, et al. 1998; Prausnitz 1977; Wreschner 1977), and Nahal Zehora I and II (Gopher 1994; Gopher, et al. 1996). The ceramic sequence from Munhata 2a provided a contextual sequence and was the primary comparative collection during the 1990’s, despite dissimilarities to the original Kaplan sites of Wadi Rabah and ᶜEn Jarba (Banning 2007b). In addition to previous material culture remains, recent excavations at Neve Yam, a submerged site off the Carmel coast, have yielded evidence of a potential off-site burial ground associated with the village (Galili et al. 2009). Nahal Zehora I provided evidence of potential secondary product use during the Early Chalcolithic with spindle whorls and loom weights on site, as well as a potential ceramic churn (Gopher and Gophna 1993). Ceramic sherds thought to be spindle whorls were found at Tabaqat al-Bûma (Gibbs 2008a) and similar evidence was found at Basatîn (Kadowaki, et al. 2008).

More recently, use of the term Wadi Rabah as a ‘culture’ has come under fire (Banning 2007a). In addition, a high degree of regional variation makes it difficult to determine whether, across the entire southern Levant, it represents a consistent and visible culture or it is a modern construct used to bridge a gap in our own knowledge between the Late Neolithic and Chalcolithic periods. Despite these difficulties, sites associated with the Wadi Rabah pre-date Tel Tsaf and provide us with starting point for understanding the Middle Chalcolithic.

2.4 The Middle Chalcolithic

Unsurprisingly, given the lack of consensus regarding the classification of this ‘period’ there is limited literature that attempts to define the Middle Chalcolithic. In his attempt to present a comprehensive synthesis of the early pottery of the southern Levant, Garfinkel (1999) defined what he believed to be a clear Middle Chalcolithic phase. He presented the ceramics of this period as distinct from both Late Neolithic and Late Chalcolithic collections and being represented by a northern tradition (Beth Shean Ware) and a southern tradition (Qatifian Ware). His Middle Chalcolithic tradition in the north was based largely on what he calls Beth Shean Ware, which he associates with Fitzgerald’s Stratum XVIII at Beth Shean (Garfinkel 1999). He also presented sherds from 22 other sites that show what he views as similar features. In general, he suggests Beth Shean Ware is handmade and of poor quality, with the exception of the Tel Tsaf pottery. The distinctive red net design on sherds from Tel Tsaf is not a common motif,

21 having only been identified at a handful of sites within the Jordan Valley (Garfinkel 1999: 186)(figure 2.3). Garfinkel notes this higher-quality pottery is likely a local tradition within what he defined as the Middle Chalcolithic sequence. The southern tradition, Qatifian Ware, had previously been (and often still is) considered a Wadi Rabah variant (Gopher and Gophna 1993). Many scholars disagree with Garfinkel’s classification system, noting methodological and interpretive errors, as discussed below (Anfinset, et al. 2001; Banning 2001a, 2007a; Blackham 2002; Bourke 2007; Braun 2008; Gilead 2007, 2011; Lovell 2001b).

Figure 2.3: An example of the distinctive Tsafian design on a sherd from Tel Tsaf.

The ceramic evidence used to define the period was largely derived from previously excavated (legacy) sites or small surveys or salvage operations. Abu Hamid II likely represents a Middle Chalcolithic occupation, providing the largest ceramic evidence, as well as radiocarbon dates (Banning 2007b; Garfinkel 1999). The excavations at Tel Tsaf (2005-2007) represent large- scale excavation of a Middle Chalcolithic site with good architectural preservation, helping to define the period further. Garfinkel cites elements such as large storage silos, evidence for long- distance trade networks and architectural forms such as large silos, walled courtyard complexes and a combination of circular and rectangular rooms as further defining features Garfinkel and Ben Shlomo 2007; Garfinkel, et al. 2007, 2009).

22

Use of Garfinkel’s term Middle Chalcolithic has been limited and even those scholars who use the term do not consistently agree what sites or features fit within it. Some scholars have outright omitted the Middle Chalcolithic period, preferring to refer to it as Late Neolithic (Anfinset et al. 2011) or Transitionary (Gilead 2007, 2011). Bourke (2007) views the Early Chalcolithic as starting earlier in the north than the south, and avoids using the term the Middle Chalcolithic. Blackham (2002), Lovell (2001) and Banning (2001; 2007) have accepted the use of the term but do not agree upon dates or sites to be included. Notably, Blackham argues that the middle Abu Hamid phases are Middle Chalcolithic but Lovell places them prior to her Middle Chalcolithic ceramics Ghassul F-D that Blackham dates to the Early Chalcolithic. Banning generally accepts the term Middle Chalcolithic, although he does highlight the inappropriate classification by Garfinkel (1999) of Tabaqat al-Buma within this period, noting that it should be Pottery Neolithic based on both context and ceramic form (Banning 2001). Like Blackham, both Garfinkel (1999) and Banning (2007) consider Abu Hamid II to be Middle Chalcolithic.

A more cutting critique of Garfinkel’s definition of Middle Chalcolithic comes from E. Braun (2008). Braun acknowledges that the period itself may exist between the Wadi Rabah (Early Chalcolithic) and Ghassulian (Late Chalcolithic), but he does not agree with any of Garfinkel’s indicators. Having conducted his own in-depth study of material from Beth Shean, Braun concludes that Garfinkel’s method of analysis is uncritical and ill-conceived (2008). He is particularly concerned by the lack of consistency in the use of the term ware, as well as the quality of the material and stratigraphic integrity upon which Garfinkel based his Beth Shean Ware. Braun further claims that none of the distinguishing ceramic features used to define Beth Shean Ware are uniquely Chalcolithic, and also argues that other defining features, including circular silos, infant jar burials and basalt chalices, are found both before and after the period. Braun raises some valid points, but wholly dismissing the evidence while still acknowledging the period may exist chronologically has not contributed to our understanding of the Middle Chalcolithic.

2.5 The Late Chalcolithic Period

Most discussions of the Chalcolithic as a period focus on its later portion, the Late Chalcolithic. There is substantial agreement regarding the classification of the Late Chalcolithic, and this

23 period is represented by a variety of sites, including residential, cemetery and ritual sites in addition to cave caches. I will begin by providing a general overview of the most common features and then address current debates regarding social-economic organization during the period.

2.5.1 Material Culture

Changes in material culture and technology are prevalent during the Late Chalcolithic period. The three main areas of distinctive innovation are ceramics, lithics and metal. Ceramic technology had been introduced during the Late Neolithic and was widely used during the earlier Chalcolithic phases, but the Late Chalcolithic shows an advance in technology and a proliferation of forms introduced during the Late Chalcolithic. Lithics decline in prevalence at sites during the period, but there are distinctive forms and evidence of central production. Copper items were introduced during the Late Chalcolithic and evidence of smelting is apparent at some sites. Other metals, such as gold and electrum, also played a part in Late Chalcolithic society. There is variation in both quantity and form of material culture through the region and the discussion below focuses on the most commonly encountered artefacts from the period.

Pottery represents the most abundant artifact type reported at most sites, demonstrating the importance of ceramic vessels during the Chalcolithic. Two broad ceramic entities have been identified during the Chalcolithic (Roux and Courty 2005). Both forms involve coiling and their fabrics have a course granular aspect. Entity A has a thick clay layer on both the internal and external surfaces. This coating is lacking on Entity B ceramics. Ceramics were crafted by hand, sometimes supplemented by Rotative Kinetic Energy or technologies (Roux and Courty 2005). There was a great expansion in forms and styles, with open and closed-rim jars being produced in various sizes. Painted, applied and impressed decoration became more frequent and regionally diverse.

Forms ranged from basic everyday pots, both open and closed versions, to more elaborate vessels, such as fenestrated stands and anthropomorphic jars. Thin-walled V-shaped bowls are the most abundant form and are found in both residential and burial contexts. Churns also represent a distinctive Late Chalcolithic form. It has been suggested that the churns, which coincide with more intensive animal husbandry in the region, were used to produce yogurt or butter (Grigson 1998), tasks that may have been performed with ceramics using organic

24 materials such as animal hides, that would not have left evidence in the archaeological record. Residue analysis supports this idea, as gas chromatography/mass spectrometry has detected fatty acids similar to milk fat (Burton 2004). Potential evidence of other ceramic forms having similar uses has been found at Basatîn where molecular and isotopic evidence on ceramic sherds suggest either cooking of animal meats or potential storage of dairy products (Gregg, et al. 2009). “Cornets”, narrow conical cups with pointed bases, are another distinctively Chalcolithic form that declines in popularity by the end of the period. Excavations at Ghassul and Abu Hof demonstrate this decline as they are abundant in early phases and scarcer by the end of Chalcolithic occupation (Burton 2004).

Chipped- and ground-stone assemblages show continuity from the Neolithic. There was a decline in the production and use of chipped-stone tools, with a notable decrease in arrowheads (although transverse arrowheads are common in the Negev and Sinai) and a focus on scrapers, sickles and other retouched blades (S. A. Rosen 1997; Rowan and Golden 2009). Most tools seem to be produced locally, but there is some evidence of specialized production of blades and fan scrapers at some sites (Gilead, et al. 2004; Roshwarb 1981; Rosen 1997). Bifacial lithic discs with holes in the centre appear in the Southern Levant during the Chalcolithic. These discs, sometimes star-shaped, have no clearly identifiable purpose but given the amount of effort required to produce them and the lack of identifiable use, it has been suggested they may have a symbolic or ritual purpose (Rosen 1997). They are a wholly Chalcolithic phenomenon (Rosen 1997). Basalt bowls, often mimicking the ceramic V-shaped bowls, and basalt figurines are also abundant during the Chalcolithic.

Copper items made their first substantial appearance in the region during the Late Chalcolithic. There is evidence of smelting from sites such as Shiqmim and Abu Matar, as well as specialized mining and smelting facilities in the Wadi Faynan region, demonstrating that processing was taking place in the Levant (Golden 2009; Golden, et al. 2001; Shugar 2001). Utilitarian artifacts that Chalcolithic coppersmiths produced include axes, adzes, chisels and awls. Despite their design as tools, it is debatable whether they were used as such implements. The functional copper artifacts are most frequently pure copper, making them soft and less effective than their lithic counterparts. However, using SEM, Greenfield (1999, 2000, 2013) has identified copper- produced cut marks on bone from the Neolithic in the Balkans, demonstrating that copper tools may have been more efficient than is commonly believed.

25

Copper is most famously associated with Nahal Mishmar, also known as the Cave of the Treasure. This cave site along the western bank of the Dead Sea produced a plethora of copper artefacts in forms including maceheads, ‘crowns’ and ‘sceptres’ (figure 2.4).

Figure 2.4: Examples of copper objects from the Chalcolithic. A and B are from Nahal Mishmar. C is a standard from Givat Ha-Oranim (Rowan and Golden 2009)

The cave also contained ivory and haematite items. The nature of the artifacts, combined with their remote location, has led to much discussion regarding the nature of the hoard. A second cave, Nahal Qanah in the western highlands, is often included in this debate. In addition to copper items, burials and basalt items, gold and electrum (a naturally-occurring alloy of gold and silver, with a minimum of 20% of either element, often also containing other metals such as copper (Das and Zonderhuis 1964) rings were recovered. This is seen as one of the greatest pieces of evidence for ranked societies during the Chalcolithic, which will be discussed below (section 2.7).

26

2.5.2 Burial Practices

Burial practices can highlight social, ideological and economic variations among Late Chalcolithic sites. A number of Chalcolithic burial caves include the Cave of the Warrior (Schick 1998); Peqi’in (Gal, et al. 1996, 1997); Nahal Qanah (Gopher and Tsuk 1991, 1996) and Azor (Ben-Tor 1975), Kissufim (Goren and Fabian 2002) and numerous others (van der Brink 1998). There is also evidence of cemeteries and constructed tombs from the Negev (Bar-Yosef, et al. 1977; 1986), Shiqmim (Levy and Alon 1985b, 1987) and Jericho (Kenyon 1960). All of the burial sites discovered date to the Late Chalcolithic. Once more, regional variation comes into play, making it difficult to identify a single unifying burial tradition for the period. Despite this, a number of observations can be made. Late Chalcolithic burials are largely located away from settlements and can be found in natural caves or constructed tombs or graves. Where we do find examples of burials within settlements, they are usually burials of infants or children in ceramic jars within courtyards. Among formalized tombs, secondary burials are dominant, often with numerous individuals buried together and the skulls cached separately from the post-cranial skeletons (Gal et al. 1996, 1997; Levy 1982, 1987; Kenyon 1960). Ossuaries make their first appearance in the Chalcolithic and often take the form of houses or display anthropomorphic features (figure 2.5). Ceramics and basalt constitute the most frequent form of grave good, often including pedestalled v-shaped bowls, violin-shaped figurines and beads.

Figure 2.5: Clay ossuaries from Azor. (Levy 1986b)

27

2.5.3 Architecture

The dominant architectural form during the Late Chalcolithic is the rectangular house, similar to those identified at Wadi Rabah sites. They are often described as broad-room houses because the entrance is usually located along one of the long walls of the building. Broad-room houses range considerably in size and configuration both among and within settlements. Layout varies greatly from complexes formed by several adjoining broad-room houses (e.g., Teleilat Ghassul) to groups of broad-rooms surrounding courtyards (e.g., Shiqmim, Teleilat Ghassul) to scattered houses (e.g., )(Banning 2010). In the Golan, sites often exhibit what are called chain- houses, where building complexes are formed by long lines of adjoining broad-rooms (Epstein 1998). Subterranean structures have been identified at a number of sites, such as Shiqmim, Tell Abu Matar, and Bir es-Safadi (Levy, et al. 1991). In these cases, it is often unclear if the subterranean rooms were used in conjunction with above-ground buildings, or if they were independent houses with no superstructure. There is also debate as to whether they were dwellings at all, with suggestions they may have been used exclusively for storage (Gilead 1994; Levy 1993) or potentially as tombs (Rowan and Ilan 2013).

The Late Chalcolithic presents evidence of buildings that also stray in size and plan from the broad-room house, such as ones at Teleilat Ghassul, Shiqmim, Tel as-Safadi and Byblos (Banning 2010). The most notable is a complex at Teleilat Ghassul that appears to have multiple rooms and covers an area of 175 m2. In addition to larger architectural features within sites, at least one ‘temple’ site, ʻEin Gedi, has been dated to the Late Chalcolithic. This one building site stands in isolation atop a large hill overlooking the Dead Sea. The lack of domestic artifact evidence, combined with the location and scale of the complex plan has led most researchers to accept ʻEin Gedi as a ritual site (Rowan and Golden 2009). Another notable structure has also been uncovered within the settlement at Teleilat Ghassul. During excavations of Mound 5, Area E, Hennessy declared the area was likely a sanctuary based on the lack of domestic surface finds and the large isolated architecture (J.B. Hennessy 1982; Seaton 2008). The shrine at Gilat appears to be composed of a complex of large buildings that surround a large open courtyard (Levy 2006; Rowan and Golden 2009). It is unclear whether the entire complex is ceremonial or if it also represents domestic space.

28

2.6 Beyond the Southern Levant

This dissertation focusses on the Southern Levant, but it is important to understand how this area relates to contemporary sites in the larger region. Table 2.4 provides an overview of important cultural phases outside the Southern Levant.

Southern Levant Byblos Upper Mesopotamia Lower Mesopotamia 3400 Early Bronze Enéolithique Late Uruk Late Chalcolithic/ 3600 Uruk Néolithique 3800 Récent Early Uruk

4000 Late Chalcolithic Late Northern Ubaid 4200 Ubaid 4

4400 Néolithique Moyen 4600 Ubaid 3 calBC Early Northern Ubaid 4800 Middle Chalcolithic 5000 Late Halaf 5200 Ubaid 2 Middle Halaf 5400 Néolithique 5600 Early Chalcolithic Ancien Early Halaf

5800 Hassuna Ubaid 1

Table 2.4: This table outlines the major cultural periods of the areas immediately surrounding the Southern Levant. The dates used for the Southern Levant are based on Garfinkel (1999) as this is the chronology employed in this dissertation. The Byblos chronology is from de Contenson (1966) and the Upper/Lower Mesopotamia classifications are from Wilkinson (2000).

Areas outside the Southern Levant have faced similar issues to those addressed here. The dates and phases represented above are by no means wholly accepted within the archaeological community and attempts have been made to revise some classifications (Campbell 2007). There

29 are also some issues with the use of cultural entities to define periods, and they are based on very early excavations, often from sites that do not actually represent typical examples of material culture in the regions or periods with which they are associated (Banning 2007a; Campbell 2007). The table is offered as a general reference to put the material discussed in this dissertation into a wider context across the Near East.

2.7 Social Organization

Scholars debate the nature of social and economic organization of the southern Levantine Chalcolithic, in particular, regarding the presence of chiefdoms. Some suggest that the appearance of ‘temple’ sites, such as ʻEin Gedi, and the presence of copper hoards suggest both elites and widespread warfare in the region (Gal, et al. 1996; Levy 1996a, 1998), leading to the first cities of the Early Bronze Age. Most scholars accept the evidence of hoards and temples as suggestive of changing social dynamics, but the nature and structure of this change is still hotly debated (Blackham 2002; Bourke 2001; Gilead 1988, 2011; Levy 2006; Lovell 2010; Philip 2011; Rowan and Golden 2009). A general overview of the background and discussions surrounding social organization in the Chalcolithic is presented below. This topic will be re- visited with specific reference to Tel Tsaf in Chapter 6.

The concept of chiefdoms is most often attributed to Marshall Sahlins (1958, 1960) and Elman Service (1962). Their work draws on a Marxist approach to economic development, with surplus freeing up workers for non-subsistence work, and eventual differential accumulation of wealth creating inequality in a society. In discussing general trends in cultural evolution, as he saw them, Sahlins (1960) introduced the idea that many societies progress through specific social organisational stages, progressing from lesser to greater integration and adaptability. Sahlins gave each of these evolutionary stages terms, from simple bands through to tribes, chiefdoms and states that then formed the foundation for Service’s work Primitive Social Organization (1962). Service’s model defines chiefdoms as differing from tribes in two important ways: greater population density and the presence of organized centers to co-ordinate economic, social and religious activities (1962:143). The concept of redistribution is central to a chiefdom level of organization because the chief’s main duty is the redistribution of surplus goods and services. A chief not only controls resources, he also organizes labour to ensure production or construction.

30

Service later stated that chiefdoms were essential to our understanding of the rise of state-level societies and implied that chiefs most likely played a prominent religious role (Service 1975).

Similar structures formed the foundation for Fried’s concept of rank societies (1967). Borrowing from political science, Fried shifts the focus from socially-driven redistributive organization to political organization. He suggests it is more appropriate to discuss egalitarian, rank, stratified and state societies as these categories better represent the political structure. He argues that political organization is one aspect of social organization that incorporates relationships and interactions that extend beyond the political purpose of managing the affairs of public policy (Fried 1967:21). He further states, when discussing social structures that dictate how a group is structured politically, that it is necessary to move away from broad social or cultural classifications in order to address the true nature of such structures. Fried suggests that rank societies allow access to positions of prestige, but not everyone with potential access to such positions achieves them (Fried 1967: 109). Such limitation to positions of valued status does not limit access to basic resources by the general population. In contrast, Fried views stratified societies as the initial step towards statehood (and inseparable from it), resulting in such societies being maintained by differential distribution of basic means of livelihood and eventual exploitation of human labour (Fried 1967: 185-6). Within his framework, rank societies share many of the same hallmarks as chiefdoms, with increased population density, redistribution and religious leadership all featuring prominently (Fried 1967: 113, 116, 137). Although he did not intend it to be used as a synonym for chiefdom, the term ranked society is often freely interchanged with chiefdom in archaeological literature today (1986b, 1995; Peebles and Kus 1977).

These socio-evolutionary approaches figure prominently in archaeological discourse from the 1960’s. More recent approaches criticise the lineal nature of this approach that focusses on progression through pre-determined social stages. There has been outright rejection of evolutionary approaches, most notably by Giddens (1984). He states that such approaches are ethnocentric and warns of the danger in attempting to compress into a specific evolutionary shape (1984:263). Others have suggested that the concept of chiefdoms is too broad and poorly defined to make it a useful term (Kristiansen 1991; Yoffee 1993). Despite these concerns, the concept of chiefdoms is still ubiquitous in archaeology. As these issues and concerns are addressed in the literature, there is ever diversified use of the term chiefdom, based

31 on expanding ethnographic and archaeological data (Carneiro 1998; Earle 1987, 1991; Feinman 1991; Haas 1982; Hayden 1995; Nelson 1995; Peebles and Kus 1977; Renfrew 1973; Shennan 1993; Stein 1998; Wright 1984). Offering a single, universally agreed upon definition is difficult, and so it is best to discuss the term within the context of the region being discussed.

Levy (1986b, 1995, 1998, Burton and Levy 2011) employs the chiefdom model for the Late Chalcolithic, but he rejects the use of the term as a pre-determined evolutionary stage (1995). He presents his own parameters, purportedly drawn largely from Renfrew (1973)(table 2.5).

Defining Features of Chiefdoms: a ranked society the redistribution of produce organized by the chief

Other Frequent Features of Chiefdoms ● greater population ● centres that ● improvement in craft density coordinate social and specialization religious/economic activity

● increase in population ● frequent ceremonies ● potential for territorial size and rituals serving expansion wide social purposes ● increase in the size ● rise of priesthood ● reduction of internal of individual residence strife groups ● greater productivity ● environmental ● pervasive inequality of situation favouring persons or groups in specialization in society associated production with permanent (redistribution) leadership, effective in fields other than economic

● more clearly defined ● specialization, not ● distinctive dress or territorial boundaries only regional or ornament for those of ecological, through high status pooling of individual skills in cooperative endeavours ● a more integrated ● organization and ● no true government to society with a greater deployment of public back up decisions by number of labour for agricultural legalized force sociocentric statuses work (irrigation), temples or other monumental architecture

Table 2.5: Renfrew’s key indicators of Chiefdoms within societies (1973). It can be argued that some of these points are not exclusive to chiefdoms, but they are included to provide an accurate summary of Renfrew’s work.

32

Renfrew’s work clearly lays out his template for evaluating whether a society can be defined as a chiefdom. Many of his ideas feature heavily in Levy’s work, but it is clear in Levy’s definition of the chiefdoms of the Chalcolithic in the southern Levant, that he has modified the list to fit the data available (Levy 1986, 1995). For example, key indicators, according to Levy, include two- tiered settlement system, craft specialization, evidence of prestige items, ceremonial centres and agricultural intensification. Indeed, excavation has produced large and small settlements, copper and gold objects and ingots, evidence of ritual sites, and grain silos, all dating to the Late Chalcolithic. All these elements play into Levy’s idea of Late Chalcolithic chiefdoms.

However, there are researchers who believe that the archaeological record for the Chalcolithic is too sparse or unclear to allow identification of chiefdoms (Gilead 1988; Bourke 2001; Joffe 2003). While it is generally agreed that there is evidence of large-scale storage, variation within site architecture and ceremonial sites such as ʻEin Gedi, the actual social implications of these are unclear for the Late Chalcolithic. Rowan and Golden describe the socio-economic organization during the period as “highly varied and loosely integrated” (2009:69). They suggest that the Beersheva region may have had some form of hierarchy within and among sites, but it is unlikely that influence extended beyond the local level. The regions to the north, east and south are lacking evidence of the same relationships, leading Rowan and Golden to believe that no region-wide hierarchy or uniform organizational system existed in Southwest Asia during the Chalcolithic period.

The Chalcolithic, including the Middle Chalcolithic, does demonstrate many new and interesting changes in settlement organization, with new technology and artifacts coming into the archaeological record, making it an exciting period but leaving plenty of room for further research. Tel Tsaf is situated in a time period that sees the beginnings of the hallmarks of the Late Chalcolithic. Discussions of social organization or potential chiefdoms need to include data from the Chalcolithic as a whole. It is the aim of this dissertation to contribute to the debate and advance our understanding of the social and economic structure of the Chalcolithic. A greater understanding of how Tel Tsaf functioned during the Middle Chalcolithic can contribute to the larger discussion about Chalcolithic society in the Levant, as addressed in Chapter 6.

33

2.8 Tel Tsaf

The site of Tel Tsaf is located on the western bank of the Jordan River in the central Jordan Valley, southeast of modern Beth Shean and near the modern Israel/Jordan border (figure 2.6). The site consists of three hills; the western hill being the largest (3.5 ha) (figure 2.7). The hills are located between the upper and lower terraces (270 – 280 m below sea level), and lie approximately 500 m south of the confluence of Nahal Bezeq tributary and the Jordan River. The site was first identified during a large-scale survey of the region by N. Tzori (1958) during the 1940’s and 1950’s.

Figure 2.6: The location of Tel Tsaf (drawn by T Woolcott).

Figure 2.7: Map outlining the three hills of Tel Tsaf. The Southern Hill contains Area B. The main excavation area and focus of this excavation is Area C, located on the eastern edge of the Western Hill. The Eastern Hill is located behind the border fence and has not been excavated (T. Woolcott after Garfinkel et al. 2007).

34

Military activity in the 1970’s and 1980’s caused damage to the eastern hill and a continued military presence in the area is witnessed today by a border fence and patrol road that run through the depression between the western and southern hills. The eastern hill is inaccessible today, being located beyond the border fence and hosting a military lookout post.

2.8.1 Regional Geology and Paleoclimate

The Jordan Valley is the upper extension of the Jordan Rift Valley. The Jordan River runs south from the Sea of Galilee to the northern tip of the Dead Sea, a distance of approximately 120 km. South of the Sea of Galilee the Jordan River is fed by two main rivers, the Yarmouk and Zarqa, running east-west from the hills around the valley. A number of smaller seasonal wadis also feed into the river.

The geology of the Jordan Valley is illustrated in figure 2.8. The valley floor in the region is dominated by alluvial deposits of gravel, sand and clay from the surrounding hillsides. There are a few areas of travertine deposits, most notably beneath modern Beth Shean, containing gravel, sand and silt. The hills forming the valley walls contain a conglomeration of late Cretaceous to Pliocene chalk, limestone, marl and chert, with three periods of volcanism resulting in basaltic flows, intrusions and volcanoclastics.

Tel Tsaf itself is located on the Lisan Formation (figure 2.8), which runs parallel to the north- south flowing Jordan River. Lake Lisan filled the Jordan Valley from the modern Sea of Galilee to the Dead Sea between 70,000 and 17,000 years ago, at which point it began receding to become the Dead Sea (Begin, et al. 1974; Begin, et al. 1980; Goldberg and Bar-Yosef 1995; Robinson, et al. 2006; Stein 2001). The remaining sediments, or the Lisan Formation, are a complex series of stratified sediments representing at least two predominant depositional environments (Bartov, et al. 2002): (i) a low-energy offshore environment with aragonite and silt-sized calcite, dolomite, and quartz; and (ii) a higher energy shore–delta environment with sandstone, gravel and conglomerates. Erosion by the Jordan River, as well as modern construction of roadways, has exposed significant cross sections of the Lisan Formation. Figure 2.9 shows one such exposure located north of Beth Shean in the Jordan Valley.

35

Figure 2.8: Geological map of the Northern Jordan Valley around Tel Tsaf (after Sneh, et al. 1998).

36

Figure 2.9: An exposed cross-section of the Lisan Formation, north of Beth Shean in the Jordan Valley (photo by A. Hill).

The modern climate in the Jordan Valley is seasonal. Summers are hot and dry, with temperatures averaging around 34°C, but often exceeding 40°C. Mild winters see average lows around 13°C and have historically recorded up to 800mm of rain in the Galilee region (Orni and Efrat 1971; Rosen 2007; Zohary 1962). Paleo-environmental studies in the region have established that the modern climate is very different from that during the prehistoric periods. Proxy data are used to infer climate conditions prior to historical record keeping. These correlated sets are derived from data related to natural phenomena directly affected by climatic shifts over times. In archaeology, pollen analysis, isotope analysis and geomorphology are three of the commonly used techniques and provide the bulk of the data for prehistoric climates in the southern Levant.

Pollen sequences from lake and bog cores provide a plethora of data for environmental reconstruction. The outer layer of a pollen grain is durable and preserves well in anaerobic environments, such as lake sediment. The abundance of pollen and its proliferation in the natural environment, combined with the unique structure of each species grains, allows researchers to

37 identify and quantify pollen at specific levels within cores (Rosen 2007). Pollen counts are plotted on diagrams that demonstrate the rise, fall or static pollen production of certain species over long periods of time (Faegri and Iversen 1989; Moore, et al. 1991). Changes in the abundance of certain species can be associated with changes in conditions that either favour or hinder the growth of those species. These trends allow researchers to identify pollen zones that indicate wetter or drier, warmer or cooler climatic conditions.

Oxygen isotopes have great potential for identifying climatic changes on a global scale, and increases in relative humidity levels more locally (Bar-Matthews, et al. 1997; Maher, et al. 2011; Rosen 2007; Weaver 2003). Two isotopes of oxygen are commonly measured: water molecules containing the more abundant isotope, 16O, are lighter and evaporate more readily than water containing the heavier and less abundant 18O isotope. Sea and ice cores incorporate fossilized 18 foraminifera that contain biogenic CaCO3 and their δ O values provide a reliable environmental proxy (Bradley 1999). This data highlights oceanic circulation, temperature and sea levels, and provides insight into glacial and interglacial phases (Maher, et al. 2011; Weaver 2003). In addition, terrestrial biogenic and abiogenic carbonates, such as land snails, lake sediments, and speleothems can also be used in this isotopic analysis (Bar-Matthews, et al. 1997; Bradley 1999; Rosen 2007).

Geomorphology allows researchers to identify how landscapes were formed, providing indications of whether specific strata were deposited by wind or water-laid on the micro-scale, or how valleys and plateaus were formed on a macro-scale. It is possible to identify evidence of flood plains, erosional episodes, and dry periods or landscape stability by examining colour, grain-size, sediment sorting, texture and composition etc. Visual examination of the landscape can identify stream terraces that indicate greater stream flows than current conditions, for example. Such analyses also give insight into the sea and lake levels and composition of a region, providing further information related to warming and cooling trends (Bartov 2002; Maher, et al. 2011).

Extensive research into paleoclimates has been conducted in the southern Levant, with much focus on the late Pleistocene through middle Holocene (Alley 2003; Bar-Matthews 2003; Bar- Matthews, et al. 1997; Bar-Matthews, et al. 1999; Rambeau, et al. 2011; Robinson, et al. 2006; Rossignol-Strick 1993; Severinghaus 1999, 1998; Thomas, et al. 2007; Torfstein, et al. 2013;

38

Weaver 2003). Numerous lines of data have been presented, creating a good framework for environmental interpretation. It should be noted, however, that many authors freely admit that there are potential pitfalls to applying localized data to a wider region or to assuming that global climate trends are direct proxies for regional climates (Maher, et al. 2011; Rosen 2007; Rambeau 2010). Data for the Southern Levant has been collected from a wide range of sites, both geographically and climatically. The Jordan Valley and Tel Tsaf lie in one of many microclimates that are still seen today in the region. General trends may be explored and new evidence from the Jordan Valley is still emerging (Litt, et al. 2012; Stein, et al. 2010), but there is no environmental data directly linked to the Jordan River levels and flooding events during the Chalcolithic period.

The onset of the middle Holocene (c. 5500 cal BC) coincides more or less with the Early Chalcolithic in the southern Levant (c. 5200 cal BC), but it should not be assumed that there is a direct correlation between environmental and cultural phenomenon (Maher, et al. 2011). The preceding early Holocene was marked by rapid warming following the dry, cool Younger Dryas event at the late Pleistocene. Marine and terrestrial pollen samples from across the Mediterranean indicate an increase in pistachio and oak (Rossignol-Strick 1999). Evidence from Lake Hula cores and the more recent Ain Gedi core, indicate oak reached its peak around 6900 cal BC, indicating warm wet conditions favoured by deciduous forests (Litt, et al. 2012). An increase in pistachio further supports this, evidencing warmer winters in the Mediterranean region (Rossignol-Strick 1999). Isotopic analysis of speleothems from Soreq Cave in central Israel yielded a reduction in δ18O values (Bar-Matthews, et al. 1999), indicating increased rainfall and moisture in the region. Lacustrine evidence indicates the early Holocene saw a period of increased lake levels in the Dead Sea.

The middle Holocene proxy data indicate a shift away from the climate trends of the early Holocene. Pollen data from Birkat Ram in the Golan Heights indicate a sharp decline in Oak around 4500 cal BC, followed by a steady increase through the Late Chalcolithic period (Schwab, et al. 2004). The brief decline may be an indicator of a short-lived cooling period coinciding with the later occupation of Tel Tsaf. Evidence from Lake Hula and Ghab show an increase in Olive pollen from 6200 cal BC, likely indicating the beginning of olive cultivation (Baruch 1999; Yasuda, et al. 2000).

39

Isotopic evidence from the Soreq Caves show that the middle Holocene is characterized by significant variation in wet and dry phases creating unpredictable farming environments (Rosen 2007). The δ18O isotopic data from the Soreq Caves indicates a sudden shift to much drier conditions c. 5000 cal BC, contrasting with the warm, moist early Holocene. The climate then returns to a more moist environment c. 4500 cal BC, but does not reach the climatic optimum of the early Holocene (Bar-Matthews and Ayalon 2004). The probable occupation dates of Tel Tsaf (5040 – 4440 cal BC) fall almost entirely within the return to drier conditions. The Late Chalcolithic and Early Bronze Age data suggest alternating moist and dry episodes, creating environmental instability in the Southern Levant (Rosen 2007).

Geomorphological evidence from the central Jordan Valley is lacking for the middle Holocene period, but studies from the Shephela and coastal plain of Israel (Rosen 1986a, b, 1991; 1997) and the Negev (Goldberg 1987; Goldberg 1994) provide potential proxy data. Stream sediment deposits found within a system of 2m - 3m terraces indicate steady flow rates and increased stream activity (Rosen 1986a, b, 1991, 1997). Thick silt and fine sand laminations lack coarse- grained, poorly sorted sands and gravels associated with flash flooding, suggesting increased, regular rainfall is responsible for the stream flow and seasonal flooding of the surrounding plains. According to Rosen (1991) the deposits date to the Chalcolithic (Late Chalcolithic) and Early Bronze I based on associated ceramic and radiocarbon data. Evidence from the Late Chalcolithic in the Negev comes from a similar terrace system near Shiqmim in the Nahal Beer Sheva wadi (Goldberg 1987, 1994). The results indicate similar increased stream flow and regular rainfall during the Late Chalcolithic. None of this data directly dates to the Tel Tsaf occupation, but one may assume that drier conditions would have resulted in reduced stream activity and potential flash flooding rather than steading flow rates.

One final environmental proxy data set comes from the analysis of lacustrine evidence from Dead Sea lake cores. Depositional history the sediment depositional sequence, notably the presence of halite, provides indicators of lake levels (Neev 1967). The Dead Sea levels were high around 9300 cal. BC, with increased halite deposits following this indicating a reduction in levels (Neev 1967; Neev and Hall 1977). These levels do not appear to rise again until the Late Chalcolithic period, indicating that the Tel Tsaf occupation coincides with lower lake levels. This is support by Frumkin and colleague’s (Frumkin 1997; Frumkin, et al. 1994) analyses of rafted wood samples from Mount Sedom caves. An absence of wood from c. 5800 to 4200 cal.

40

BC suggests low lake levels, with a marked increase in samples between 4200 and 2300 cal. BC indicating a rise in the Dead Sea levels. The Tel Tsaf occupation falls within the low level period. Klinger, et al. (2003) present geomorphological evidence from the alluvial fans of the Dead Sea basin that indicate terraces formed during low lake levels, with one dating to 5000 – 4200 cal BC.

The proxy data from the Mediterranean and southern Levant suggest that the Middle Chalcolithic period climate was relatively dry with reduced rainfall and falling Dead Sea water levels. Either side of this period is marked by increased rainfall and conditions more favourable to farming. There was likely an associated drop in the Jordan River water levels during the Middle Chalcolithic period, and less consistent flooding which may have affected agricultural production. Historically, the Jordan River averaged an annual streamflow of c. 1400 MCM (million cubic metres) when it entered the Dead Sea. Damming and extensive irrigation projects have reduced the annual streamflow to around 400 MCM (Salāmah and Bannayan 1993). No specific environmental data regarding streamflow rates during the Chalcolithic are available but given the indications of a dry climate with Dead Sea reduction, it should be assumed that it was likely less than 1400 MCM as the climate improved steadily through the modern period, but likely higher than the current 400 MCM, as there is no indication of man-made irrigation at the time.

Micromorphology samples from in situ Lisan Formation sediments in the lower fields near adjacent to Tel Tsaf contain well-lain, stratified calcareous clays (see Chapter 3.3.3) that lie on Quaternary alluvium. The Lisan sediments and alluvium deposits run parallel to the Jordan river through much of the Jordan Valley, which is noted for its fertility, being called the bread basket of the region in modern times (Al-Weshah 2000). Modern growth is greatly aided by irrigation, but the surface geology has remained relatively stable during the mid - late Quaternary. Today, the area surrounding Tel Tsaf is extensively used for agricultural production. Tel Tsaf’s proximity to the Jordan River combined with the fertile natural sediment, would have allowed successful crop cultivation even during drier periods. The Middle Chalcolithic likely experienced variable weather and flooding events year to year, but it is reasonable to assume the climate and geological factors would have allowed the occupants of Tel Tsaf to engage in significant, although potentially inconsistent, farming.

41

2.8.2 History of Excavation

Between 1977 and 1989, three seasons of small-scale excavations were carried out by R. Gophna (Gophna and Sadeh 1988/1989). Gophna excavated eight 5 m x 5 m squares on the eastern edge of the western hill, including two test pits dug down to virgin soil. This revealed a wealth of artefacts, including pottery, flint, faunal remains and charred wood. Excavations also uncovered well-preserved mudbrick architecture. Extensive survey of the region in 2000-2001 by Kohn- Tavor further refined our understanding of the extent of the site, suggesting that it was c. 20 ha in size with evidence of occupation beyond the three hills to the north and west (Garfinkel et al. 2007).

Interest in the site renewed in 2004, when a team directed by Y. Garfinkel from the Institute of Archaeology, Hebrew University, conducted survey and test excavations at the site. The team opened up 26 5 m x 5 m squares in various parts of the site, including the western and southern hills, the nearby area of Tel Jema’in, and areas to the north and west of the central hills (figure 2.7). The finds included hundreds of ceramic sherds and flint artefacts occurring in various densities that provided a better understanding of the boundaries of the site. The western hill yielded the best evidence for dense population, suggesting it was the main occupation area of Tel Tsaf.

Following the survey season, Garfinkel conducted three seasons of large-scale excavation between 2005 and 2007. The first season uncovered a well dug into riverbank deposits and several occupation surfaces on the southern and western hills. The second season extended horizontal exposures of the occupation area on the western hill, while the final season in 2007 focused on the lowest occupational horizons within Area C.

2.8.3 Area B

Area B, located on the upper portion of the southern hill, was excavated in 2004 and 2005 (Garfinkel et al. 2007). A well was visible during the 2004 survey, as almost 2 m of the shaft had been exposed by road cuts and erosion (Garfinkel et al. 2007). The shaft is approximately 5 m deep and 1.2 m in diameter (figure 2.10). At the base of the well, 20 partially damaged ceramic vessels were discovered. The vessels included three bowls and small and medium jars, many with four handles around the outside circumference. The sizes of the jars and the handle

42 placement suggest that these were used to draw water from the well, with the four handles stabilising the vessels as they were lowered by rope into the well shaft. It appears that the shaft was filled intentionally once out of use, possibly in response to contamination by a goat whose remains were found at the bottom of the well (Garfinkel et al. 2007). It is also possible that the goat was sacrificed when the well was closed.

Figure 2.10: Area B well (circled). The shaft was originally exposed by a road cut, and was excavated during the 2004 season (after Garfinkel et al. 2007).

At the top of well, on the surface of Area B, Garfinkel et al. (2007) report a debris scatter, probably representing a living surface. This area was located about 1 m from the mouth of the well, and extended almost 6 m to the east. Excavations during the 2005 season expanded to cover an area of 5 m x 7.5 m over this scatter. Various fragments of limestone and basalt implements were found, including vessels, grinding stones and cores. Furthermore, there were concentrations of stones, grinding implements, burnt animal bone, and charcoal, which Garfinkel et al. (2007) interpret as the remains of several hearths. However, there was no architecture in relation to any of these features and it appears, instead, to be an open-air work area. There was

43 also a plethora of other artifacts in the area, including pottery, animal bones, obsidian, shell, imported minerals, and a zoomorphic figurine. An abundance of beads, sometimes unfinished, fashioned from raw greenstone minerals and accompanied by flint drills, further suggests this was a production area.

2.8.4 Area C

Excavations began in Area C during the 2005 season. There appear to be at least two Chalcolithic occupation phases in this area; Phase 3 and Phase 4 (figures 2.11 and 2.12). Garfinkel has further divided the occupation into sub-phases 3a, 3b and 4a, 4b, 4c (table 2.6). The rationale for the sub-phasing is not clearly outlined in the published material so in this overview I will discuss the excavations in relation to the main Phases 3 and 4 exclusively, rather than referring to sub-phases.

Each phase is represented by a series of building complexes. Each complex consists of an enclosing wall surrounding rooms, probable silos (see Chapter 4.2 for discussion) and other features. A total of four complexes have been identified by Garfinkel et al. (2007, 2009). Building Complex I is the only building that spans Phases 4 and 3. Building Complex II is labeled as Phase 3 (figure 2.11) and Building Complex III and Building Complex IV are assigned to Phase 4 (figure 2.12).

During the 2005 excavations, architecture from Phase 4 was uncovered near the southeast corner of the Tel. A number of rectangular rooms were uncovered along with a number of paved circular features (figure 2.13). As I was not present during this season, no micromorphological or soil samples were taken from the southern portion of the excavated area (eventually Phase 4, Building Complex III). Phases 3 and 4 in Building Complexes I and II, IV, however, were sampled extensively and form the focus of the present analysis (discussed below).

44

Architectural features

Phase Phase

Period (Garfinkel et al) et (Garfinkel

1? 2 intrusive burials, Floor surface NE corner of area C Late Late Byzantine

2? pottery, Early Islamic coin Early Islamic Wind deposited soil, over 1m deep Building I: 1 rectangular Building II: 2 circular 3a room (70); 4 silos; 9 rooms (230, 263); 2 silos; roasting pits small enclosure

Building I: 1 Building II: 3 circular 3b rectangular room (70); 4 rooms (230, 2 new silos; 9 roasting pits unnamed); 3 silos

Building I: 1 Building IV: 1 rectangular rectangular room (70); 5 room (612); 3 silos; 2 low 4a silos, nothing N of room walls dividing N and S 70 courtyard

Middle Middle Chalcolithic Building IV: 2 rectangular Building I: 1 rooms (612; 662); 2 silos; 4b rectangular room (70); 5 1 low wall dividing N and S silos courtyard

Building III: multi-celled building; 1 silo; paved 4c area; oven?

Tentatively identified by Gopher and Sadeh (1988- ? 89), based on pits dug into virgin soil, 1 Pottery Neolithic sherd Neolithic

Table 2.6: Overview of phases for Area C at Tel Tsaf as observed by Garfinkel et al. (2007a, b; 2009).

45

Figure 2.11: Phase 3 at Tel Tsaf (drawn by T. Woolcott after Ben-Shlomo et al. 2009).

46

Figure 2.12: Phase 4 at Tel Tsaf (drawn by T. Woolcott after Garfinkel pers comm).

47

Figure 2.13: Site plan of the area exposed during the 2005 excavation. The southern portion of Building I was exposed, as well as the excavated extent of Building III. No soil samples were taken during these excavations. (drawn by T. Woolcott after Garfinkel pers comm.)

2.8.4.1 Building Complex I

Building Complex I shows very little change in layout between Phases 4 and 3 (figures 2.11 and 2.12). The most prominent feature is Broadroom 70, measuring approximately 10 m x 5 m (figure 2.14). This room presents at least four plastered surfaces during its use and, beyond a slight alteration in the construction of the wall between phases, there is no other apparent gap in use or occupation.

48

Figure 2.14: Aeriall view of the eastern courtyard of Building Complex I. Silos are the round structures along the top and left.

Figure 2.15: Roasting pit from Building I, Phase 3. This is a good example of a stone-lined pit. (Ben-Shlomo et al. 2009)

49

One of the key features of Phase 3 at Building Complex I is the abundance of earthen ovens or roasting pits within the courtyard. These are not present during Phase 4 and do not occur at all in Building Complex II. The bean-shaped pits were often lined with stone or mudbrick and contained levels of fine, ashy sediment (figure 2.15).

Also located within the courtyard are several brick-lined silos ranging in size from 1.8 m to 3 m in diameter (figure 2.16). These features were present in both Phases 3 and 4 and were clearly distinguished from rooms as they were lined with mudbrick. Silos were prominent at the site and demonstrate that the ability to store large quantities of goods was important. Their presence plays a major role in the discussion of the site of Tel Tsaf, and will be addressed in more detail in later chapters.

Figure 2.16: Silo 339, Building I. This is a good example of how the silos at Tel Tsaf were constructed. This photo was taken following the excavation of a burial by our team, explaining the cut down the middle. The silos were normally fully paved.

Building Complex I contained two adult burials, both associated with silos in the courtyard. The first burial was uncovered during the 2005 season and was located next to Silo 74 in the southwest corner of the courtyard (Garfinkel, et al. 2009; Garfinkel and Ben Shlomo 2007; Garfinkel, et al. 2007). The body was flexed and had no associated grave goods. The proximity to Phase 4 Silo 74 combined with post-depositional disturbance by a Phase 3 roasting pit dates this burial to Phase 4. The second burial was uncovered in the centre of Silo 339, located in the north of the courtyard (Garfinkel, et al. 2009). The body was flexed and found with 1668 beads

50 in situ around the waist (figure 2.17) and a highly corroded copper awl near the chest (Garfinkel, et al. 2014). Silo 339 is associated with Phase 3.

Figure 2.17: Burial from Silo 339 (photo by A. Hill).

2.8.4.2 Building Complex II

Building Complex II (Phase 3) differs from Building Complex I most noticeably in the lack of any rectangular structures within the excavated area. Both complexes contain silos but the courtyard of Building Complex II encloses two circular rooms (230 and 263) (figure 2.18). The courtyard also has two silos, a number of clay installations, including rounded basins, and a low enclosure or wall along the western periphery.

During excavation, no roasting pits or ovens were uncovered within the walled courtyard or buildings. The circular rooms also lacked any evidence of plastered floors. Excavation uncovered two infant burials within the building complex. The first was a jar burial discovered east of Silo 568. The neonate was on its side, with knees flexed (Garfinkel, et al. 2009). The second infant, approximately one year old, was found in a pit cut into the Phase 4 level of Silo

51

568. Again, the body was flexed and there were no associated grave goods (Garfinkel, et al. 2009).

Figure 2.18: Final excavation photo of Building Complex II, Phase 3.

2.8.4.3 Building Complex IV

Building Complex IV, Phase 4 is located directly beneath Phase 3 of Building Complex II. The building is difficult to define clearly as there appears to be a spill of mudbrick into the courtyard area, making it difficult to distinguish its architecture (figure 2.19).

52

Figure 2.19: Final excavation photo of Building Complex II, Phase 4. The central silo (Silo 272) was intentionally sectioned which is why half is missing in this photo (photo by A. Hill).

During excavation, a large ashy level was uncovered directly east of Silo 548. The area was not confined to a pit or enclosed area and exhibited fragments of fuel ash slag (figure 2.20). Fuel ash slag is formed when the ash of spent fuel combines with siliceous material and it is very light in weight. It can be a by-product of smelting, but is just as likely to form in other contexts, such as bonfires or burnt buildings (McDonnell 1983). Along with the fuel ash slag, three fragments of limonite (iron ore, similar to haematite) were uncovered. It should also be noted that beneath the walls of Room 612 there was a clear burn layer with ash and phytoliths visible (figure 2.21).

53

Figure 2.20: The area directly east of Silo 272, Building IV contained a widespread ashy level (left). This area contained a number of fuel ash slag fragments (right)(both A. Hill)

Figure 2.21: Burn level located beneath the excavated boundary of room 612.

The only architectural features present in both Building Complex II and Building Complex IV are the silos. Silo 272 was present during all occupation phases, although it became smaller during Phase 3.

2.8.5 Summary of Excavations

As expected, ceramics comprised the largest artifact class uncovered across the site during all phases. The pottery has yet to be analysed but the majority of sherds appear to be locally made, often with Tsafian designs. Garfinkel has identified at least five sherds that he has called ‘Ubaid and suggests they may have been imported from northern Mesopotamia (Garfinkel, et al. 2007). This claim is speculative and based on a very limited sample of small body sherds. Other potential trade items may include sea shells (as yet unidentified), obsidian (less than ten pieces) and two basalt base fragments. A few small animal figurines and two anthropomorphic figurines

54 were uncovered. A small stone seal and a number of clay sealings were recovered from a pit dug into Building Complex II Room 230 (Phase 3).

The architecture and material finds from Tel Tsaf present an interesting starting point for an assessment of the Middle Chalcolithic. The significance of the ceramics has yet to be explored but there is a potential wealth of data. The variation in architectural forms between and within Phases 3 and 4 is tantalising. The remainder of this dissertation will present my attempt to gain further understanding of the site and its relationship to the larger region. Addressing the questions related to architectural form and use related to resource management at Tel Tsaf will then be put into the larger context of storage, redistribution, potential wealth and social organization during the Chalcolithic period in the region.

55

Chapter 3 Methods and Results

Investigating the use of space at household and community levels requires evidence that demonstrates the relationship between the built environment and human activities. The horizontal and vertical exposures at Tel Tsaf reveal a variety of architectural forms within its built community. Reconstructing the distribution of these forms and their use-life over time, combined with insights into how the site’s inhabitants used space, will enhance our understanding of the socio-economic relationships within and among households in Middle Chalcolithic society.

This chapter provides an assessment of the reconstructed geomorphological processes involved in site-formation and diagenesis at Tel Tsaf in order to shed light on the on-site human activities within various phases of occupation and site use. These large-scale geoarchaeological assessments are then refined in scale through the use of pH testing, bulk sediment analyses, and soil micromorphology. The results of these fine-grained analyses allow us to move between the macro- and micro-scale to begin discussion of the use of space as presented below. 3.1 Geoarchaeological Assessment and Site Formation

In Chapter 2, an overview of the dating and phasing of Tel Tsaf was presented. Before proceeding with the results of my analysis, I will revisit the table from the previous chapter with the addition of geoarchaeological observations. This will provide the reader with a better sense of site-formation processes at Tel Tsaf and an explanatory framework (table 3.1).

In table 3.1, Phases 1 and 2 are defined on the basis of surface finds and limited evidence from small test pits. They include Byzantine and Early Islamic artifacts, but no radiocarbon dates are available for these phases. Some micro-geoarchaeological assessments of floor surfaces associated with Byzantine artifacts are discussed in table 3.1, with additional detail in section 3.3.6.4. However, there does not appear to have been extensive occupation during these periods. A layer of purportedly windblown sediment, approximately 1 m in thickness, was removed from Area C prior to the 2005 excavation in order to access the Middle Chalcolithic levels below (Garfinkel, et al. 2007: 11).

56

Building Complex III yielded the only architectural evidence exposed during excavation of the earliest Chalcolithic Phase 4c, but micromorphological samples also identified an occupation level below Phase 4b in Building Complex IV. These samples revealed the presence of floor levels directly beneath the Phase 4b destruction layer. The relationship between these Building Complex IV Phase 4c deposits and the Building Complex III Phase 4c remains unclear at this time.

Phase 4b at the site is represented architecturally by Building Complex I and Building Complex IV. Building Complex IV is located in the northeastern portion of the site. The main feature associated with Building Complex IV is Silo 232 and the associated walls that extend from its southern end. The walls separate the silo from the area directly south, which appears to be unoccupied during this phase. The extent of other architectural features during this phase is unclear, but a few observations can be made. The uppermost level of Phase 4 (a) is dominated by what appears to be a mud-brick spill across the eastern portion of the complex, likely representing a destruction level (figure 3.1). It is reconstructed to have been only one course thick and is poorly defined, so it is unlikely to have been an in situ building wall. Two rooms (612 and 662) were cut into this spill during excavation, revealing the inconsistent nature of mud-brick distribution beneath the spill layer. As illustrated in figure 3.1, some areas have more than 1 layer of mud-brick, but these are not evenly distributed or laid in an intentional manner. Some of the bricks are not even horizontally oriented. Other sections of the Room 612 and Room 662 baulks did not contain any evidence of mud-bricks, further highlighting the enigmatic nature of the mud-brick layer.

57

Architectural features Geoarchaeological Observation

Phase Phase

Period (Garfinkel et al) et (Garfinkel

1? 2 intrusive burials, Floor surface NE corner of area C Late Late Byzantine

2? pottery, Early Islamic coin Early Islamic Wind deposited sediment, over 1m deep Building I: 1 Building II: 2 circular 3a rectangular room (70); 4 rooms (230, 263); 2 silos; silos; 9 roasting pits small enclosure

Building I: 1 Building II: 3 circular northern portion of room 70 rebuilt; leveling and replastering evidence from room 70; multiple floor surfaces evident in 3b rectangular room (70); 4 rooms (230, 2 new Courtyard building I through 3/4 silos; 9 roasting pits unnamed); 3 silos Mudbrick spill across NE portion of Building IV S courtyard of Building IV (directly beneath Phase 3b 230) shows no occupation evidence, fine well sorted clay/silt, clay coated sediment. Building I: 1 Building IV: 1 rectangular rectangular room (70); 5 room (612); 3 silos; 2 low 4a large, uncontained burn area between silo and mudbrick spill, Fuel Ash Slag and limonite found = production? Destruction? silos, nothing N of room walls dividing N and S 70 courtyard

Middle Middle Chalcolithic evidence of well sorted sediment with clay coating above occupation levels and below mudbrick spill. rooms 612, 662 cut through mudbrick spill - no visibly constructed walls, 662 is circular in site photos

Building IV: 2 rectangular Destruction level: thick burn line above final excavation level of rooms 612, 662. Building I: 1 rooms (612; 662); 2 silos; Evidence of floor level directly beneath burn level indicates pre-destruction occupation. Evidence of disturbed floor fragments 4b rectangular room (70); 5 1 low wall dividing N and S above = post-destruction occupation. silos courtyard Building I lower courtyard levels suggest Phase 4/3 was constructed on uninhabited soil.

Building III: multi-celled building; 1 silo; paved 4c area; oven?

Tentatively identified by Gopher and Sadeh (1988- 89), based on pits dug into virgin soil, 1 Pottery Neolithic sherd Neolithic

Table 3.1: Overview of Tel Tsaf occupation phases and Geoarchaeological Observations

54

Figure 3.1: The extent of the mud-brick spill is not clear, nor are the boundaries of Rooms 612 and 662. Insets A and B highlight the unoriented and inconsistent nature of mud-brick distribution. They also clearly demonstrate that the excavated boundaries of each room do not represent the extent of these areas, as the mud-brick does not represent walls.

Micromorphological analysis of these rooms reveals disturbed floor levels beneath the brick spill (section 3.3.7). It is unlikely that rooms 612 and 662 were utilized in their excavated form during Building Complex IV Phase 4 occupation, as their original extent has been obscured by the mud- brick spill that covers portions of the original building walls.

The lowest level of Phase 4b within Building Complex IV is marked by evidence of destructive fire, indicated by a thick layer of ash and charcoal running beneath the Phase 4 structures (see figure 2.19). The destruction level does not run into the courtyard area next to Silo 232, which was also unaffected by the mud-brick spill. It is clear that the northeast portion of Building Complex IV was more greatly affected by these events than was the northwestern area.

55

Building Complex I, Phase 4a-c contains the large rectangular Room 70 and five silos. There is no evidence that the destruction events affecting Building Complex IV had an impact on Building Complex I. This phase has five silos: 633, 66, 171, 74 and 415.

Phase 3 is represented architecturally by Building Complex I and Building Complex II. In Building Complex I, there was an episode of wall renovation of the northern portion of Room 70, with a new course of bricks laid on top of the Phase 4 wall (figure 3.2). This reconstruction does not significantly impact the shape or size of Room 70. The room remained the same size. It appears silos 633, 66, 171 and 74 went out of use and silos 286, 399 and 53 were constructed during Phase 3. Silo 415 remained in use.

Figure 3.2: Room 70, Building Complex I. The arrows indicate the Phase 4 and Phase 3 walls. The overall shape and general size of the room remains the same, but it is clear that there was some modification to the walls, likely during Phase 3.

Building Complex II in the northeast portion of the site is located directly on top of Building Complex IV and sees the continued use of Silo 232. Its dimensions, however, are smaller than the Phase 4 structure. Two circular rooms, 230 and 263 are constructed.

3.2 Sediment pH

Prior to discussing the microscopic analysis of the soils from Tel Tsaf, it is important to identify any indications of potential differential preservation across the site. Sediment pH was measured from various contexts throughout the areas under discussion in order to test if pH values and,

56

therefore, preservation potential for organic and inorganic material was consistent throughout the site’s deposits (figure 3.3). This ultimately allows us to assume that presence or absence of chemical and organic signatures from these deposits results from human activities, rather than differential diagenesis dictated by micro-level environmental parameters.

Figure 3.3: Plans of Area C, Phase 3 and 4 at Tel Tsaf. The dots mark the locations of bulk sediment samples used in the pH test discussed in text.

During excavation, bulk sediment samples were taken from various contexts within the site, as well as from the surrounding area, for microscopic analysis. The full details regarding bulk sample collection appear in Section 3.3. For the pH analysis, 13 bulk samples were chosen to represent varied contexts across the site, including surface samples from on and near the occupation area (figure 3.3).

Sediment pH analysis was carried out in the Department of Geology at the University of Toronto. From each bulk sample, a 50 g subsample of sediment was measured out and placed in a graduated glass cylinder. 50 ml of de-ionized water was added to the subsample and these were stirred until thoroughly mixed (approximately five minutes), then left to stand for three minutes. This process was repeated for five stirring and standing cycles, and then left to stand a final time for six minutes.

A Mettler Toledo Seveneasy pH meter was calibrated using three single-use 20 ml standard solutions of 4.01, 7.00 and 10.00 pH. The electrode was then inserted into the upper portion of

57

the sediment/water mixture. The electrode was left in the sample for six minutes before temperature and pH were recorded. This process was repeated for each selected Tel Tsaf sample.

The results for each sample are listed in the table below (table 3.2).

Sample pH Context GS07-7 7.84 surface sample, on Tell near road SS15B-8 7.89 70, fill level TT15A-3 7.84 70, plaster floor level TT15A-4 7.71 Building I courtyard, roasting pit, dark ashy soil TT16C-2 7.76 Building I, courtyard TT16C-3 8.06 Building I courtyard, roasting pit, dark soil UU15A-10 8.06 Interior 230, fill - ashy UU15A-7 7.96 Interior 230, ash and soil levels UU15D-4 7.77 Ashy layer UU16A-1 7.72 top fill from square, no visible architecture UU16C-18 7.88 Fill from silo section UU16D-5 7.91 south of 263, courtyard UU16D-6 7.84 263, fill above 'floor' level

Table 3.2: Soil pH results from Tel Tsaf. The context of each sample listed in the right hand column corresponds to those shown in figure 3.3

The pH on site was consistent throughout all contexts and was neutral with an average pH of 7.9. The arid conditions of the Jordan Valley, combined with neutral to slightly alkaline pH values for all measured samples, indicate that preservation of environmental and archaeological data at Tel Tsaf is good. The conditions are not dry enough to preserve most organic material but bone, phytoliths, shell and dung spherulites survive quite well. The lack of variation across context, including the roasting pits where one may expect the most discrepancy, suggests no significant differential preservation across the site. Thus, variations in organic and mineral matter are real, not a factor of post-depositional disturbance.

3.3 Micromorphology

Micromorphology was employed as the primary technique for the analysis of sediment from the site because it allows one to examine microscopic details of floor construction, organic and micro-artifact inclusions, and site-formation processes associated with living and destruction surfaces in situ (Courty, et al. 1989; Fitzpatrick 1993; Goldberg 1983, 1992; Goldberg and Berna 2010; Goldberg and Macphail 2006; Matthews 2003; Matthews, et al. 1997; 2006). This

58

technique can provide evidence that highlights the character of floor levels and the activities associated with their use. A variety of social and economic considerations contribute to choices made during construction and use of space. Microscopic analysis can highlight these choices, offering insight into this process. This method of analysis has great potential to identify activity areas on prehistoric sites (Matthews 1995; 2003; 2012; Matthews, et al. 1997; MacPhail, et al. 1997).

Micromorphology differs from other geoarchaeology techniques as it informs researchers of site formation processes and traces of activities that are unidentifiable at a macro-scale or in unconsolidated sediments. It provides context for taphonomic and depositional process and the relationship between sediment deposition and anthropogenic data (Goldberg 1992; Goldberg and Macphail 2006; Matthews, et al. 1997). Macro-scale artefact distribution has great potential to aid our understanding of the use of space but it has been acknowledged that artefacts found in household contexts are often not in their original locations. Within the life history of a house, there are three main stages: habitation, abandonment and post-abandonment (LaMotta and Schiffer 1999). During these phases, any number of accretion and depletion process can affect what is eventually found within archaeological deposits. Hayden and Cannon (1983) use an ethnographic study of modern Maya household disposal to demonstrate that “there is much more variability in refuse disposal behaviour than archaeologists generally assume, and that this variability must be understood if interpretations based on artefactual debris at the household level…are to be meaningful” (153). The integration of microstratigraphic and micromorphological analysis of samples taken from vertical exposures can identify the pre- and post-depositional history of archaeological deposits, as well as identifying microscopic evidence of on-site activities even when macro-evidence has been removed. Using micromorphology at Tel Tsaf offered the opportunity to use a new data set to assist in understanding the daily use of space at the site.

3.3.1 Sampling Strategy

The first season of micromorphology sampling took place in 2006. The excavation was aimed at horizontal exposure of the northern portion of Area C. The lack of previously exposed architecture within the excavation area required a flexible sampling strategy. The primary goal of sampling was to determine if micromorphology was an effective analytical tool for the

59

deposits at Tel Tsaf. The main archaeological questions were aimed at understanding construction of floor surfaces and the use of architecturally distinct areas.

Sampling in 2006 was opportunistic, with samples extracted when suitable contexts were exposed among the existing excavation squares. Contexts included the floor levels in Room C70, as well as Building Complex I courtyard sections exposed during the extraction of a Byzantine burial and walls of excavated roasting pits. When a large circular building was exposed during the excavation of Building Complex II, the southern portion was cordoned off to create a baulk specifically for micromorphology sampling. A total of 17 samples were taken, of which 15 were processed for analysis.

The results of 2006 demonstrated the suitability of micromorphology as an analytical tool at Tel Tsaf. Therefore, a more extensive and focused sampling strategy was employed during the 2007 season. The following contexts were specifically targeted for micromorphology sampling: 1) Building II, Room C268 2) Building II, Room C330, lower floor levels 3) Building II, courtyard 4) At least one silo feature 5) Building I, northern courtyard 6) Exposed Lisan Formation deposit

In addition to these specific locations, sampling continued to be opportunistic, with samples taken from any appropriate contexts as they were exposed. A total of 53 samples were taken during the 2007 season, and 41 of these were processed for analysis. A total of 56 (15 from 2006, 41 from 2007) micromorphology samples were analyzed for the present study (appendix I) and the results are discussed below.

3.3.2 Sample Extraction and Processing

All samples were collected from baulks and profiles after excavation. In preparation for sampling, an excavation baulk was first scraped with a palette knife or trowel to create a smooth surface, and to clean and expose the stratigraphy within the section. Samples were then cut from the section using a 4-inch Swiss Army knife blade. Each sample was cut in accordance with the constraints of the material and size of the baulk, but generally were 12 cm by 5 cm in size (figure

60

3.4). The block was then removed by hand, laid out on paper towel, and wrapped and secured with packing tape (figure 3.5). Each sample was labeled on the front with the top indicated.

Figure 3.4: Building Complex II, Room 230. Preparing a micromorphology block for extraction from the baulk in the southern portion of the room.

Figure 3.5: Building Complex II, Room 230. After extraction the micromophology sediment block laid out on paper towel in preparation for wrapping (assisted by K. Sayers).

61

Samples were sent to Spectrum Petrographics, Vancouver, WA for impregnation and slide preparation. All samples were impregnated at low temperature in a vacuum with a clear epoxy resin. The impregnated blocks were then cut and a thin slice of sediment was mounted onto large format slides (5 x 7 cm). These mounted sediment slices were then ground and polished to a thickness of 30 µm. The 2006 samples were left uncovered; however, the slides from the 2007 season were covered with a coverslip. The orientation of each sample was physically marked on each slide, as was the sample number.

Discussion of the micromorphology results below follow the format presented by Goldberg, Miller, et al. (2009) and begin with an overview of the most commonly found components in the Tel Tsaf samples. The distribution of these constituents, both at the sample and site level are considered along with orientation and composition to establish six microfacies. These microfacies represent the six identifiable depositional activities identified during this analysis. A discussion of each Building Complex follows, with specific reference to these microfacies.

3.3.3 Sample Components

Two geological samples were taken from non-archaeological sediment exposures near Tel Tsaf in order to identify a sediment profile for comparison. Samples GS07-1 and GS07-2 were taken from a test pit created in 2006 that exposed an in situ section of the local Lisan Formation (figure 3.6). GS07-1 consists of water-lain, stratified calcareous clays (figure 3.7). GS07-2 is composed of calcareous clays; however, the sediment is less consolidated. A final ‘geological’ sample, GS07-8, was taken from the exposed southern face of the Southern Hill of Tel Tsaf (figure 3.6). The sample was largely composed of a calcitic, silty clay with a notable coarser grained component of basalt and plagioclase feldspar-bearing gabbro. The igneous grains represent re- deposited sediment from within the Jordan Valley catchment. The presence of charcoal fragments within sample GS07-08 suggests that re-deposition of sediment from elsewhere in the valley (presumably upstream of the site, or from adjacent hillsides) included anthropogenic input. The hills of Tel Tsaf are situated on the Lisan Formation, and contain a mixture of Aeolian and fluvial deposits of the Jordan Valley alluvial material that derive from local limestone and calcitic sand, silt and clays, as well as igneous clasts and, as mentioned, some anthropogenic input.

62

Figure 3.6: Regional map of Tel Tsaf with location of micromorphology samples GS07-1, GS07-2 and GS07-8 indicated (after Garfinkel et al. 2007).

Figure 3.7: Water-laid, calcareous clays from the Lisan Formation near Tel Tsaf (GS07-1, Left: PPL, Right: XPL, 50 X).

The main components of the Tel Tsaf, Area C slides are detailed in table 3.3.

Component General Observations Samples Image Building - building materials that are not in situ TS06-1, TS06-2, TS06-4, fragment - origins can include floors, ceilings, TS06-5, TS06-9, TS06-10, , ovens TS06-11, TS06-13, TS06- - often characterised by densely packed 14, TS06-14, TS06-15, local sediment with sharp, clear edges TS06-16, TS07-1, TS07-8, - common in all contexts TS07-12, TS07-18, TS07- 19, TS07-20, TS07-20B, TS07-30, TS07-37, TS07- 40, TS07-44, TS07-43, TS07-44, TS07-45, TS07- 46, TS07-47, TS07-48, TS07-49, TS07-50

Mud-render - floors are constructed of local calcitic TS06-3, TS06-4, TS06-5, surface clays from the Lisan deposits TS06-10, TS06-11, - contain plant voids from chaff use - mineral and grog inclusions with some examples of re-surfacing - two types on site - fine floors and rough floors - post-depositional gypsum often present in plant voids, usually destructive

63

Pottery - macro-sherds are occasionally visible in TS06-1, TS06-3, TS06-7, samples TS06-8, TS06-14, TS07-9, - smaller fragments (grog) are often used TS07-13, TS07-20, TS07- to temper floors and mudbrick 36, TS07-41, TS07-42, TS07-47, TS07-48, TS07- 49

Plant voids - most often associated with mud-render TS06-3, TS06-5, TS06-5, floor and mudbrick temper TS06-9, TS06-10, TS06- - form when chaff or other organic material 15, TS06-16, TS07-2, used as temper decays, leaving TS07-9, TS07-10, TS07- characteristic hollowed areas 12, TS07-20, TS07-23, - some examples of natural plant voids are TS07-30, TS07-38, TS07- present. In these cases naturally 39, TS07-40, TS07-43, occurring organic material is present in or TS07-44, TS07-45, TS07- disturbs the archaeological sediment. 46, TS07-48, TS07-49, TS07-50

64

Ash -results when organic material is partly or TS06-7, TS06-8, TS06-9, ` fully oxidized during combustion TS06-14, TS07-2, TS07- -most common at higher temperatures 22, TS07-25, TS07-30, (above 500°C) and in high oxygen TS07-38, TS07-43, environments (Smart and Hoffman 1988) - microcrystalline calcium carbonate deposits - often found in association with charcoal, but not exclusively.

Charcoal - results from burning plant material TS06-1, TS06-2, TS06-3, - often only small fragments of ash are TS06-4, TS06-8, TS06-13, visible in samples (a) TS06-14, TS06-16, TS07- - sometime cell structure is visible in 1, TS07-2, TS07-6, TS07- microscopic samples (b) (Woody 8, TS07-9, TS07-18, TS07- Charcoal) 22, TS07-24, TS07-25, - charcoal is most likely to form in low TS07-30, TS07-31, TS07- temperature fires (less than 500 C), 37, TS07-38, TS07-42, reduced oxygen environments or matter TS07-43, TS07-47, TS07- with high water content (Smart and 48, TS07-49, TS07-50 Hoffman 1988)

65

Phytoliths - silica skeletons of plant cells that remain TS06-8, TS06-9, TS06-13, after the organic material has decayed TS06-14, TS06-15, TS07- - soluble silica in the soil is absorbed into 1, TS07-2, TS07-8, TS07- the lining of cell walls 9, TS07-12, TS07-25, - common forms include grasses, reed TS07-31, TS07-36, TS07- stems, bulliforms and wheat/barely and 38, TS-7-39, TS07-40, buliforms -good TS07-41 preservation across the site, easily viewed in samples.

Fecal - calcium carbonate spheres that form TS06-8, TS06-9, TS06-13, spherulites within the digestive tracks of animals - TS06-14, TS07-1, TS07-2, range in size from 5 to 15 µm. TS07-8, TS07-12, TS07- -most visible in cross-polarized light, 24, TS07-25, TS07-30, appearing white or yellow with a cross TS07-31, TS07-38, TS07- through the middle 39 -preserve best in slightly alkaline sediment, with a pH of 7.7 or above - vary in density across contexts, either in clumps (indicating fecal pellets), or dispersed.

66

Chert - most frequent in courtyard and fill TS06-1, TS06-2, TS06-3, contexts TS06-5, TS06-8, TS06-9, -some likely micro-debutage, but there is TS06-10, TS06-11, TS06- not enough variation in density across 13, TS06-14, TS06-15, samples or contexts indicate a workshop TS06-16, TS07-6, TS07- -chips may represent naturally deposited 10, TS07-31, TS07-44, fragments, disposal of waste material or TS07-45, TS07-48 touch-up activity.

Igneous Rock - basalt, gabbro and lava grains common TS06-1, TS06-2, TS06-3, (basalt, in courtyard and fill slides TS06-5, TS06-8, TS06-10, gabbro, lava) - likely originate from alluvial deposits TS06-11, TS06-14, TS06- within the Jordan Valley 15, TS07-6, TS07-10, - it is possible some fragments may be TS07-13, TS07-18, TS07- the result of grinding activities 20B, TS07-31, TS07-36, - none of the samples present sufficient TS07-37, TS07-39, TS07- quantities to make conclusive 44, TS07-45, TS07-47 observations at this point

67

Micro-fossils -microfossils of foraminifera originate in TS06-7, TS06-8, TS06-11, (limestone) limestone deposits TS07-1, TS07-2, TS07-8, - marine origin of Lisan sediment suggests TS07-9, TS07-10, TS07- local origins 13, TS07-18, TS07-22, - most often associated with re-deposited TS07-23, TS07-24, TS07- contexts such as silos and courtyards. 43, TS07-45, TS07-47, TS07-48, TS07-49, TS07- 50

Gypsum - gypsum and anhydrite is post- TS06-1, TS06-3, TS06-4, depositional TS06-5, TS06-8, TS06-9, - often destructive TS06-10, TS07-12, TS07- - most often associated with plant voids in 20, TS07-23, TS07-43, floor levels TS07-44, TS07-45 - individual crystals are visible in these deposits, confirming it was not ground and used as a construction material

Table 3.3: Main components found in the Tel TsafFigure micromorphology 3.4 Main components samples. of the Tel Tsaf, Area C micromorphology samples 68

69

There are three types of human-made components present in the Area C samples; building fragments, mud-rendered surfaces and pottery. Building fragments are the most common constituent and include fragments of mud-brick, mortar, collapsed roof or floor material. Oven fragments are included in building fragments with the acknowledgement they are less closely related to other materials in the category. All Chalcolithic building material was made from local clay and silt deposits. Clear, sharp borders distinguish building fragments from compressed fill material along with increased densities of plant voids.

Macroscopically, mud-rendered surfaces come in two varieties; fine or roughly-lain. In thin- section, the same basic composition is present in floors and mud-bricks. In these cases, size of features, orientation and context were used to distinguish between the two. All are constructed of local clays and silts, with fine floors or mud-brick containing fewer inclusions and more plant voids than roughly constructed floors and mud-brick. In all cases, mud-rendered surfaces at Tel Tsaf exhibited largely in situ horizontal orientation, but some do show evidence of post- depositional disturbance. In these cases, the surface was discontinuous and exhibited isolated changes in orientation.

Some of the micromorphology samples contained pottery sherds, resulting in visible cross sections of the Tel Tsaf ceramics. There are very few examples of fired mud-brick on site, so the well-fired ceramics are easy to distinguish by their red or orange colour and clearly defined borders. Smaller fragments of ceramics (grog) are commonly found as temper in mud-bricks and floor levels, and ceramic fragments in general are common to all fill deposits.

Plant voids occur in anthropogenic features when vegetal stabilizers decay leaving voids within the material (Matthews 2010). The addition of reeds, stems, grasses or seed husks as chaff in floors, mud-brick and mortar is common, and plant voids are visible in all these contexts at Tel Tsaf. The frequency of voids can help establish the proportion of chaff to clay and silt and, thus, provide information on construction techniques.

Charcoal in the Tel Tsaf slides falls into two categories: large pieces of wood charcoal and small fragments or flecks of unidentifiable charcoal. Woody charcoal pieces are large, often visible

70

macroscopically, and frequently reveal the anatomical features of the plant. It is often possible to identify the type of wood based on these features depending on its orientation and how visible the morphology of the charcoal fragment is in a two-dimensional slide (Fahn, et al. 1986; Matthews 2005, 2010; Matthews, et al. 1997). There are a few examples of well-preserved woody charcoal at Tel Tsaf, but it was not possible to identify the samples to species, except to say they likely originated from a shrub (Matthews pers comm). Smaller, less well preserved charcoal fragments are also present in some slides. These examples are much smaller, more widely distributed throughout individual slides and do not present any characteristic cellular features.

Ash deposits are also present in a number of Tel Tsaf samples. Unlike charcoal, ash results when plant matter is mostly or fully combusted, often at high temperatures and in high-oxygen environments (Smart and Hoffman 1988). Most of the ash at Tel Tsaf is calcareous and associated with charcoal fragments, especially in courtyard and destruction contexts. There is also evidence of ash in-fill in void spaces in at least one sample (TS07-2).

A number of micromorphology slides demonstrate good preservation of phytoliths in the sediment. Phytoliths are silica skeletons of plant cells that remain after the organic material has decayed. Soluble silica in the soil is absorbed through plant roots, eventually lining cell walls and interiors (Piperno 2006). Once the organic plant material has decomposed, the silica skeleton can remain preserved in wet or dry conditions in soil pH ranging from 3-9 (Pearsall 2000; Piperno 1988, 2006; Rosen 1992). The unique cell shapes of different plants help archaeologists identify the species represented by phytoliths visible in microscopic analysis, which is especially useful when identifying crop plants such as barley and wheat (Ball, et al. 1993; 1996; Rosen 1992). Studies involving phytolith analysis have also identified harvest and irrigation methods (Harvey and Fuller 2005; Rosen 1999; Rosen and Weiner 1994), as well as dung-derived sediment on sites (Powers-Jones 1994). Studies have also aided in confirming fuel sources as either dung- or plant-based (Gur-Arieh, et al. 2013; Katz, et al. 2007; Lancelotti and Madella 2012; Madella 2003). The most common phytolith form visible in the Tel Tsaf micromorphology samples is from elongated reeds, similar to those found at Çatalhöyük

71

(Matthews 1996; L.-M. Shillito, et al. 2011; 2011), with elongated grasses and buliforms also present.

Fecal or dung spherulites are microscopic calcium carbonate spheres that form within the digestive tracks of animals. Once food is ingested, it begins primary breakdown prior to reaching the stomach or abomasum (in ruminants) where it is immersed in pepsin and hydrochloric acid excreted for proteolysis (pH 3). The resulting mass, known as chyme, then enters the duodenum. Here it is exposed to alkaline secretions from the gut wall, gall bladder and pancreas. This continues as it moves through the large and small intestine, with it reaching a pH of 7 midway through the small intestine. It is believed that the chyme in the abomasum must contain both Ca2+ and Cl- ions as a result of the hydrochloric acid breakdown of plant matter that contains calcium-rich cell walls, calcium oxalates as well as any calcite from the soil that may be ingested. These ions react when neutralized by the sodium bicarbonate-rich secretions in the small intestine, causing the calcium carbonate spherulites to form (Canti 1999).

The spheres are very small, ranging in size from 5 to 15 m. The mineral composition of the spherulites (calcium carbonate) causes them to appear most clearly in cross-polarized light, and they are not easily distinguished in plane-polarized light. In cross-polarized light they appear white or yellow in colour with a cross through the middle. They preserve best in alkaline soils, with a pH of 7.7 or above (M.G. Canti 1997; Canti 1999). They dissolve easily in water, including distilled water, and are best preserved in dry or sheltered areas. Archaeologically, the presence of dung spherulites indicates areas of animal penning or dung use or storage (Albert and Henry 2004; Brochier, et al. 1992; Courty, et al. 1991; Lancelotti and Madella 2012; Macphail, et al. 1997; Matthews 2005; Matthews, et al. 1997; 1994; Shahack-Gross 2011; 2005; Shahack-Gross and Finkelstein 2008; Shahack-Gross, et al. 2003; M. Shillito, et al. 2011).

Bone fragments are present in a few slides, most often associated with unoriented, unconsolidated sediment. Most examples are chips or splinters from larger bones, with no identifiable characteristics. There are a couple of examples of small bones from either a bird or small mammal, which are distinguished by the morphology of the bone fragments. These examples have not been identified to species. Some slides also contain calcined bone. The

72

quantity and distribution of the bone within the Tel Tsaf samples do not show any significant patterns but do not contradict the interpretation of the macro-faunal analysis (Ben-Shlomo, et al. 2009; Hill 2011), which is discussed further in Chapter 4.

Deposits of gypsum and possible anhydrite are frequent in the Tel Tsaf samples. Naturally- occurring gypsum was discovered macro- and microscopically during excavation across all contexts, and is expected given the geology of the region. Within the micromorphology samples, the most common context for gypsum was within mud-rendered floor deposits. The post- depositional formation of gypsum within plant voids is destructive, in some cases breaking apart floor levels completely. It has been suggested that the formation of gypsum in plant voids may be the result of the deposition of authigenic (forming in situ) minerals following plant degradation (Katz, et al. 2007). Observations at Grar in the Negev have led the excavators to propose that gypsum nodules within the archaeological sediment formed in this manner were due to the high content of sulphate salts in the local vegetation (Katz, et al. 2007). It is possible that the gypsum and anhydrite at Tel Tsaf are also the result of degraded local plant material.

Chert fragments are common in a number of contexts on site, but are most frequent in courtyard and fill samples. The presence of microscopic chert flakes may be an indicator that tool retouch took place at the site. Alternatively, the flakes may be the result of tool use within various areas of the site. The third possible explanation may be that tool production took place within courtyards and only microscopic flakes remain after the area was cleaned. With no macroscopic evidence of knapping areas on site, and the limited presence and distribution of microscopic chert, it is difficult to draw any further conclusions.

Volcanic rocks such as basalt and gabbro appear most frequently in courtyard and fill contexts. They appear in thin section most frequently as small rock fragments found within unconsolidated sediment. The presence of volcanic outcrops to the west and north of Tel Tsaf may explain the inclusion of igneous elements within the local sediment. It is also possible that some of the basalt fragments are related to the use of basalt grinding stones on site (Garfinkel, et al. 2007). None of the slides demonstrate significant densities of basalt that would suggest a specific

73

grinding area, but the relative increase of occurrence within courtyards may support this hypothesis.

Microfossils of foraminifera and shell fragments indicate that limestone constitutes a significant portion of the groundmass at Tel Tsaf. Shell may also represent archaeological artifacts or modern land-snails, but given the sedimentary nature of the local sediment and the frequency of microfossils, it is likely that most of the shell eroded from the local limestone.

3.3.4 Microfacies

Every component above is found in multiple samples and contexts across Tel Tsaf, making it unreasonable to draw any conclusions about the site based solely on the distribution of one component. However, patterns emerge when combinations of components, and their distribution and orientation within individual slides are examined. These small-scale changes in slide composition can help determine what human, natural or post-depositional activities contributed to the formation of each sample. In order to define and classify these patterns the concept of microfacies is used following Goldberg, Miller, et al. (2009). The microfacies at Tel Tsaf (table 3.4) are based on common features. Interpretations and locations for each of these are outlined based on distribution across all micromorphology samples analyzed.

Microfacies (MF) 1a and 1b are composed of highly compacted sediment. Microfacies 1a is composed of a compacted fine fraction (clays) containing plant voids with a strong horizontal orientation. Microfacies 1b is composed of fine- and medium-sized particles (clays and silts) and contains fewer plant voids (< 60%) with no strong orientation. Inclusions are larger and include charred plant material, minerals and rock grains. Given the coarse and fine fractions, plant void densities and appearance, both 1a and 1b are interpreted as constructed features, either floors or mud-brick depending on excavation context. Microfacies 1a is present in both well-laid mud-rendered floors and mud-bricks. The interpretation is context-dependent and relies on macroscopic excavation data. Well-laid mud-rendered floors have been identified in Room 70 and in the Byzantine test pit. Microfacies 1b represents roughly laid courtyard surfaces and has not been identified in any room interiors. Samples from the Building Complex I courtyard

74

present the best examples of this microfacies, with fragments also present in the Room 612 samples. Mud-brick construction varies across contexts, falling into similar categories as the floor surfaces – either well-made (MF 1a) or roughly constructed (MF 1b). Recent studies at Çatalhöyük have investigated the social implications of variation in construction of mud-brick, advocating greater investigation of these features (Love 2012). The use of local materials in mud-brick construction results justifies their inclusion with floor surfaces rather than their own microfacies. However, acknowledgment of variation in construction method and components result in two categories of mud-brick – MF 1a and MF 1b.

Microfacies 2 is characterized by horizontally laminated organic matter, with charcoal and phytoliths being the dominant components. Long-celled phytoliths within a silty matrix and charcoal fragments are found in relation to microfacies 1b (courtyard floors) in Building Complex I. These features are interpreted as surfaces resulting from the laying down of organic material on the courtyard floor, combined with probable rake-out from roasting pits or hearths.

Microfacies 3a and 3b consist of phytoliths, spherulites and consolidated dung. In microfacies 3a, these features are horizontally laminated and compressed in situ. In contrast, the features are unconsolidated, with no clear orientation in microfacies 3b. Both cases are interpreted as related to animal-penning activities, similar to those at Çatalhöyük (Matthews 2005) and Tel Dor (Albert, et al. 2008; Shahack-Gross, et al. 2005). The horizontal laminate of microfacies 3a most likely represents in situ penning activities. The combination of enlongated phytoliths from reeds embedded in clustered fecal spherulites is very similar to natural dung deposits at the sites mentioned above. If dung was re-worked for fuel use and stored, one would expect a more uniform distribution of fecal spherulites, as well as likely plant-voids from chaff that would decay within the dung-cake matrix. Microfacies 3b contains the same components as 3a; however, they are unoriented and randomly distributed. The samples contain elongated reed phytoliths and clusters of fecal spherulites, but the two are unrelated within the slide matrix. The samples appear to be either re-deposited or highly disturbed penning material.

Microfacies Example Interpretation Locations MF1a building floor, Building compacted fine fraction mudbrick Complex I a) plant voids + horizontal orientation

MF1b courtyard floor, Building compacted medium mudbrick Complex I, fraction Room 612 a) plant voids b) organic inclusions c) mineral inclusions -anthropogenic inclusions

75

MF2 courtyard surface, Building laminated organic matter on top of floor Complex I a) charcoal b) phytoliths + general horizontal orientation with some bioturbation/disturbance

MF3a animal pen Room 230 laminated organic matter a) phytoliths b) fecal spherulites + amorphous organic matter + horizontally oriented

76

MF3b disturbed animal Room 263 non-laminated organic pen or re- matter deposited pen a) phytoliths material b) fecal spherulites + amorphous organic matter + un-orientation

MF4 oven? south of natural and anthropogenic Building material Complex IV a) phytoliths b) charred bone c) charcoal d) ash + clay +amorphous organic matter + scattered fecal spherulites + un-oriented

77

MF5 destruction Room 612, natural and anthropogenic material a) charcoal b) amorphous dung c) building fragments + ash + phytoliths + fecal spherulites + un-oriented

MF6 fill Silo 272, Silo un-oriented natural and 336, Room anthropogenic matrix 70, Building a) geological inclusions Complex II b) building material courtyard fragments + anthropogenic inclusions + unoriented

Figure 3.5 Microfacies present in Tel Tsaf, Area C micromorphology samples.

Table 3.4: Microfacies found in the Tel Tsaf micromorphology samples. 78

79

Microfacies 4 is characterized by unconsolidated charred bone, wood and clay within amorphous organic matter, phytoliths, fecal spherulites and ash. It is only present in one context, found in the southern courtyard of Building Complex IV. The feature sampled appears as a discrete line of ash with definitive edges. It has been interpreted as an oven or potential hearth feature. The presence of burnt bone and wood supports this, and the phytoliths and fecal spherulites indicate dung was also used as a fuel source.

Microfacies 5 contains an unoriented matrix of building fragments, charcoal, ash, dung, phytoliths and fecal spherulites. This appears to represent a conflagration level that caused extensive damage to the site. All of the samples with this microfacies come from the burn level within the ‘wall’ of Room 612.

Microfacies 6 consist of unoriented building material fragments as well as geological and anthropogenic inclusions. Geological inclusions include basalt, chert flakes, microfossils, shell, gypsum/anhydrite and quartz. The most common anthropogenic inclusions are pottery fragments. The unorientated, unconsolidated nature of the material suggests it is a fill level, re-deposited either intentionally or naturally. This is the most widespread microfacies at Tel Tsaf, and is often present in conjunction with other microfacies.

Applying the described microfacies to the analysis of the micromorphology samples from Tel Tsaf provides specific categories for description and understanding of activities across the site. Similarities and differences of slide features are easier to classify and more consistently interpreted when such a structure is used. It also allows easier comparison of activity areas, and provides a more precise framework for comparison to other data sets, such as bulk sample analysis. The remainder of the discussion of the micromorphology results will refer to the microfacies outlined above.

3.3.5 Building Complex I – Phases 3 and 4

The architecture of Building Complex I remained relatively consistent between Phase 3 and Phase 4 at Tel Tsaf. It is, therefore, difficult to distinguish at what level each phase begins or ends. In addition, the evidence from this building suggests that its use was consistent

80 throughout its entire life history. For this reason, the evidence from Building I will be addressed as one unit, referring to its use through Phase 4 into Phase 3 (figure 3.8).

Figure 3.8: Building Complex I, Phases 4 and 3. The micromorphology samples marked are those that were processed for analysis and are discussed below. 3.3.5.1 Room 70 – Rectangular Room

MF1a MF1b MF2 MF3a MF3b MF4 MF5 MF6 TS06-1 TS06-2 TS06-3 TS06-4

TS06-5

Table 3.5: An overview of micromorphology samples from Building Complex I, Room 70. The black squares represent the microfacies present in each sample slide.

81

All samples from Room 70 contain evidence of fill levels (MF 6) dominated by unconsolidated silty-sand. Among the inclusions are small fragments of ceramics, igneous aggregates and microfossils. Three samples exhibit well-laid mud-rendered floor surfaces (MF 1a). These well- sorted, fine-grained renders contain horizontally-oriented plant voids. All three surfaces are highly disturbed by post-depositional gypsum.

Excavation of the lower portion of Room 70 during the 2005 season at Tel Tsaf left a baulk exposed in the southern area of the structure (figure 3.9). The baulk was 37.4 cm high, with at least four plaster levels visible. The thickest floor level was 3 cm thick and located 11.8 cm from the top of the baulk. The section also showed burnt clay areas in the lower portion of the baulk, associated with visible ash lenses. Pottery and bone fragments were also clearly visible, most densely clustered in the eastern half of the baulk. All five samples from the Room 70 surface were taken from the western half of the section.

Figure 3.9: Building Complex I, Room 70, baulk highlighted by white lines. The area from which micromorphology samples were extracted is highlighted by the red circle.

Samples TS06-1 and TS06-2 are dominated by MF 6 and represent the ‘fill’ levels between the visible floors. The micromorphology slides contain fragments of building material, generally in the form of mud-brick fragments that make up 40-60% of each slide. Both slides also contain significant numbers of flint fragments (max. diameter 2 mm) and basalt fragments. Bone fragments are also present, but rare. There is strong evidence of post-depositional disturbance in

82 the form of gypsum (and possibly anhydrite) crystals, especially in TS06-2. The crystals are intact and appear as naturally-occurring, post-depositional features, rather than as crushed gypsum plaster. The evidence from the two slides suggests that the deposits between the floors are fills to level the surface in preparation for a new floor layer.

The largest visible mud-rendered floor level (3 cm) is represented in TS06-3. The upper portion of the slide contains loose fill, with fragments of building material, including mud-brick (MF 6). The fill material is unoriented and loosely packed. The lower portion of the slide is represented by the mud-rendered floor (MF 1a), indicated by fine, tightly packed sediment and oblong plant voids (figure 3.10).

Figure 3.10: Room 70, Building Complex I. Left: fill level above floor surface (MF6). Circled area highlights the unoriented, loosely packed fill material. Right: fine, mud-rendered floor surface (MF 1a). Arrows highlight plant voids within the tightly packed floor material (TS06-3, PPL, 50X).

Observation reveals that the floor is constructed of mud and clay, rather than lime or gypsum plaster. Horizontally oriented voids indicate that the mud-rendered floor was tempered with plant fragments; likely reed stems. There is also evidence of pottery being used to temper the floor material. Examination of the upper limit of this level produced evidence of a new application of mud-rendering at least once during the use-life of the floor. The upper, finer layer of mud-render does not contain plant voids, and the boundary between the upper and lower floor indicates in one location that there may have been a reed mat on the larger, lower portion prior to the re-plastering (W. Matthews pers comm) (figure 3.11).

83

Figure 3.11: Room 70, Building Complex I. The original floor level is topped by a fine layer of new mud-rendered. The white arrows highlight the undulating boundary between the two surfaces, which is likely the result of an impression made by a woven reed mat on the original surface (TS06-3, PPL, 50X).

As with the fill slides, the floor samples show extensive post-depositional disturbance caused by post-depositional gypsum. It is clear that the gypsum was not part of the building material, as it tends to form within the plant voids, and displays classic sharp edges rather than a homogenous granular appearance indicative of ground gypsum used as plaster. Samples TS06-04 and TS06- 05 both contain evidence of floor levels; however, both have suffered disturbance from the gypsum. In both cases, there are large fragments of the floors that have been shifted or decayed by the intrusive mineral. In both samples, the floor levels themselves are approximately 1 cm thick, significantly thinner than in TS06-3. The fill levels for these slides contain the elements expected based on the other slides from the section. There is an assortment of building material, flint, bone, basalt and shell, with mud-brick fragments representing the majority of the material.

84

3.3.5.2 Courtyard

MF1a MF1b MF2 MF3a MF3b MF4 MF5 MF6 TS06-7 TS06-8 TS06-9 TS07-6 TS07-30

TS07-31

Table 3.6: An overview of micromorphology samples from the Building Complex I Courtyard. The black squares represent the microfacies present in each sample slide.

Four of the Building Complex I Courtyard samples (TS06-8, TS06-9, TS07-6, TS07-30) contain evidence of courtyard floors (MF 1b) or surfaces (MF 2). These contain mostly laminated organic material dominated by charcoal and phytoliths, although there is greater bioturbation in slide TS07-30. The MF 2 components represent likely rake-out or redistribution of hearths or ovens combined with chaff spread throughout the courtyard. Three samples contain evidence of rough-laid courtyard floors, onto which the MF2 components were deposited. Fill material (MF 6) is common in most of these slides, and two slides (TS06-7 and TS07-6) contain only fill material, and the earliest deposits contain evidence of some water-lain sediment.

The courtyard was sampled in three separate locations over the course of the two years of data collection. The first samples, taken during the 2006 season, were taken from a baulk created by the extraction of a Byzantine burial cut through the Chalcolithic levels of the site. This area was located in the eastern portion of the courtyard, oriented NE-SW, with the NE portion of the excavation baulk cut into the wall dividing Building I from Building II. The second set of samples was taken during the 2007 season, from beside the western wall of Room C70, where a baulk was left during excavation. The final samples were taken from the southern wall of a second Byzantine burial pit straddling Building Complex I and Building Complex II.

Samples TS06-7, TS06-08 and TS06-09 represent the first samples taken from the courtyard. The Byzantine burial pit measured 1.58 m in length and ranged in width from 1.31 to 1.39 m

85 with a depth of 63 cm. The samples were taken from the southern wall, which exhibited the best preserved stratigraphy.

TS06-7 was dominated by fill, but the other two samples have clear evidence of burning, with ashy levels and macro-fragments of charred wood (MF 2). However, these levels are bounded by layers of both loose and compressed fill. There appear to be alternating levels of compressed floors (MF 1b) and organic-rich, burnt levels (figure 3.12). The floor levels consist of tightly- packed, sandy silt and clay, but are not as clearly defined as the mud-rendered floors in C70. Both slides have evidence of flint fragments, calcinated bone, pottery and building debris. The organic/ash levels consist of loose sand, silt and clay combined with ash and coprolitic material and are characterized by a high degree of organic staining. There is also evidence of horizontally-bedded reed stem phytoliths associated with the organic layers (figure 3.13). The upper portion of TS06-08, containing a mixture of ash, loose sediment and charred wood, and likely indicates either a street or courtyard area, with fresh organic material scattered.

Figure 3.12: Building Complex I, Courtyard. Left: Rough laid, courtyard floor surface (MF 1b). Arrows indicate plant voids within the tightly-packed floor material. Right: Organic-rich, burnt courtyard surface debris (MF 2). Arrows highlight charcoal and circles indicate areas with organic staining (TS06-8, PPL, 50X).

86

Figure 3.13: Building Complex I, Courtyard. The white highlights laminated phytoliths in combination with charcoal (red arrows) and other fill material (circled) (MF 2) (TS06-8, PPL, 50X).

The courtyard samples from the 2007 season (TS07-30 and TS07-31) were less clearly defined in the micromorphology samples. The upper portion of the section represented the lower limit of the 2006 excavations, with the lower portion defined by the end of the 2007 excavations. The final baulk measured 31 cm in depth. Visual examination of the section indicated alternating levels of compacted and loose material. Pebbles and rocks are visible in the section at varying levels with no apparent pattern. There also appears to be a lens of fine charcoal in the lower portion of the baulk. Five micromorphology samples were taken from the section but only two samples survived the sampling process and both of these samples are from the lowest portions of the baulk.

The micromorphology slides present dense, well-sorted marl and clay, likely water-lain. TS07- 31 clearly shows banded sands, well-sorted by water deposition (figure 3.14). Partially burned coprolitic material and organic staining are prevalent in both samples. These slides were taken

87 from a lower level than the other two courtyard slides. It is possible that excavations here reached below the levels of Phase 3 and these slides represent the pre-occupation levels of Building I. The evidence of phytoliths as well as dung spherulites from TS07-30 and TS07-31 may also indicate that the previous occupation was briefly abandoned prior to Phase 3.

Figure 3.14: Building Complex I, Courtyard. Well sorted, water-lain deposit, including a) larger grains b) fine fraction c) marls indicating this was likely not disturbed by the Chalcolithic inhabitants of Tel Tsaf (TS07-31, PPL, 50X).

The final sample was taken once a portion of the courtyard section was revealed during the excavation of a Byzantine burial pit. The baulk was located between silos 339 and 415 in the northeastern portion of the courtyard. TS07-6 contains large fragments of rough-laid courtyard floors (MF 2), in combination with phytoliths, partially ashed dung and clusters of charcoal. It appears to be highly disturbed, courtyard surface material combined with broken floor material and general fill.

88

3.3.5.3 Room 70 – Wall samples

MF1a MF1b MF2 MF3a MF3b MF4 MF5 MF6 TS07-20 TS07-33

Table 3.7: An overview of micromorphology samples from Building Complex I wall surfaces. The black squares represent the microfacies present in each sample slide.

Two samples were taken from wall surfaces within Building I in order to examine its construction material. As expected, both samples contain fine mud-rendered material and plant voids (MF1a). The TS07-20 material represents a mud-brick, while the MF 1a material in TS07- 33 appears to be a surface render on the wall. In each case, the material contains mostly fine fraction, plant voids and few inclusions.

TS07-20 was taken from the northern face of the southern wall of Room 70. It appeared to be a plaster surface, varying in thickness and discontinuously preserved along the exposed wall. The sample came from just above the 2006 floor level. The second sample, TS07-33 was taken from the southeast wall of the courtyard. The sample comes from between two courses of mud-brick on the wall and appeared rippled in the field (figure 3.15).

Figure 3.15: Building Complex I, Room 70. Material sampled from the north face of the southern wall.

89

It was assumed that TS07-33 was a sample of mortar used during the construction of the wall. The upper extreme of the sample, representing the visible fabric in the field, contained fine sediment consistent in colour with other construction materials on the site (MF 1a). However, immediately beneath that layer are two distinctly reddish layers (figure 3.16). This suggests that the material was heated or fired, creating the two grades of colouring. It is most likely that the visible layer in the field was a thin layer of mortar or mud-rendering spread over a baked mud- brick.

Figure 3.16: Building Complex I, Room 70. Flatbed scan of micromorphology slide of a fired mud-brick. The arrow indicates the first layer of brighter red material, with the core of the brick immediately beneath (TS07-33, original slide size 5 x 7.5 cm).

Slide TS07-20 contains a thin, distinct mud-rendered layer between a fill level and potential mud-brick (figure 3.17). The rendered level contains a large number of plant voids, indicating

90 the liberal use of vegetal matter as temper (MF 1a). The voids are not as horizontally oriented as the mud rendered floor from the same building. There is no evidence of multiple surfacing events in this sample.

Figure 3.17: Building Complex I, Room 70. Flatbed scan of a wall sample with mud rendered surface between a) fill and b) mud-brick layers. (TS07-20, original slide size 5 X7.5 cm)

3.3.5.4 Silo 339 – Chalcolithic Burial Pit

MF1a MF1b MF2 MF3a MF3b MF4 MF5 MF6 TS07-34 TS07-36

TS07-37

Table 3.8: An overview of micromorphology samples from the interior of Silo 339. The black squares represent the microfacies present in each sample slide.

The silo samples are dominated by roughly-made mud-brick fragments (MF 1b), which is to be expected as the interment appears to be dug through the silo construction. Unlike the mud-brick

91 from Room 70’s wall (TS07-33), the silo bricks contain few, if any plant voids and often contain fragments of unworked, water-laid clay. Two samples also contain fill material, likely packed around the bricks during construction of the silo.

After the excavation of the human interment from Silo 339, the side of the burial pit within the silo was sampled (figure 3.18). One sample (TS07-34) was taken from the eastern end of the northern wall of the pit, and the remaining two were taken from the southern wall. The exposed sections did not contain any obvious stratigraphy and the burial pit was not lined, so it was unclear whether the exposed wall belonged to the silo or whether it was remaining fill from the burial. There were visible ash and charred seeds within the edges of the burial pit.

Figure 3.18: Building Complex I, Silo 339. Northern wall of the Chalcolithic burial pit. Sample TS07-34 was extracted from this surface (indicated by white box).

The sample from the northern wall (TS07-34) was taken from directly above a mud-brick visible in the section. The slide contains well-sorted silt, clay and marl, as well as fragments of unworked clay (figure 3.19). The layering suggests that the sediment was water-lain. Sample TS07-36 also shows well-sorted clay and marl, but no unworked clay fragments. The final sample overlapped TS07-36, and the upper portion presents similar features. However, the lower

92 portion of the slide is less consolidated and appears to contain fragments of building material and pottery (MF 6).

Figure 3.19: Building Complex I, Silo 339. An example of unworked clay incorporated into the mud-brick used in construction of the silo. (TS07- 34, PPL, 100X)

The samples indicate that the exposed burial walls were unlikely to represent the original pit walls. The presence of fine clay and marl, in addition to fragments of unworked clay suggests that constructed or fill levels within the silo were made with natural, unworked sediment from the river.

3.3.6 Building Complex II - Phase 3

As previously discussed, Building Complex II is only present during Phase 3 and the area is renamed in Phase 4a as Building Complex IV. A map showing the sampling locations for the Phase 3 occupation appears in figure 3.20.

93

Figure 3.20: Building Complex II, Phase 3. The micromorphology samples marked are those that were processed for analysis and are discussed below. 3.3.6.1 Room 230 – Circular Structure

MF1a MF1b MF2 MF3a MF3b MF4 MF5 MF6 TS06-13 TS06-14 TS06-15 TS06-16

TS06-17

Table 3.9: An overview of micromorphology samples from Building Complex II, Room 230. The black squares represent the microfacies present in each sample slide.

94

The samples from Room 230 fall into 2 categories: animal penning material (MF 3a, MF 3b) and fill material (MF6). The 2 samples dominated by evidence for penning were taken from the interior of Room 230 and showed evidence of layering in the baulk cross-section. In contrast, the three remaining samples were taken from the same baulk, but from a portion bisecting a post- occupational pit (figure 3.21).

During the 2006 season, a baulk was created across the southern portion of Room 230 during excavations. This section bisected a large pit in the floor of the room (Feature 346) and measured approximately 20 cm in height from the floor level (figure 3.21).

Figure 3.21: Room 230, Building Complex II. The baulk (highlighted in white) transects the southern portion of the interior of the building, bisecting a later pit. The areas sampled are indicated by red circles.

The baulk’s eastern portion appeared to show at least two distinct, white layers that were originally identified during excavation as ‘living floors’. These layers, in an orangey-brown matrix, where interpreted as floor and fill levels at the macroscopic level in the field. These white layers were not continuous across the section and sloped slightly to the west towards the

95 pit (figure 3.22). They appeared to correspond to a number of light grey, phytolith-rich layers discovered during excavation of the room. The baulk appeared to show evidence of four compact mud horizons of the same colour as, but differing density than, the intervening fill levels. The sloping of the eastern stratification of white and orangey-brown layers towards the pit suggests it was dug through these occupation layers at a later date, rather than being contemporaneous with the lower levels.

Figure 3.22: Building Complex II, Building 230. Idealized section drawing reconstructed from field notes (drawn by T. Woolcott).

Two micromorphology slides, TS06-13 and TS06-14, were taken from the eastern portion of the baulk. Slide TS06-14 revealed layers of reed stems embedded in clusters of dung spherulites (MF 3a). Calcareous dung spherulites form on the intestinal tracks of animals and an association with intensive animal activity has been established at archaeological sites (Brochier, et al. 1992; M.G. Canti 1997; 1998; Courty, et al. 1991; Matthews, et al. 1994). Clusters of these spherulites were evident within laminate layers of reed stem phytoliths (figure 3.23). Slide TS06-13 contained some sections of horizontally oriented material (MF 3a), but also has some material that was more highly disturbed (MF 3b).

There is no microscopic evidence of constructed floor levels, contrary to original field observations. A similar situation arose during excavations at Tel Dor where the original ‘floor’ levels were instead compressed layers of phytoliths and dung (Albert, et al. 2008; Shahack- Gross, et al. 2005). At Tel Tsaf, the microscopic material is in situ and vertically compressed

96 with limited post-depositional disturbance and strongly suggests this area was occupied by animals for a continuous period.

Figure 3.23: Building Complex II, Building 230. Laminate phytolith and fecal spherulite matrix (MF 3a). Left: PPL. Right: XPL. The arrows indicate the phytoliths, which appear as voids in XPL. The fecal spherulites are not visible in PPL, but appear in abundance in XPL (a cluster is circled to demonstrate to the reader how fecal spherulites appear in this image)(TS06-14, 100X).

There is no microscopic evidence of constructed floor levels, contrary to original field observations. A similar situation arose during excavations at Tel Dor where the original ‘floor’ levels were instead compressed layers of phytoliths and dung (Albert, et al. 2008; Shahack- Gross, et al. 2005). At Tel Tsaf, the microscopic material is in situ and vertically compressed with limited post-depositional disturbance and strongly suggests this area was occupied by animals for a continuous period.

The micromorphology slides from the samples from the Room 230 pit (TS06-15 and TS06-16 and TS06-17) contained unconsolidated fill of sand and silt interspersed with organic material and charcoal (MF6). The compact mud levels seen in the baulk were not well-defined in the micromorphology slides. Indeed, slide TS06-17, taken from the pit beneath the floor level, demonstrates a similar make up to TS06-15 and TS06-16. All three slides contained fragments of mud-brick, as well as bone, shell, basalt and flint fragments. The consistency in the groundmass and inclusions in all three slides combined with the lack of any correlated orientation for this material confirms that the pit was dug through the layers sampled from the eastern portion of the room.

97

3.3.6.2 Room 268 – Circular Structure

MF1a MF1b MF2 MF3a MF3b MF4 MF5 MF6 TS07-1 TS07-2

Table 3.10: An overview of micromorphology samples from Building Complex II, Room 268. The black squares represent the microfacies present in each sample slide.

The two samples from Room 268, both extracted from an interior baulk left during the 2007 season (figure 3.24), exclusively contain disturbed or re-deposited animal penning material (MF 3b). TS07-02 contains a small section of laminate, horizontally oriented material, but the rest of the slide matrix is too disturbed to determine whether the orientation represents in situ activity or coincidental orientation. The phytoliths in these slides are much smaller and fragmented than in the Room 230 samples and there are no other building fragments or large inclusions in the Room 268 slides. There was no clear stratigraphy evident at the macroscopic scale, and the slides showed greater bioturbation and re-deposition of its constituents.

Figure 3.24: Room 268, Building II. A small baulk was left along the southeast wall at the beginning of the 2007 season. The upper limit of the section represents the end excavations from 2006. The red circle indicates where the micromorphology samples were extracted.

98

The slides did however provide further evidence of re-deposited animal dung through the presence of dense concentrations of dung spherulites and some scattered reed phytoliths (figure 3.25), in addition to greyish ash with high birefringence, which suggests melted coprolitic material (MF 3b). Room 263 was likely either intentionally filled with debris that included dung and reeds, or it was used as an animal pen and suffered from greater post-depositional disturbance than Room 230. While the evidence is not as conclusive as for Room 230, the slides from Room 263 do show some indicators of likely animal penning.

Figure 3.25: Building Complex II, Room 268. The same field of view is presented in both images (MF 3b). On the left (PPL) arrows highlight some examples of fragments of phytoliths within the amorphous organic matrix. The arrows in the right image (XPL) highlight examples of fecal spherulites for the reader (TS07-2, 100X). 3.3.6.3 Silo 272

MF1a MF1b MF2 MF3a MF3b MF4 MF5 MF6 TS07-43 TS07-44 TS07-45 TS07-46 TS07-47 TS07-48 TS07-49 TS07-50

Table 3.11: An overview of micromorphology samples from Building Complex II/IV, Silo 272. The black squares represent the microfacies present in each sample slide.

99

Only two microfacies are present within Silo 272 samples. The slides contained fine mud-brick (MF 1) or fill (MF 6) or both. The mud-brick was well-constructed with some plant voids and no evidence of unworked clay material.

During the 2007 excavation season, a series of micromorphology samples were taken from Silo 272 to test the hypothesis that the paved circular features were grain storage silos. During excavation, it became apparent that Silo 272 was present during Phase 3 and Phase 4; however, it underwent changes in size from the earlier to the later phase, with the earliest (Phase 4) being the largest. The entire preserved silo was approximately 70 cm high, with three stepped levels. In order to address the question of its purpose, half the structure was excavated, revealing a section midway through the structure (figure 3.26).

Figure 3.26: Building Complex II and IV, Silo 272. The silo was sectioned exposing alternating layers of hard mud- brick and softer fill material (indicated with arrows).

During excavations of the eastern half of the structure, two compact floor levels were identified. These surfaces consisted of hard, packed mud and corresponded to visible packed mud in the section. There was loose fill between these floors that produced only a few pottery sherds. However, there were areas directly above the floors that produced 'sheets' of white phytoliths. One section of the floor clearly has phytoliths embedded in its surface (figure 3.27). This evidence supports the idea that the structures were used for grain storage. Alternatively, the silos

100 could have been lined with reed mats. It was hoped the micromorphological analysis could provide greater information regarding the relationship between the fill and floor levels.

Figure 3.27: Building Complex II and IV, Silo 272. A piece of hard-packed floor material from the silo has visible white phytoliths embedded in it, as indicated by the white arrows.

A number of difficulties were encountered during sampling of the structure. There were visible distinctions between compact mud and loose fill in the section and samples were designed to bridge the boundaries between these elements. However, the compact mud proved difficult to cut through. At the same time the loose fill crumbled very easily. Despite these difficulties, seven blocks (approximately 10 cm each) were taken for analysis (figure 3.28).

101

Figure 3.28: Building Complex II and IV, Silo 272. The locations of the micromorphology samples are circled.

Analysis of the slides created from these samples was not as illuminating as hoped. The compressed mud floor surfaces are constructed from the same material as the fill and as a result the slides were generally homogeneous in terms of material present; however, the density of components varied. There appears to be a mixture of mud-brick fragments (MF1a) and loose fill (MF 6). There were a few fragments of shell and bone and occasional burnt fragments of wood; however, these were limited and not associated with specific features or elements. Overall, the slides are more indicative of general fill, with mud-brick fragments, loose sediment and other items mixing with no apparent structural relationship.

3.3.6.4 Byzantine Test Pit (North East corner of excavation area)

MF1a MF1b MF2 MF3a MF3b MF4 MF5 MF6 TS06-10

TS06-11

Table 3.12: An overview of micromorphology samples from the Byzantine Surface, north of Building Complex II. The black squares represent the microfacies present in each sample slide.

The two samples were taken from a shallow test pit dug through the uppermost Byzantine levels in UU16B (figure 3.29). The test pit was 23 cm deep with dimensions of 57 cm by 61 cm. There were two mud rendered floor levels visible, interspersed with reddish soil. The floor levels, while more refined than those from the Chalcolithic levels, are primarily calcareous with

102 limited mineral inclusions (figure 3.30). There are clear boundaries between the floor and fill levels, but the floor levels appear to be disturbed by post-depositional forces. The fill from these sections largely consists of building fragments and patches of gypsum (figure 3.30).

Figure 3.29: Byzantine Surface, located north of Building Complex II (circled). The test pit is indicated by the square. Micromorphology samples were taken from the northwest wall of the pit.

Figure 3.30: Byzantine Surface. The floor level (Left) consisted of very fine fraction, densely packed with few plant voids (MF 1a). The fill levels (Right) are consistent with similar fill on site (MF 6) (TS06-10, 50X).

103

3.3.7 Building Complex IV – Phase 4

The lower levels of Building Complex II revealed a large ‘spill’ of mud-brick along the eastern limits. Within this ‘spill’, a circular room (662) and a rectangular room (612) were excavated. The exact extent and configuration of these rooms is difficult to distinguish as they were not constructed of mud-brick, but rather dug through a mud-brick spill (figure 3.31).

Figure 3.31: Building IV, Phase 4. Circular room 662 and Rectangular room 612 (both circled) were cut through the destructive mud-brick spill. 3.3.7.1 Room 662

MF1a MF1b MF2 MF3a MF3b MF4 MF5 MF6 TS07-18 TS07-19 TS07-20B TS07-22

TS07-23

Table 3.13: An overview of micromorphology samples from Building Complex IV, Room 662. The black squares represent the microfacies present in each sample slide.

All the samples from Room 662 are dominated by fill material (MF 6). Samples TS07-22 and TS07-23 were taken from a baulk left near the centre of the room and TS07-23 presents the only evidence of a floor surface, similar to other courtyard floors (MF 1b). Small concentrations of

104 amorphous organic material, ash and charcoal in sample TS07-22 suggest the presence of courtyard surface material (MF 2) similar to Building I.

Macroscopic views of the walls from Room 662 did not reveal any visible stratigraphy, but three samples were taken for analysis (TS07-18, TS07-19 and TS07-20B). These samples consisted of consolidated sections of clay and marl within a matrix of loose, unconsolidated material (MF 6). The consolidated fragments were not linear and they were not consistent in orientation. Inclusions within the consolidated and unconsolidated sections included shell, bone and mineral aggregates. There is no evidence to demonstrate this was a living area; however, the samples were taken from the walls rather than the floor area. Overall, these samples were not able to shed any light on the extent or function of this area and indicate that the mud-brick spill is immediately on top of an unoccupied or intentionally filled level. These samples were taken from a wall that did not show visible burning so the relationship between the sediment and the destruction phase remains unclear.

A small section was cordoned off and left unexcavated to create a small baulk in the northern portion of the room after the first three samples were taken. Samples TS07-22 and TS07-23 were taken from this section. Both samples contained dense material and TS07-23 presented a surface similar to the courtyard floor in Building I (MF 1b). The material appears to be a compressed floor surface with evidence of plant voids (figure 3.32). This is only visible on the upper portion of the slide, with the remainder being unconsolidated (MF 6). This suggests that there was at least one floor surface within Room 662; however, it appears to be a compacted earthen floor, perhaps outdoors, rather than an interior mud-rendered surface as seen in Room 70.

105

Figure 3.32: Building Complex IV, Room 662. Example of a courtyard floor fragment with some characteristic features indicated (MF 1b). a) charcoal b) mineral aggregate c) plant void (TS07-23, PPL, 50X). 3.3.7.2 Room 612

MF1a MF1b MF2 MF3a MF3b MF4 MF5 MF6 TS07-38 TS07-39 TS07-40 TS07-41 TS07-42

Table 3.14: An overview of micromorphology samples from Building Complex IV, Room 612. The black squares represent the microfacies present in each sample slide.

106

The samples from Room 612 are dominated by evidence for destruction and fill layers consistent with the very clear burning visible within the wall. However, TS07-39 and TS40 also present evidence of rough-lain, mud-rendered floors (MF 1b). These do not appear to be in situ, but floor fragments are not found in general fill levels and are most often associated with living areas. Given their association with destruction (MF 5) and fill (MF 6) levels, it can be assumed that the surfaces were damaged during the conflagration.

Room 612 contains evidence of a destruction layer in the form of a thick black burn level visible along all three walls. The level runs under the mud-brick spill into which Room 612 is cut. Samples were taken from both the burn levels and the wall.

Two samples were taken from the western wall of the room (TS07-38, TS07-39). TS07-38 was taken through the ash layer, with the lower extreme of the sample beginning at the floor level (figure 3.33). The sample largely contains unconsolidated material with abundant ash and charcoal. In addition there are pockets of dense phytoliths and spherulites indicating burnt or melted coprolite fragments within the section (MF 5). The second sample (TS07-39) was taken from beneath the burn level, with the upper portion incorporating the very lowest level of ash. This slide presents evidence of compressed floor levels as found in Room 662 and the courtyard in Building I (MF 1b). These levels are not clearly defined and appear disturbed. There is also evidence of long, compressed coprolites as indicated by dense phytoliths and spherulites (figure 3.34). These are most abundant in the upper portion for the slide, near the burn level. This indicates that there was an occupation surface beneath the destruction level; however the current excavated area does not represent the full extent of the occupied area.

107

Figure 3.33: Building Complex IV, Room 612. Wall area where sample TS07-39 was extracted: A) ash and charcoal b) ash c) mud-rendered surface fragment.

Figure 3.34: Building Complex IV, Room 612. Amorphous organic material with ash infill (best seen in PPL on Left). Circles indicate areas of concentrated fecal spherulites, visible in XPL (Right) (TS07-39, 100X).

108

The three remaining samples from Room 612 were taken from the northern wall. All three were extracted from above the burnt level. All three slides produce evidence of fragments of building material, indicated by large, sharp-edged pieces surrounded by loose fill material (MF 6)(figure 3.35). TS07-40 contains fragments that look similar to those in the courtyard floor levels (MF 1b) but they are not oriented or consolidated. There are also sections of dense organic-rich fill within all three slides. These slides indicated there was likely some form of occupation between the destructive burn level and the mud-brick spill. It is possible the event causing the mud-brick to tumble also led to the disturbance of the occupation levels.

Figure 3.35: Building Complex IV, Room 612. Left: fragment of courtyard floor level with arrows indicating plant voids within tightly-packed fine sediment (MF 1b). Right: Fill (MF 6) within which floor fragments were located. The sediment is more coarse and more loosely-packed than the floor level (TS07-39, PPL, 50X).

The presence of floor material beneath the burnt level suggests that there was a living surface that extended beyond the excavated limits of Room 612 prior to the destruction level. The samples from above the burn level suggest that there may have been additional post-destruction floor levels; however, the existing evidence makes it difficult to speculate on their size or extent. It is apparent that the final excavation configuration for Building IV Phase 4 does not fully represent the extent and layout of the site during this phase. Furthermore, there is evidence that the area was extensively used both before and after the two destruction phases seen during excavation.

109

3.3.7.3 Southern Courtyard

MF1a MF1b MF2 MF3a MF3b MF4 MF5 MF6 TS07-08 TS07-09 TS07-10 TS07-12 TS07-13

Table 3.15: An overview of micromorphology samples from the Courtyard south of Building Complex IV. The black squares represent the microfacies present in each sample slide

The samples from the courtyard south of Building Complex IV contain a variety of microfacies, but do not provide a clear indication of how the space was used or whether different activities were performed in different areas. The Building Complex IV Phase 4 courtyard samples were taken from the north wall of the pit dug through Phase 3 Room 230. The mud-brick structure, Room 230, did not extend below the 2006 excavation levels, suggesting that the pit samples represent the courtyard of the earlier Phase 4 occupation. Both examples of MF 1b are mud- brick but they were not part of specific structure. The remaining three samples contain unconsolidated material similar to fill (MF 6) elsewhere on site. They also contain clusters of fecal spherulites and phytoliths reminiscent of MF 3b material from Room 268. It is possible this material was re-deposited at lower layers during the digging of the pit the samples were taken from.

The samples do not present any evidence of floor levels. The material is generally loose and unconsolidated, with the exceptions of TS07-10 and TS07-13 (MF 6). These slides are very dense and well-consolidated and may represent mud-brick fragments (MF 1a) (figure 3.36). They do not present evidence of plant voids from reed-stem temper as seen in other mud- rendered surfaces. In addition, they do not demonstrate the same characteristics as the compacted courtyard floors. The material is fine and densely packed with few inclusions. The material is very similar to other mud-brick fragments from other contexts such as Silos 339 and 272.

110

Figure 3.36: Building Complex IV, Southern Courtyard. Possible mud-brick, with indicators such as densely packed material, some plant voids, charcoal and mineral aggregates (TS07-10, PPL, 100X).

There are some clay coatings present on aggregates within the groundmass of these samples, most notably in slide TS07-9 (figure 3.37), that suggest water movement through these deposits at some point. The slide does not contain evidence of any compacted or mud-rendered surfaces, nor does it contain any evidence of mud-brick fragments. All the slides contain phytoliths, as well as spherulites. However, there is no direct evidence of animal penning, such as in situ coprolites, as seen in Phase 3. The lack of anthropogenic features such as floors or bricks, combined with the well-sorted water-laid sediment suggests that the southern portion of the courtyard was not inhabited during Phase 4.

111

Figure 3.37: Building Complex IV, Southern Courtyard. Some building fragments from this area had a clear, dark clay coating on the lower exterior surfaces (TS07-9, PPL, 50X). 3.3.7.4 Oven Feature, Courtyard

MF1a MF1b MF2 MF3a MF3b MF4 MF5 MF6 TS07-24 TS07-25

Table 3.16: An overview of micromorphology samples from an Oven in the courtyard south of Building Complex IV. The black squares represent the microfacies present in each sample slide.

112

Excavation revealed a burnt context along the edge of a pit feature in the courtyard of Building III that relates to the earliest occupation phases excavated during the 2005 season (figure 3.38). This was interpreted as an oven feature (MF 4) and two samples were taken from it.

Figure 3.38: Building III, Courtyard. This burnt level was interpreted as an oven feature and is located in the south- east portion of the courtyard.

The evidence from slide TS07-25 contains a high density of melted dung and organics as well as charred wood (MF 4). This is especially dense in the upper portion. Similar evidence is visible in some areas of TS07-24 as well. Both contain small fragments of burnt bone and charcoal (figure 3.39). Overall, this suggests that the feature was likely an oven and, as in the roasting pits, dung was used as fuel.

113

Figure 3.39: Building Complex III, Oven Feature. Fragments of a) charcoal b) ash and c) burnt bone, combined with melted coprolitic material suggest this area was likely an oven (MF 4) (TS07-24, PPL, 5X). 3.4 Summary of Results

Building Complex I appears to have been occupied consistently through all excavated Middle Chalcolithic phases at Tel Tsaf. The large open courtyard was used as an outdoor activity space, but the exact nature and location of activities are unclear. The main room, Room 70, was likely used as the main residence with evidence of well-constructed plaster floors found during both Phase 3 and 4. During Phase 3 there is evidence of extensive use of roasting pits, with at least nine of these features excavated. There is evidence of both wood and dung being used as fuel within these pits. Silos were present during both phases, although some were abandoned or rebuilt throughout the use of the building.

Building Complex II was only used during Phase 3 and appears to have had a very different use than Building Complex I. The two circular rooms, Rooms 230 and 263, appear to have been used as animal pens during Phase 3 as evidenced by the high numbers of phytoliths and

114 spherulites found in the micromorphology slides from these rooms. The baulk samples suggest that animals were not likely roaming freely within the courtyard, as spherulites are noticeably less dense in courtyard contexts. There is no evidence of roasting pits within the courtyard, and there are only two silos. This evidence suggests that the building complex, or a portion of it, served a purpose outside of everyday household activities. I suggest that it was a specialized penning area and potential processing area, rather than a residence.

Building Complex IV raises more questions than it answers. The Phase 4 architecture east of the main Silo 272 has not been exposed to its full extent so it is difficult to hypothesize its use(s). There are two potential destruction levels in the form of a mud-brick spill and large burnt level. There does appear to be some occupation between the two events, but its nature is unclear. The area directly east of Silo 272 appears unaffected by these events. A large burnt area was uncovered in the courtyard. There do not appear to be any clear indicators of an oven or pit, but the ashy area did contain fuel ash slag and three fragments of limonite. The microscopic evidence indicates that dung was used as fuel within this area. I would suggest the area was potentially a processing area, perhaps used to dye wool. Limonite, also known as yellow ochre, provides a vivid pigment when used as a dye. The use of ochre as a natural colourant is still prevalent today

3.5 Bulk Samples

Bulk samples were taken during the 2006 and 2007 seasons. These samples were intended to provide information from a wider variety of contexts than possible with micromorphology. The in situ data from micromorphology is invaluable, but time constraints combined with financial and excavation limitations make it impractical to sample every exposed context. The less costly alternative, bulk samples were taken from horizontal exposures as well as pits, such as the roasting features in Building Complex I. Samples were also taken to correlate with micromorphology samples. 119 samples were collected across all phases and features excavated.

Samples were taken with a trowel and placed in plastic zip lock bags. The bags were left open over night after sampling to allow moisture to evaporate. There was no pre-determined sample size, each ranging between three and ten trowel scoops. A total of 45 samples were selected and

115 processed for microscopic analysis. In order to ensure every context was represented, samples were arranged by context and random samples were selected from each grouping. A total of 20 samples were selected in this manner. The remaining 25 samples were randomly selected.

Initial analysis of micromorphology slides from 2006 identified excellent preservation of phytoliths and fecal spherulites in the sediment. The unique cell shapes of different plants help archaeologists identify the species represented by phytoliths visible in microscopic analysis. In the Near East, phytolith studies identifying wild and domesticated wheat and barley have provided an excellent framework for further analysis in the region (Rosen 1992; 1993; 1999; Rosen and Weiner 1994). Standard preparation of phytolith-bearing sediment involves separating phytoliths using heavy-liquid flotation following chemical treatment and wet-sieving (Piperno 2006). This technique was not employed during the present analysis. The preservation and abundance of phytoliths made them clearly visible on the mounted sediment slides from Tel Tsaf (figure 3.40). Samples were not taken in the field specifically for phytolith analysis and this project was not intended as a phytolith study. Fecal spherulites were also clearly visible in the bulk sample slides (figure 3.41).

Figure 3.40: Phytoliths from Tel Tsaf bulk sample. The upper left form (circled) is a wheat or barley fragment. The two other circled examples are grass forms. (sample 230-3, PPL, 200X).

116

Figure 3.41: Example of fecal spherulites preserved at Tel Tsaf. The circles highlight some of the most clear examples, but the spherulites are present throughout the slide (sample BC I – 6, XPL, 20X).

In order to understand the distribution of phytoliths and fecal spherulites at Tel Tsaf further, the fine fraction was mounted on slides for analysis. Each bulk sample was sorted using a mechanical shaker for 15 minutes. The sediment was sieved with the following screen sizes: 2 mm, 1 mm, 710 μm, 500 μm, and 250 μm. The finest fraction, <250 µm, was then sent to the Petrographic Thin Section Lab at McMaster University for preparation and mounting on thin section slides. Prior to mounting the sediment, a number of standard glass microscope slides were frosted on one side with 5 µm polishing media to assist sample adhesion and smearing. Approximately 5 mg (+/-0.1 mg) of each sample was weighed onto each slide. Several drops of 95% ethanol, adjusted to 8.0 pH, were placed on the slide, which was then rolled back and forth to smear the sample. After all the ethanol had evaporated, a thin layer of epoxy resin (Buehler Epoxide and hardener) was applied on top of the sample. The slide then cured for 24 hours and was ground to approximately 50 µm in thickness using a diamond cup wheel lapper (Buehler Petrothin system). Sample thickness was further reduced by hand lapping on a glass plate using a slurry of 5 µm aluminum oxide powder and light mineral oil, until a final sample thickness of 25-30 µm was reached.

117

Each sample was viewed in plane-polarized and cross-polarized light at various magnifications. Densities for fecal spherulites and phytoliths were classified as sparse, moderate or abundant. Phytoliths were, in general, denser than fecal spherulites across all samples. In both cases features were considered sparse if they appeared in a few fields of view or not at all. Moderate density is marked by the presence of one or two features in more than half the fields of view and abundant density is represented by the appearance of at least one feature in the majority of or all fields of view. The table below provides examples of sparse, moderate and abundant distributions for both phytoliths and spherulites (table 3.17).

118

Table 3.17: Examples of densities of fecal spherulites and phytoliths from Tel Tsaf samples. In each example either fecal spherulites or phytoliths are circled as appropriate. All fecal spherulite images are in XPL and all phytolith images are PPL. All images were taken at 200X magnification.

119

Bulk samples were collected from various locations and elevations across the site. Figure 3.42 shows the location of the samples discussed in the remainder of this chapter.

Figure 3.42: Location of bulk sample collection. The dots represent samples discussed in the text, not all samples taken. Building Complex I and II are from phase 3 and Building Complex IV is from Phase 4.

120

The main architectural features on the site are room interiors, courtyards and silos. Roasting pits were visibly present within the Building Complex I courtyard, and micromorphology identified the presence of potential animal pens in Building Complex II. Therefore, the results of the bulk sample analysis will be discussed by context: room interiors, courtyards, silos, roasting pits and animal pens. In order to illustrate the relative densities numerically, each category was assigned a value on an ordinal scale: absent = 1, sparse = 2, moderate = 3, abundant = 4. Below is a table summarizing the distribution of densities by context (table 3.18). Frequency spherulite phytolith total (#) 1 2 3 4 1 2 3 4 samples

Room 3 3 2 0 1 2 3 2 8 Interiors Courtyard 2 5 10 1 0 2 9 7 18 Roasting Pit 0 0 1 2 0 0 1 2 3 Silo 0 2 3 0 0 1 3 1 5 Animal Pen 0 1 6 3 0 4 1 5 10

Table 3.18: Bulk sample results by context. The number of samples per relative abundance category for both fecal spherulites and phytoliths are represented.

Figure 3.43: Cumulative frequency graph illustrating fecal spherulite relative abundance by context.

121

A cumulative frequency graph (figure 3.43) demonstrates the variation in fecal spherulite relative abundance distribution by context. Almost 80% of room interiors have only sparse spherulites, while they and silos never have more than moderate spherulite content, demonstrating that very few, if any, samples contain dense abundance of fecal spherulites. It is notable that 75% of all room interior samples contain sparse or absent fecal spherulites. The roasting pit samples do not contain lower than moderate relative abundance and the majority of samples contain abundant fecal spherulites. Unfortunately there are only three samples in the roasting pit category, making the results important but less statistically significant. The highest percent of animal pen samples contain moderate relative abundance , but approximately 30% contain abundant fecal spherulites.

Figure 3.44: Cumulative frequency graph illustrating trends in phytolith relative abundance by context.

The phytolith cumulative frequency graph (figure 3.44) demonstrates the relative abundance distributions across contexts. Room interiors are the only contexts that contain no phytoliths at all. The remaining interior samples are relatively evenly distributed between sparse, moderate and abundant relative density. In the courtyard context, 80% of samples have moderate or abundant relative density of phytoliths, with 50% of those being moderate. The majority of silo samples (c. 60%) contain moderate relative abundance of phytoliths with the remaining samples evenly distributed between the sparse and abundant categories. Roasting pit phytolith densities

122 directly correspond to the fecal spherulite densities with no samples containing less than moderate relative abundance and the majority containing abundant phytoliths, although the sample size is low. The animal pen samples plateau between sparse and moderate phytolith relative abundance but raise sharply with over 50% of samples containing abundant density.

Figure 3.45: Cumulative frequency graphs of fecal spherulites and phytoliths by context. These graphs demonstrate the relationship between the relative abundance of both features in each discrete context sampled at Tel Tsaf.

123

Figure 3.45 illustrates the overall patterns in fecal spherulite and phytolith densities by context. In summary, the roasting pits represent the highest densities of both fecal spherulites and phytoliths. Animal pens also tend towards higher densities of both, but there does not appear to be a direct correlation between fecal spherulite and phytolith relative abundance. In contrast, the room interiors, courtyards and silos consistently demonstrate lower relative abundance of fecal spherulites than phytoliths. The room interiors present the lowest relative abundance for both features. The courtyard and silo samples are most similar, with densities of both fecal spherulites and phytoliths predominantly in the moderate range, with mostly moderate density of each.

Grain-size analysis was undertaken on the bulk samples to determine whether there was any correlation between the fecal spherulite – phytolith data and sediment grain-size distribution. Having successfully noted certain trends in distribution of fecal spherulites and phytoliths across various contexts at Tel Tsaf, the percent of sediment at each sorting interval was plotted. After each bulk sample was sorted following the protocol discussed above, the sediment from each sieve was weighed (appendix II). There was no standardized sampling size, so the total weights for each grain-size category was normalized by calculating what percent of the sample fell within each grain-size. The mean grain-size of each context identified above (room interiors, courtyards, silos, roasting pits, animal pens) were plotted against the mean grain-size of all the bulk samples from the site (figure 3.46).

124

Figure 3.46: The mean grain-size distribution from each context at Tel Tsaf and the overall mean distribution of all samples. The standard deviation (1 σ) for the mean grain-size of all bulk samples is plotted in red. Intervals on the x- axis are non-linear and represent screen sizes rather than an equal distribution of grain-sizes.

The graph in figure 3.46 demonstrates that there is very little variation in grain-size distribution across all contexts at Tel Tsaf. With the exception of silos, all contexts follow the same distribution trend as the overall average, suggesting there is very little variation in grain-size distribution across contexts. To test this, the standard deviation of each context mean was plotted to 1 σ (68% confidence) (figure 3.47). It is clear that there is no statistically significant variation in grain-size distribution across contexts at Tel Tsaf. There is some variation in individual samples within each context that will be discussed below, along with the fecal spherulite and phytolith data, but it should be noted that sediment size is not a significant indicator of use of space within this study.

125

Figure 3.47: Mean grain-size distribution from all contexts at Tel Tsaf with Standard Deviation (1 σ) for each context and grain-size. Intervals on the x-axis are non-linear and represent screen sizes rather than an equal distribution of grain-sizes.

126

3.5.1 Room Interiors

The room interior samples were taken from Room 70 (Building Complex I) and Room 662 (Building Complex IV).

Absent Sparse Moderate Abundant (1) (2) (3) (4) 70-1

70-2

70-3

70-4

70-5

70-6

612-1

612-2

Fecal Fecal Fecal Fecal Phytolith Phytolith Phytolith Phytolith Spherulite Spherulite Spherulite Spherulite

Table 3.19: Relative abundance of fecal spherulite and phytolith densities from all interior room contexts at Tel Tsaf.

A comparison of the relative abundance data demonstrates significant variation in densities across the samples, with three groups apparent (table 3.19). The first group, samples 70-1, 70-2, 70-3, contain sparse or absent fecal spherulites and sparse to moderate phytoliths. The second group, also from Room 70 (70-5 and 70-6) has moderate fecal spherulite distribution and moderate to abundant densities of phytoliths. The final group, Room 612 samples (612-1 and 612-2), contain sparse or absent fecal spherulites and sparse to high densities of phytoliths.

The sparse/absent fecal spherulite and phytolith cluster of samples from Room 70 (70-1. 70-2, 70-3) contain exposed mud-rendered floor levels. The limited evidence of fecal spherulites within the floor samples indicates that dung was not used as a building material. In contrast to the floor cluster, the second group from Room 70 (samples 70-5 and 70-6) are both from fill levels. Room 70-5 was taken from above the mud-rendered floor and Room 70-6 was taken from a niche in the eastern wall of the room. These fill levels are more likely to contain a mix of

127 materials, which explains the increase in both fecal spherulites and phytoliths. These results clearly demonstrate the difference between floor and fill levels within room 70.

The samples from Room 612 are low in fecal spherulites and highly variable in phytoliths. They do not show the characteristics of either floor or fill levels from the Room 70 groups, but it is not surprising as no constructed floor levels were exposed in the room. As discussed in section 3.1, the actual extent of Room 612 was not defined during excavation, and walls and floors were not apparent. Given the variation from the Room 70 samples, it seems unlikely the third group represents a discrete room interior type at Tel Tsaf.

The grain-size distribution analysis from the Room Interiors also break down into three discrete, yet similar groups (figure 3.48).

Figure 3.48: Grain-size distribution of bulk samples from room interiors at Tel Tsaf. Intervals on the x-axis are non- linear and represent screen sizes rather than an equal distribution of grain-sizes.

Samples 612-1, 612-2 and 70-4 contain 55% or more sediment with a grain-size of less than 250 µm, with the remaining sediment falling in line with the other interior samples. The Room 612

128 samples likely represent disturbed fill layers, and 70-4 was taken from fill levels in Room 70, which likely explains the similarity in distribution. Samples 70-1, 70-2, 70-5 and 70-6 all contain very similar profiles with the exception of an increase in sediment in the less than 500 µm range in sample 70-1. There is no obvious explanation for this variation. This group of samples contain both fill and floor sediment from Room 70, making any meaningful analysis difficult. The final sample, 70-3, shows the greatest divergence from the general distribution of building interior samples. It contains very little fine sediment, and consists of over 60% sediment over 2 mm in size. This sample contained large fragments of a white floor level and theses pieces of mud-rendered floor likely account for the variation in distribution.

Overall, the fecal spherulite and phytolith distribution does correlate with sample context within room interiors. It is possible to make a reasonable assumption regarding fill and floor levels based on the distribution of these features. The grain-size analysis, unfortunately, does not demonstrate a significant correlation between sample context or fecal spherulite and phytolith distribution.

129

3.5.2 Courtyards

Courtyard samples were taken from all three building complexes (I, II,

Absent Sparse Moderate Abundant (1) (2) (3) (4) BC I - 1

BC I -4

BC II - 8

BC II - 9

BC IV - 11

BC I - 3

BC I -5

BC I - 6

BC I - 7

BC IV - 16

BC IV - 17

BC I - 2

BC IV - 10

BC IV - 12

BC IV - 13

BC IV - 14

BC IV - 18

BC IV - 15

Fecal Fecal Fecal Fecal Phytolith Phytolith Phytolith Phytolith IV). Spherulite Spherulite Spherulite Spherulite

Table 3.20: Relative abundance of fecal spherulites and phytoliths from all courtyard contexts at Tel Tsaf.

Courtyard contexts at Tel Tsaf generally contained moderate to high densities of phytoliths and varying levels of fecal spherulites with five groupings visible in the chart (table 3.20). The first group (BCI-1, BCI-4, BCII-8, BCII-9, BCIV-11 ) shows remarkable consistency in densities of both elements with all samples containing sparce fecal spherulites and moderate phytoliths. The second group contains samples from Building Complex I (BCI-3, BCI-5, BCI-6, BCI-7) and Building Complex IV (BCIV-16, BCIV-17). All these samples have moderate fecal spherulite and abundant phytolith densities. The third group contains both Building Complex II samples

130

(BCII-8, BCII-9) and Building Complex IV (BCIV-11), and is marked by sparse fecal spherulite and moderate phytolith relative abundance. The next group contains two samples from Building Complex IV (BCIV-14 and BCIV-18) that do not contain any fecal spherulites and only sparse phytoliths. The final sample is Building Complex IV-15, and contains abundant fecal spherulites and phytoliths.

The first group of courtyard samples had sparce fecal spherulites and phytoliths, and most slides also contained very small phytolith fragments and ash, as well as burnt material. This is consistent with the evidence from the micromorphology slides from the Building Complex I courtyard (TS06-8, TS06-9) which displayed evidence of ash and charred wood (MF 2 – courtyard surfaces). The presence of moderate relative abundance of fecal spherulites suggests that dung was likely used as fuel in the roasting pits in the courtyard. The Building Complex IV samples (12, 13) that contain similar evidence are from an extensive burn level located northeast of Silo 272. The burn level was not bounded by any architectural features that suggest its function, and it is possible the area was an ash tip.

The second group (BCI-3, BCI-5, BCI-6, BCI-7 and BCIV-16, BCIV-17) also contains moderate fecal spherulite densities, but more abundant phytoliths than the first group. Other features, such as ash and charcoal, are also present as the samples were taken from contexts similar to those from group one. The similar fecal spherulite relative abundance further supports the use of dung as fuel within roasting pits in Building Complex I. The moderate fecal spherulite densities from the Building Complex IV samples, also taken from the burn level in the courtyard, suggest that whatever the origins of the ash, the responsible fire was fueled by dung. This is further supported by bulk sample BCIV-15 which contained high densities of both fecal spherulites and phytoliths. This sample was taken from the same burnt area in the Building Complex IV courtyard as the other BCIV samples already discussed. The BCIV-15 sample included material directly associated with fuel ash slag. Given the use of dung within the roasting pits of Building Complex I and the presence of fecal spherulites, it can be assumed that the burnt ashy layer in Courtyard Complex IV originated from intentional cooking or production fueled by dung, rather than unintentional means.

131

The third courtyard type apparent in table 3.20 contains all the Building Complex II samples, and is defined by the absence of fecal spherulites and only moderate levels of phytoliths. This is somewhat surprising, given the evidence of animal pens from the buildings within the courtyard of Building Complex II. However, the limited evidence of fecal spherulites suggests that any animals housed in Rooms 230 and 263 were not granted access to the courtyard. In addition, the lower phytolith densities strongly suggest that very different sets of activities were taking place in Building Complex II. The lack of roasting pits and absence of rooms with mud-rendered floor surfaces already indicate that the excavated portion of Building Complex II was not a domestic dwelling, and the dissimilarities between bulk sample results further indicate that the building was used for different tasks than Building Complex I.

The final courtyard relative abundance group is more enigmatic, as the samples contained within it (BCIV-14 and BCIV-18) were taken from the same general area as the other Building Complex IV courtyard samples but contain no fecal spherulites and only sparse phytoliths. BCIV-14 was taken from the lowest level of the burn area, with sediment from the layer beneath incorporated. This may explain the lower fecal spherulite densities from these samples as there was less burnt material. BCIV-18 included visibly white sediment within the burn area, likely representing burnt building material or dense ash.

Given the variability in fecal spherulite and phytolith densities in the courtyard samples, the grain-size distribution is remarkably uniform within all samples (figure 3.49).

132

Figure 3.49: Grain-size distribution of bulk samples from courtyards at Tel Tsaf. Intervals on the x-axis are non-linear and represent screen sizes rather than an equal distribution of grain-sizes.

There does appear to be some variation in actual percent of sediment less than 250 µm (between 35% and 65%), but with the exception of BCI-2, all the courtyard samples contain less than 20% of any other grain-size. The one noticeable outlier, BCI-2, was taken from the northern area of the Building Complex I courtyard. There is nothing regarding the sample location that would explain the variation seen in the grain-size distribution.

Once more the fecal spherulite and phytolith distribution demonstrated some correlation with sample material from courtyard contexts allowing identification of courtyard rake-out rather than a potential ash tip or processing area. The densities also demonstrated a difference in distribution between Building Complex I and Building Complex II that coincides with other evidence that suggests each courtyard was used for a different purpose. The grain-size distribution did not contribute to these observations.

133

3.5.3 Silos

During excavations, the interiors of two silos were exposed, allowing access to interior bulk sediment samples. The silo samples come from Silo 339 in Building Complex I and Silo 272, which is present in both Building Complex II and IV.

Absent Sparse Moderate Abundant (1) (2) (3) (4) 339 - 2

272 - 3

339 - 1

272 - 1

272 - 2

Fecal Fecal Fecal Fecal Phytolith Phytolith Phytolith Phytolith Spherulite Spherulite Spherulite Spherulite

Table 3.21: Relative abundance of fecal spherulites and phytoliths from all silo contexts at Tel Tsaf.

All of the silo samples generally contain sparse to moderate fecal spherulites and phytoliths, but two groups could be discussed (table 3.21). The first group (339-2 and 272-3) contains sparse fecal spherulites and moderate phytolith densities. In contrast, the second group (339-1, 272-1, and 272-2) contains moderate fecal spherulite densities. The phytolith densities are varied with 272-1 containing abundant, 272-2 displaying sparse and 339-1 containing moderate densities. During sampling of both Silo 339 and Silo 272, sediment was collected from various levels. There is nothing in the sampling strategy that provides a specific explanation for the varied densities in the five samples. One potential explanation is that different products may have been stored at different times. In addition, some samples may represent storage levels and others may be fill layers. In general, the moderate, rather than abundant, phytolith levels are surprising as one might expect a high relative abundance if grain was stored in the silos. Furthermore, the presence of sparse to moderate fecal spherulite densities is unexpected. It is not clear why fecal matter would show up in storage areas. It is possible that the increased levels of fecal spherulites may help identify fill levels, especially if material from the courtyards was used to level off or close the silos.

134

Grain-size data is only available for three of the five silo samples and demonstrate some degree of variation in distribution (figure 3.50).

Figure 3.50: Grain-size distribution of bulk samples from silos at Tel Tsaf. Intervals on the x-axis are non-linear and represent screen sizes rather than an equal distribution of grain-sizes.

The grain-size distribution from the three silo samples available (339-1, 339-2, 272-3) follow the same general distribution pattern seen in other contexts with the smallest and largest fractions being most abundant. There is some variation in the percent of each grain-size in each silo sample, but this is likely related to whether it is from a storage or fill level.

The sample size for silos is quite small, but some potential patterns can be seen. Potential fecal spherulite and phytolith relative abundance and grain-size distribution could help distinguish between and fill or storage deposits. However, given the lack of secure provenience and small sample size from silos, it is premature to make any definitive statements.

135

3.5.4 Roasting Pits

Three separate roasting pits were sampled in the Building Complex I courtyard.

Absent Sparse Moderate Abundant (1) (2) (3) (4) RP - 1

RP - 2

RP - 3

Fecal Fecal Fecal Fecal Phytolith Phytolith Phytolith Phytolith Spherulite Spherulite Spherulite Spherulite

Table 3.22: Relative abundance of fecal spherulites and phytoliths from all roasting pit contexts at Tel Tsaf.

The roasting pits present the highest densities of fecal spherulites and phytoliths at Tel Tsaf (table 3.22). The first two samples, RP-1 and RP-2, contain high densities of both features. The third sample has noticeably fewer fecal spherulites and phytoliths. This, however, can be explained by the sample context. RP-3 was taken from the very bottom of roasting pit 284, and the sample thus contained a mix of pit and courtyard sediment. The presence of non-pit sediment likely contributed to the reduced relative abundance of spherulites and phytoliths. All samples contain predominantly grass phytolith forms. The presence of both fecal spherulites and phytoliths in high densities is indicative of dung being used as a fuel source within the roasting pits.

The grain-size distribution for the roasting pits’ sediment demonstrates little variation among samples (figure 3.51).

136

Figure 3.51: Grain-size distribution of bulk samples from roasting pits at Tel Tsaf. Intervals on the x-axis are non- linear and represent screen sizes rather than an equal distribution of grain-sizes.

There is nothing significant about the grain-size distribution of the roasting pits, as they follow the same general pattern seen previously. The strongest indicator of roasting pits at Tel Tsaf remains the high relative abundance of both fecal spherulites and phytoliths in the bulk samples.

3.5.5 Potential Animal Pens

Bulk samples were collected from the two circular rooms, 230 and 263, in Building Complex I. Both rooms are from the Phase 3 occupation and no similar structures were found in the Phase 4 complex.

137

Absent Sparse Moderate Abundant (1) (2) (3) (4)

230-1

230-2

230-3

230-4

230-5

263-1

263-2

263-3

263-4

263-5 Fecal Fecal Fecal Fecal Phytolith Phytolith Phytolith Phytolith Spherulite Spherulite Spherulite Spherulite

Table 3.23: Relative abundance of fecal spherulites and phytoliths from all animal pen contexts at Tel Tsaf.

The data from the two animal penning contexts groups vary clearly by room. Room 230 contains moderate to abundant densities of fecal spherulites and abundant phytoliths. The phytolith fragments are generally larger than in other contexts at Tel Tsaf and exhibit a mix of wheat/barley and grass forms. In contrast, there are sparse to moderate phytolith densities in Room 263 and the fragments are generally smaller. The samples also contain sparse to moderate fecal spherulites.

The micromorphology slides from the two rooms also show some marked difference, leading to the conclusion that the sediment in Room 263 was either re-deposited or suffered greater post- depositional bioturbation. The bulk sample results further support this difference. The similar fecal spherulite densities suggest the presence of significant dung deposits in both rooms and the lack of ash and charcoal indicates a direct deposit rather than burnt dung. The larger phytolith fragments in Room 230 support in situ penning deposits, and the smaller, fragmented phytoliths from Room 263 may be the result of greater post-depositional disturbance or redeposit of penning material. The most dissimilar sample from Room 263 (263-1) was taken from the

138 entrance to the building, which likely explains the sparse distribution of both phytoliths and fecal spherulites.

The grain-size distribution is fairly consistent, but a few observations can be made based on the animal pen samples (figure 3.52).

Figure 3.52: Grain-size distribution of bulk samples from animal pens at Tel Tsaf. Intervals on the x-axis are non- linear and represent screen sizes rather than an equal distribution of grain-sizes.

The samples can be roughly grouped into two distribution patterns: Those with greater than 50% sediment measuring less than 250 µm, and those with less than 50% measuring less than 250 µm. All but one of the samples with over 50% are from Room 230 and only one of the samples in that same group is from Room 263 (263-1). There are four samples in the less than 50% group (230- 1, 263-3, 263-4, 263-5) and each also demonstrates a higher percent of sediment from the >1mm sieve. In fact, these samples tend towards an overall higher percent of larger fraction than the other samples. This demonstrates a general agreement with the variation in fecal spherulite and phytolith distribution, with variation seen between Rooms 230 and 263.

139

Overall, the animal penning samples are distinctive in their fecal spherulite and phytolith densities, with some variation between in situ and disturbed penning sites. These variations may also be correlated with grain-size distribution, with a higher percent of fine fraction (< 250 µm) associated with the in situ sediment in Room 230. This may be related to trampling and cleaning activities within this area. The increase in larger fraction material (>710 µm) in the Room 263 material is likely a result of the disturbed nature of the sediment. The absence of charcoal and increase in phytolith size and type (barley and wheat) can distinguish pens from roasting pits and courtyard rake-out.

3.6 Chapter Summary

The micromorphology data identified six microfacies associated with human activity at Tel Tsaf: constructed floors (smooth and rough); courtyard surface material; animal pens (in situ and disturbed); ovens; conflagration and general fill. Analysis of the relative abundance of fecal spherulites and phytoliths in bulk samples demonstrates a correlation with distinct activity areas: room interiors; courtyards; silos; roasting pits and animal pens. Grain-size distribution did show some potential in aiding in identification of floor or fill levels in room interiors, and potentially distinguishing between in situ and disturbed animal pens, but overall it is not a significant indicator of use of space. The correlation between the micromorphology data and the bulk sample data is discussed by building complex below.

Building Complex I provides good data for understanding the correlation between the micromorphology and bulk data sets (figure 3.53). The room interiors are dominated by MF 1a and MF 6, which represent well-laid mud-rendered floors and fill levels. The bulk samples showed similar clustering, with samples from observable floor levels containing either no fecal spherulites and phytoliths, or sparse evidence of each. The fill levels are represented by samples containing moderate to high densities of both components.

The courtyard micromorphology samples contain MF 1b, MF 2 and MF 6, indicating rough-laid floors, courtyard surfaces including rake-out from roasting pits, and general fill. The bulk samples all contain moderate fecal spherulite densities and moderate to abundant phytolith

140 densities. All the samples also contain charcoal, indicating their correlation to the courtyard surface (MF 2) levels in the micromorphology slides.

Figure 3.53: Building Complex I noting all bulk samples and relative abundance of fecal spherulites and phytoliths are plotted and microfacies present in each context.

141

The silo sampled contained evidence of roughly constructed mud-brick (MF 1b) and general fill (MF 6). The bulk samples contained sparse to moderate densities of fecal spherulites and phytoliths. The bulk results appear to be consistent with fill levels from Room 70. No micromorphology samples were taken from the roasting pits, but the bulk sample data is distinct from this context. The highest densities of fecal spherulites and phytoliths were found in the roasting pits, indicating that dung was used as a fuel source. These densities also explain the composition of the floor surface material in the surrounding courtyard, which likely contains rake-out from these pits.

Building Complex II provides an excellent overview of the correlation between micromorphology and bulk sample data in animal pens. Room 230 presented the best in situ micromorphology evidence of penning activity (MF 3a) and the relative abundance of fecal spherulites and phytoliths support this interpretation. The bulk samples generally contained high densities of both components, but did not have any charcoal and are therefore distinguishable from the roasting pit samples. The fill levels (MF 6) in Room 230 were taken from the pit dug into the Phase 3 levels, and no bulk samples were analyzed from this context. The disturbed penning (MF 3b) evidence from Room 263 is also reflected in the bulk sample data. These samples contain fewer fecal spherulites and far fewer phytoliths than the Room 230 samples. The sediment in Room 263 suffered greater bioturbation, or was potentially re-deposited. The reduced densities reflect this activity.

142

Figure 3.54: Building Complex II noting all bulk samples and relative abundance of fecal spherulites and phytoliths are plotted and microfacies present in each context.

The Silo 272 micromorphology samples contained well-constructed mud-brick (MF 1a) and generally fill (MF 6). The bulk samples present the same general abundance of fecal spherulites and phytoliths as seen in the Silo 339 samples. It is likely these levels represent general fill, which is characterized by sparse to moderate relative abundance of fecal spherulites and moderate to abundant phytolith relative abundance.

143

Figure 3.55: Building Complex IV. All bulk samples and relative abundance of fecal spherulites and phytoliths are plotted and microfacies present in each context are noted

Building Complex IV is the least well defined architectural unit at Tel Tsaf. Room 612, the extent of which has not been fully identified, contains micromorphology evidence of rough-lain courtyard floors (MF 1b) but no evidence of courtyard surfaces (MF 2). The bulk sample data does not indicate such a surface either, with relative abundance more in line with the Room 70 floor levels. The other microfacies present, MF 5 and MF 6 (conflagration and fill levels) are not represented at all in the bulk data.

144

The majority of the bulk data from Building Complex IV comes from the courtyard area to the west of Silo 272. This area contained an extensive ash and charcoal scatter that included slag fragments. The scatter is not contained in a pit, nor is it defined by any architectural features. Some bulk samples contain moderate to abundant densities of fecal spherulites and phytoliths, while others contain sparse of no evidence of either. This variation is generally explained by sampling context, with the samples with lower densities containing fill sediment, and those with higher densities taken from areas with larger concentrations of burnt material. It is likely the ashy and burnt sediment was the result of dung-fueled fires, and may represent an ash dump.

Room 662 presents micromorphology evidence of rough-laid floors (MF 1b), courtyard surfaces (MF 2) and general fill (MF 6). The full extent of this room has not been defined, but it does appear that the material has been highly disturbed. The courtyard area to the south of the main Building Complex IV architecture presents micromorphology evidence of rough-made mud- brick (MF 1b), disturbed penning (MF 3b), general fill (MF 6) and what appears to be an oven (MF 6). It is unclear whether this courtyard is directly associated with Building Complex IV. There is no bulk sample evidence from Room 662 or the southern courtyard.

An overall summary of the main context and the indicative microfacies, and fecal spherulite and phytolith densities is presented in table 3.25

145

Fecal Spherulite Microfacies Phytolith Density Density MF 1a well-constructed absent - sparse sparse - moderate mudrender floor Room Interior MF 6 moderate moderate - abundant general fill MF 1b rough-laid mudrender floor Courtyard MF 2 moderate moderate - abundant courtyard surface MF 1a well-constructed mudbrick

Silo MF 1b rough-made mudbrick

MF 6 sparse - moderate sparse - abundant general fill

Roasting Pit abundant abundant

MF 3a moderate - abundant moderate - abundant in situ animal pen Animal Pen MF 3b moderate sparse - moderate disturbed animal pen

MF 5 Conflagration conflagration

MF 4 Oven oven

Table 3.24: An overview of the main contexts identified at Tel Tsaf and corresponding micromorphology and bulk sample evidence.

Having established the use of architecturally distinct areas at Tel Tsaf, it is important to establish how these contexts are formed, and their significance to the population of Tel Tsaf. Chapter 4

146 will discuss how these spaces may have been used by the inhabitants of the site, and the role each area may have played in daily life.

147

Chapter 4 A Life History of Tel Tsaf

The geoarchaeological data presented in Chapter 3 suggest some useful results regarding how space at Tel Tsaf was used. These analyses allow identification of rough floor surfaces, mud- rendered floors, animal pens and dung-fueled fires that were otherwise invisible macroscopically during excavation of these building complexes. These elements individually tell us about the types of activities that took place on the site. This chapter will bring that information together to help create a more complete picture of the archaeological remains and to detail what life at Tel Tsaf may have been like during its occupation in the Middle Chalcolithic beginning with a discussion of the formation processes involved in creating the contexts identified through micromorphology and bulk-sample analysis in Chapter 3. I will then discuss the silos and present evidence for what they may have been used to store. The discussion will then move onto livestock and their role in domestic economies at Tel Tsaf, with particular attention to the use of animal pens. Evidence for dung as fuel in both cooking and manufacture raises questions regarding the use of pyrotechnology at Tel Tsaf. Finally, I will address the significance of changes and continuity in architecture between Phases 4 and 3 at the site.

4.1 Site Formation Processes

Following from the discussion of sediment components and microfacies in Chapter 3, we can look at the anthropogenic processes that created the contexts at Tel Tsaf. Below are three schematics providing an overview of activities that likely contributed to the creation of specific areas on site (figures 4.1 to 4.3). Each Building Complex is displayed separately, alongside flat- bed scans of two or three micromorphology slides from different contexts. This overview is intended to demonstrate the visible variability of each context, even at the macroscopic level, and illustrate how humans and animals may have interacted within each surface to create distinctive signatures in the sediment record. The figures also provide greater insight into what activities were likely to have taken place within each occupation area.

148

Figure 4.1: Formation processes evident in two micromorphology samples from Building Complex I. A comparison of the visual aspects and formation activities highlight the differences between interior living areas (Room 70) and unroofed courtyard spaces.

149

Figure 4.2: Formation processes evident in three micromorphology samples from Building Complex II. The three slides clearly demonstrate the variation in activity during Phase 3. It is easy to see the difference between the Room 230 sediment (top) and the pit from the same baulk (middle). The bottom slide highlights the difference between in situ penning activity (top) and disturbed or re-deposited penning material. TS07-2 has some microscopic components illustrated, as there is very little macroscopic material to discuss.

150

Figure 4.3: Formation processes evident in three micromorphology samples from Building Complex IV. There are some similarities between the samples from Room 662 (top) and Room 612 (middle). They both contain fragmented courtyard floors, but 662 is far more disturbed. The Room 612 sample contains some evidence of courtyard surface spread, similar to Building Complex IV, but is less distinctive. The silo sample consists entirely of large and small mudbrick fragments and fill material.

151

4.2 Silos

The identification of the circular, paved structures from the Tel Tsaf courtyard as silos is an interpretation that is broadly accepted within the current literature (Banning 2010; Ben-Shlomo, et al. 2009; Eitam 2009; Garfinkel, et al. 2009; Garfinkel and Ben Shlomo 2007; Garfinkel, et al. 2007; Graham 2014; Hill 2011; Hubbard 2010; Rowan and Golden 2009; Rowan and Kersel 2014; Shennan 2011). The classification of similar features as silos is common in the southern Levant and widely applied from the Neolithic through the historic periods (Banning 2010; Currid 1985; Kuijt 2008). This interpretation is influenced by iconography from ancient Egypt. An Old Kingdom model from the IV dynasty tomb at Kamena illustrates 12 rounded “beehive silos” within a courtyard (Currid 1985). It appears most Egyptian and Mediterranean silos consisted of a rounded base, with a plastered corbel roof that inspired the term beehive silos. Images from Thebes and Bene Hasan clearly illustrate this shape and illustrate the use of ladders to pour grain into a door at the top of the structures. They also show doors at the base, presumably to remove grain as needed (Currid 1985).

Excavations at the Pre-Pottery Neolithic A site of Dhra’ uncovered a number of paved circular features measuring c. 3 m in diameter (Kuijt and Finlayson 2009). These also contained upright stones, presumably to support a wooden floor. Kuijt and Findlayson (2009) suggest the structures would have had a superstructure that would have been effective in keeping the grain dry and free of pests. Micromorphology and phytolith analyses have identified higher levels of grain and legume evidence than anywhere else on the site. Other examples of probable Neolithic silos with similar designs are found at Yiftahel (Garfinkel, et al. 1987), Ain Ghazal (Rollefson 1997), Netiv Hagdud (Bar Yosef and Gopher 1997) and Jericho (Kenyon 1981). There is no evidence of superstructures preserved in any of these examples, as is the case at Tel Tsaf. Later period sites in the Levant provide the best data for beehive silos similar to those from Egypt, with corbelled superstructures preserved at Late Bronze Age II Bir el-Abd (Oren 1973) and Persian or Hellenistic period Tel Gemme (van Beek 1972).

Currid (1985) notes that early interpretations of stone- or brick- paved circular features include fortresses, palaces, sanctuaries, threshing platforms or potential hut foundations. It should be noted that the extensive geographic and chronological range of the above examples does limit the

152 degree to which similarities in architectural forms directly translate to actual function during different periods. However, as more evidence has been uncovered, and contexts at each site are considered, the interpretation of paved, rounded features as silos has become the most probable and widely accepted one. At Tel Tsaf, the combination of structure, courtyard location and large quantities of unprocessed cereal grains and local weeds (Graham 2012), strongly indicate that the features are silos.

The appearance of structures designed specifically for storage of surplus is always exciting as it has great implications for the discussion of wealth accumulation and socio-economic variability (see Chapter 2). When wealth is identifiable in the archaeological record, it becomes possible to discuss the relationships between those with wealth and those without. Grain silos have been cited as one of the potential clues in our detection of this emerging divide between the haves and have-nots (Banning 2010; B. F. Byrd 1994; Flannery 1972; Garfinkel, et al. 2009). Garfinkel et al. (2009) present a useful discussion of some of the implications of the silos at Tel Tsaf and their potential significance. While my data generally agrees with their conclusions with regard to silo area, my views on the impact these silos would have had on socio-economic organization at the site level are more conservative.

Circular silos, as well as large store jars, bins, storage rooms and other similar features occurred in the southern Levant from the Neolithic onwards (Amiran 1978; Bienert and Vieweger 1999; Braun 1989; Eisenberg, et al. 2001; Epstein 1998; Garfinkel 2004; Garfinkel and Miller 1999; Milevski, et al. 2012). A great deal of literature exists on interpreting the form, capacity and significance of such facilities (Banning 2010; Bourke 2001; B. F. Byrd 1994; Epstein 1998; Garfinkel, et al. 2009; Gilead, et al. 1995). There is limited evidence of storage facilities in sites of the Wadi Rabah cultural complex, with some evidence from Tabaqat al-Bûma (Banning 2010) and potential paved storage areas at Wadi Rabah (Kaplan 1958) and Basatîn (Kadowaki, et al. 2008). Rounded platforms from the Néolithique récent level at Byblos may have been silos similar to those at Tel Tsaf (Dunand 1973). Concentrations of olive stones in storage pits at Neve Yam indicate storage areas at the submerged site (Galili, et al. 1997). In most of these Wadi Rabah related sites the facilities appear to be associated with specific households, indicating private storage (Banning 2010). By the Late Chalcolithic, the most dominant storage

153 facilities are in-ground, stone-lined silos such as those at Tel Te’o (Eisenberg, et al. 2001) and Golan Site 18 (Epstein 1998). Alternative storage methods included large store jars, bell-shaped pits and paved store rooms (Banning 2010).

The silos at Tel Tsaf present some of the best evidence for storage during the late prehistory of the region and their presence within bounded courtyards raises interesting questions regarding their public versus private use. Table 4.1 provides an overview of the silos from the site, their diameter and estimated capacities if reconstructed as ~2 m in height (based on Garfinkel et al. 2009). The enclosed courtyard of Building Complex I implies that the silos are directly associated with the inhabitants of this complex. The presence of silos in Building Complex II also indicates different ownership or management than those of Building Complex I. It makes a strong case for private ownership of grain stores.

154

Silo Building Phase Diameter Area Estimated Number (m) (m²) capacity (m³) 286 I 3 3.00 7.05 14.10 339 I 3 3.00 7.05 14.10 415 I 3 2.50 4.90 9.80 53 I 3 2.90 6.60 13.20 288 II 3 2.20 3.80 7.60 272 II 3 3.00 7.05 14.10 415* I 4 2.50 4.90 9.80 633 I 4 1.60 2.00 4.00 66 I 4 1.90 2.80** 5.70** 171 I 4 2.20 3.80 7.60 74 I 4 2.00 3.15 6.30 288* IV 4 1.80 2.55 5.10 272* IV 4 4.00 12.60 25.20

Table 4.1: Number and dimensions of the silos from Phases 3 and 4 at Tel Tsaf. The estimated capacity is based on the assumption that each silo was ~2 m in height (data from Garfinkel et al 2009).

*These silos were renamed in Garfinkel et al 2009 during Phase 4 as follows: 415 = 603; 288 = 565; 272 = 568. The dimensions vary between phases but, since the location is unchanged, I have referred to these silos by their Phase 3 designations throughout this dissertation.

** These numbers have been changed from Garfinkel et al 2009 as the calculations provided did not correspond with the diameters as published. The above figures are based on a diameter of 1.90 m.

We can see an increase in total silo area in Building Complex I from the earlier Phase 4 to the later Phase 3. We can be certain that silo 633 was no longer in use during Phase 3, as Silo 53 is superimposed over the earlier structure. There is also a decrease in total area from Phase 4 Building Complex IV to Phase 3 Building Complex II (table 4.2). The overall change in size and layout of this complex likely plays a role in this variation. The architecture of Building Complex I remains otherwise constant in Phases 4 and 3, so the noticeable increase in silo floor area is even more enigmatic, assuming all silos were used contemporaneously. This will be addressed further in section 4.5, with social and economic implications discussed in the next chapter.

155

Phase 4 Phase 3

Silo Building Area Silo Building Area Number Complex (m²) Number Complex (m²) 415 I 4.90 286 I 7.05 633 I 2.00 339 I 7.05 66 I 2.80 415 I 4.90 171 I 3.80 53 I 6.60 74 I 3.15 TOTAL 16.65 TOTAL 25.60

Silo Building Area Silo Building Area Number Complex (m²) Number Complex (m²) 288* IV 2.55 288 II 3.80 272* IV 12.60 272 II 7.05 TOTAL 15.15 TOTAL 10.85

Table 4.2: Overview of silo area by building complex and phase.

4.3 Livestock

Much of the discussion about the accumulation and storage of surplus in prehistory focuses on grain storage, and this is especially true at Tel Tsaf. The remains of physical structures, combined with the physical labour required to build the structures and to farm the grains to fill them lend themselves well to such discussions. However, the evidence for animal penning areas at Tel Tsaf, as well as for the use of secondary products (discussed below), suggests that livestock were just as important in the socio-economic structure at the site as crops.

In order to understand the significance of the circular rooms in Building Complex II, it is first important to review the faunal evidence from the site. Faunal evidence of domesticated species tells us what animals were present and how they were used by the inhabitants of Tel Tsaf. These factors influence likely scenarios regarding the use of the Building Complex II pens. During excavation, 100% of the excavated sediment was screened through a 5mm mesh. Each field student was trained to spot faunal remains by the site specialist, Austin Hill. In total 19,983 identifiable specimens (bone and teeth) were recovered from Areas A and C. The collection was

156 primarily analyzed by relative abundance based on Number of Identifiable Specimen (NISP), as well as Minimum Number of Elements (MNE) and Minimum Animal Units (MAU) (Hill 2011).

Not surprisingly, sheep/goat represent the largest portion of faunal remains at Tel Tsaf, with a relative abundance (by NISP) of 40% of the assemblage representing these small ungulates. The composition of the assemblage suggests that sheep and goat were raised with the primary goal of meat production and herd maintenance. Mortality patterns were calculated by creating post- cranial bone fusion mortality profiles based on Silver (1969) as most previous Chalcolithic collections also used Silver’s profile (Hill 2011). Kill patterns suggest that most animals were killed prior to maturity, but no set of the population is slaughtered particularly young either. There is an even kill pattern across both male and female sheep and goat, with 70% of usable elements representing individuals that survived to be at least 1.5 years old. There is a shift in age distribution from Phase 4 to Phase 3, with a slight increase in older animals in the later phase. While it is unlikely that this pattern represents intensive exploitation of wool or dairy, it is likely that both were being used on some scale during Phases 3 and 4.

Cattle represent the least abundant domestic animal taxon at Tel Tsaf. Cattle bones and teeth make up 13% (by NISP) of the faunal assemblage. In contrast to the sheep/goat kill patterns, bone fusion and dental analysis show that the majority of identified elements (70%) represent cattle that lived past their 5th year. Dental analysis suggests that 30% of teeth come from cows that were at least 6 years or older (Hill 2011). While this pattern may suggest that the cattle were exploited for milk, there is no characteristic slaughter of calves after weaning which often accompanies herds maintained specifically for dairy production. It is likely that dairy production was taking place to some extent, but not on a large scale. It is also very likely that the cattle were raised and used as draft animals. This is further supported by pathological evidence prominent on the first phalange of cattle from Tel Tsaf. Using De Cupere, et al. (2000) as a model, Hill (2011) identified that 52% of the cattle’s first phalanges demonstrated some degree of pathology consistent with use as draught animals. In addition, the prominence of silos at the site and the need for large-scale storage may be the result of the use of draft animals greatly increasing crop yield, making it easier to clear and plant larger areas of land.

157

Pigs represent 30% (by NISP) of the faunal assemblage. Evidence suggests that pigs were domesticated in the Near East by at least 7000 BC (Flannery 1983; Larson, et al. 2005; Rosenberg and Redding 1998; Zeder, et al. 2006). Hill (2011) also compared the average size of pig bone within the Tel Tsaf sample against the mean size of modern wild pigs from Turkey (Payne and Bull 1988). The Tel Tsaf population was significantly smaller, demonstrating diminution, which given that most of the animals lived to adulthood is one of the key indicators of pig domestication. The bone fusion mortality profile at Tel Tsaf indicates that the majority of elements were from animals killed between the ages of one year and 2.5 years. Only 17% of elements were from pigs that survived to the age of 3.5 years, likely kept for reproductive purposes (Hill 2011). This kill pattern supports the idea of raising pigs for meat and hides.

The faunal evidence presents a picture of small-scale exploitation of animals for numerous resources. Sheep and goat were managed to maintain herd size, with no evidence of intensive exploitation for either milk or wool. Cattle appear to be maintained for meat and milk but with good evidence of specialized exploitation as draft animals. As with sheep/goat, the animals do not appear to be intensively raised for one specific purpose, but rather maintained to maximize small-scale use for a number of purposes. Lastly, pigs were raised primarily for meat.

Beyond examining what species were being exploited at Tel Tsaf, and why, we can now begin to look at where animals were kept on-site. It seems likely that sheep and goat were largely kept off site, as it is an effective way to maintain a large flock. If this was the case, the circular structures on site may be related to the use of animals for secondary products, such as milk, yogurt and cheese or wool. Ethnographic observation in the region supports this hypothesis with numerous examples of small circular structures being associated with modern pastoralist communities (Banning and Köhler-Rollefson 1992; Cribb 1984; Palmer, et al. 2007; Shahack- Gross, et al. 2003). These structures are used to separate young sheep or goat from their mothers, especially when they wish to limit the suckling of infants in order to increase milk yields for exploitation. Alternatively, Brouchier, et al. (1992) describe pens from Sicily that are designed to confine sheep and goats to facilitate easier milking of the animals. The Sicilian pens are more elaborate, with a slab for supporting the milking bucket as well as vertical stones to protect the bucket. While the dimensions and layout at Tel Tsaf are not identical, the enclosed area,

158 complete with step down, may have been used as a milking area, confining a limited number of animals and allowing easier containment during the process.

Beyond ethnographic analogies, it is important to consider alternative explanations for the circular structures, especially to explain the significant proportion of pig bones found at Tel Tsaf. Despite their seemingly single purpose as meat producers, pigs offer a number of advantages and may have been exploited for leather after death. Indeed, pigs are often overlooked in discussions of the secondary products revolution and changing social structures within the prehistoric context (Nelson 1998). They can be used as garbage disposal on site, as they will eat and gain nutrition from almost any food source including bone. This ability makes them relatively cheap to keep on site, as they do not compete directly for grain or meat resources with humans. Pigs are receptive to reproduction year round, have a short four-month gestation period and can produce litters of up to 15 piglets (Zeder 1996). This makes them a highly productive food and fuel/fertilizer source (dung), at very low labour and resource costs.

Archaeologists often assume that pigs were raised at the household level, with the animals co- habiting humans in housing complexes (Diener, et al. 1978; Grigson 2007; Hesse 1990; Hesse and Wapnish 1997; 1996, 2003). There are greater difficulties associated with keeping a drift of pigs in a field. They do not herd as well as sheep, goat or cattle, and can often run amok if left unsupervised. In general, foraging pigs have a large impact on their environment, partially because they eat everything but also because they will dig for food if needed. It is possible to keep and maintain a drift off-site, but it is generally assumed archaeologically that they were kept on site. This seems likely at Tel Tsaf, but I would suggest that, rather than being kept within a dwelling complex, they were housed in the circular buildings in Building Complex II.

The roasting pits within the courtyard of Building Complex I demonstrate that food preparation and consumption took place within the open courtyard. Having pigs rummaging within the same space would not have been ideal. If the structures in Building Complex II did represent storage for pigs, each pen could have housed one to two sows and their piglets. Alternatively, sows and piglets may have been kept elsewhere on site, and the Building Complex II pens may have housed male pigs or pigs ready for slaughter. There is also evidence that male pigs tend to fight with each other (Zeder 1996), so having multiple small pens would allow the inhabitants to keep

159 them separate. Furthermore, it would be expensive and unnecessary for each individual household to keep a male pig and a female pig, as the male is only required for a brief period of time. The pens in Building Complex II could represent communal male pigs.

The advantages of having the pens on site, nearby, but separate from houses, are many. The pens could have been used as waste disposal, as feeding kitchen refuse to the pigs would have lessened the need for middens. In addition, dung could frequently be collected from the sties, making fuel readily available. Being near housing complexes, such as Building Complex I, the Building Complex II pens would have made waste disposal and fuel gathering an easy task. However, having these areas separated from the actual habitation area would lessen the impact of the smells associated with raising animals in enclosures.

There is debate regarding the use of content and morphological indicators in archaeologically- preserved feces to identify the species of origin (M. G. Canti 1997; Goldberg, Berna, et al. 2009; Rasmussen, et al. 2009; Shahack-Gross 2011). As a response, work has begun to focus on chemical and biomarker indictors in archaeological samples (Bull, et al. 2005; 2002). Recently, L.-M. Shillito, et al. (2011) have demonstrated that Gas Chromatography – Mass Spectrometry can be used to differentiate between omnivore and ruminant feces based on sterol content and between omnivore species based on dominant bile acids. Their work from Çatalhöyük revealed that feces originally thought to be of animal origin were actually from humans. Such studies have tremendous potential and would greatly enhance our understanding of animal storage at Tel Tsaf. With renewed work at the site, it may be possible to pursue this avenue of testing in the future.

4.4 Cooking and Manufacturing

The soil samples from Tel Tsaf present evidence of dung through the excellent preservation of fecal spherulites in micromorphological and bulk samples. Combined with preserved phytoliths, there is good evidence of dung in penning and burning contexts. The abundance of dung spherulites within roasting pits as compared to the general courtyard indicates that dung was being used as a fuel source. There are also examples of charcoal and wood fragments as well as

160 botanical remains from flotation (Graham 2014), so it is likely that both wood and dung were used in these roasting pits.

Dung as fuel is a well-noted phenomenon among several cultures today (Sillar 2000). Domesticated animals provide a renewable supply of animal dung, making it an ideal fuel for cooking and manufacturing. Dried dung burns at a steady rate and produces relatively constant temperatures over a long period, making it useful for both cooking and ceramic manufacture (Shahack-Gross 2011; Sillar 2000). Open-air dung-fueled fires can reach temperatures as high as 630°C and maintain at least 400°C for hours (Shahack-Gross 2011). It has also been suggested that enclosed dung fires could reach temperatures in excess of those indicated above (Matthews 2010).

The best evidence for in situ use of dung as fuel comes from the roasting pits of Building Complex I, Phase 3. These pits contained high concentrations of dark sediment and ash, as well as some animal bone. Pig bone represents almost 51% by NISP of bone recovered from these contexts. This is significant, as pig bone only makes up about 30% (NISP) of the total faunal assemblage at the site. This fits well with pigs being raised as a primary food source. There is also evidence of sheep/goat (24% by NISP) and cattle (18% by NISP) bone in the roasting pits. This suggests that these animals were also cooked on site. The co-mingling of bones of different species within the pits suggests that the pits were used more than once. Unfortunately, all attempts to take micromorphology samples from the pit interiors were unsuccessful. The loose matrix and high ash content resulted in blocks crumbling prior to or during their extraction. It is therefore not possible to assess the potential re-use of pits via microstratigraphy.

During excavation, 11 roasting pits were identified in Building Complex I, Phase 3. Two of these pits, 349 and 261, were located in mudbrick enclosures and all others were bean-shaped pits dug into the ground in the courtyard. These features were not identified in any other building complex or phase. While the number of pits and their capacity for cooking are significant, as with the silos, we need to consider a number of factors. It is possible that the pits were used simultaneously for a single roasting event. If this was the case, it would indeed represent an impressive feast, as suggested by Ben-Shlomo et al. (2009). It is more likely that the pits were used individually, perhaps each representing a single roasting event. In addition, if

161 the pits were re-used then one pit may represent a weekly, monthly or even annual cooking cycle. This is one potential explanation for the variation in species found within some of the roasting pits. I believe it is reasonable to assume these pits represent cooking facilities for the inhabitants of Building I used on a rotating basis, rather than simultaneously for large-scale events.

The use of animals for secondary products has been discussed above. Kill patterns indicate herd management with meat exploitation being the likely primary goal. There is no evidence of intense, specialized breeding of sheep/goat for secondary products such as dairy and wool. It can be, however, be assumed that some secondary products were being used by the inhabitants of Tel Tsaf, but perhaps on a small scale. As discussed in Chapter 2.8.4, three fragments of limonite associated with ash and slag were uncovered in Building Complex IV, Phase 4. These could be by-products of dying wool. Limonite, an iron ore, continues to be used in dyes today. The ashy area where these items were found is not contained within a specific hearth or oven, but rather appears to be a surface spread. This may mean that the ash was dumped in this area and the actual processing took place elsewhere. The bulk sample slides from this area showed an abundance of spherulites and phytoliths, strongly suggesting that dung fueled fires here, for whatever purposes they may have been used.

4.5 Architectural changes between phases

4.5.1 Building Complex I – Continuity

It appears that Building Complex I changed relatively little in either form or function during the excavated sequence of occupation at Tel Tsaf (figure 4.4). The rectangular room and silos remain constant forms of architecture, although there is evidence of renovation or repair to Room 70 between Phases 4 and 3. The most visible change is in the number and arrangement of silos. Phase 4 appears to have five silos associated with it and the Phase 3 courtyard contains four silos. It is possible that the phasing of these features is not completely accurate, especially as we know that the silos were not all constructed at the same time. Silo 633 lies beneath the eastern portion of Silo 53, indicating that it definitely dates to a latersub- phase 4. Despite the apparent

162 decrease in silo numbers between Phase 4 and Phase 3, there is an increase in actual surface area. If the silos were each used to store different products, the increase in surface area may indicate intensification in production of some products, necessitating larger storage areas. Alternatively, the change may be related to the introduction of roasting pits as a means of cooking during Phase 3. Larger storage capacity in a smaller number of silos may have taken up less room in the courtyard, leaving more space for cooking in the courtyard.

Figure 4.4: Phases 4 and 3, Building Complex 1. There is relatively little change in the main features between the phases. The complex is still dominated by Room 70 and a series of silos.

Room 70 demonstrates evidence of occupation during the entire use-life of Building Complex I. The four plastered floor levels, separated by fill layers to level the surface, suggests that this building was renovated at least three times. There is also a slight change in the room construction, likely repair or renovation between Phases 4 and 3 (figure 4.5). This is evident in the layout of the mudbricks in the northernmost portion of the room. The upper course (Phase 3) is laid on top of the earlier wall, but the two courses do not directly line up. This suggests that there was fairly extensive re-building between the phases.

163

Figure 4.5: Room 70, Building Complex I. The arrows indicate the Phase 4 and Phase 3 walls. The overall shape and general size of the room remains the same; however, it is clear that there was some modification to the walls during Phase 3.

There is no clear indication of abandonment or major renovation of the floor surfaces within the building, so it is likely the architectural redesign was a repair rather than re-build after a destruction event. The Phase 3 construction appears slightly smaller but overall room area does not change significantly.

The lack of cooking features, such as roasting pits, in Phase 4 suggests that cooking did not take place in the same way during this phase. Alternatively, cooking may have been done in an unexcavated area of the complex. It is also possible that the courtyard of Phase 4 was not excavated down far enough to reveal these features in the earlier level. The micromorphological evidence indicates that there were at least two roughly compressed floor levels in the courtyard. There is also evidence of charcoal, ash and phytolith levels that may have resulted from spreading hearth remnants that would not have been clearly distinguished during excavation. The presence of sheep/goat, pig and cattle bones through all levels does suggest that meat was being prepared and consumed within this area throughout its use.

There is some evidence of cultic or ceremonial practices within Building I. The 2005 excavations uncovered a stash of horn cores in the southern portion of the courtyard. The horn cores all displayed evidence of burning and were mainly from male gazelle (Hill 2011). It is

164 unclear with which phase they are associated but it was likely Phase 4. There was also a human burial (C98) near the horn core stash, adjacent to Silo 74 in the south of the courtyard. Nearby, two stone violin-shaped figurines were also found (Garfinkel et al. 2007, 2009) and these are usually interpreted as symbolic or ritually important objects (Rowen and Golden 2009). A second burial was in Silo 339 (C555). Both bodies were in a flexed position with their heads oriented to the east. The burial from Silo 339 had 1688 ostrich-shell beads laid out in six rows around the waist (Garfinkel et al. 2009). It is likely that the burial (C98) dates to Phase 4 because of its association with the Phase 4 Silo 74. Silo 339 is dated to Phase 3 and the burial most likely dates to the end of the silo’s use. The social implications of these features will be addressed in more detail in the next chapter.

4.5.2 Building Complex IV to Building Complex II - Why the change?

Despite the continuity between phases seen in Building Complex I, the complex directly east changed dramatically during its use. The change in architecture and function is so significant that the complex is labeled Building Complex IV during Phase 4 and Building Complex II during Phase 3 (figure 4.6).

Figure 4.6: Phases 4 and 3, Building Complexes IV and II. The difference in architecture between Building Complex IV and Building Complex II is significant. Phase 4 is dominated by the large Silo 272 with evidence of some potentially rectangular buildings. During Phase 3 two circular buildings are present, as well as two smaller silos.

165

Building Complex II shows an expansion of the courtyard area, as well as the addition of two circular rooms within the complex. Silos 288 and 272 decreased in size, with the total area of about 10 m². The two round rooms, 263 and 230, present evidence of being animal pens as discussed in earlier. The bulk samples from within the courtyard lack the high relative abundance of spherulites and phytoliths found in the room interiors of 263 and 230. This indicates that animals (possibly pigs) being penned here did not roam the courtyard freely. The silos thus may represent storage of feed for the animals. Alternatively, perhaps during Phase 3 they were used to store dung mucked out of the pens. The lack of roasting pits within the courtyard suggests that food was not being cooked within this area. The excavated portion of Building Complex II does not appear to be a residential area especially considering the evidence from Building Complex I. Building Complex I contains the hallmarks of an active residential complex, but Building Complex II lacks many of the key indicators, such as roasting pits and constructed mud-render floors. The difference in architectural form between the two buildings also suggests form is tied to function. A more in-depth discussion of possibilities and implications can be found in the next chapter.

4.6 Tel Tsaf during the Middle Chalcolithic

The excavations at the site have uncovered at least one residential area; Building Complex I. It appears to have been continuously inhabited during Phases 4 and 3. There is ample storage and evidence of cooking areas within this complex, demonstrating that domestic tasks were carried out within the courtyard. The silos likely stored grain, but may also have been used for storage of other products such as dried dung bricks for use as fuel or legumes such as lentils. Room 70 represents the primary habitation structure, with evidence of clean, well-lain plastered floors.

Building Complex II was separated from Building Complex I during Phase 3 by a mudbrick wall with no visible openings. It also had silos for storage but the circular architecture likely acted as animal pens rather than residential areas contrary to Garfinkel et al (2007, 2009). The variety of domestic species represented in the faunal data presents a number of options for the use of these pens. I suggest that the pens were used to house pigs rather than sheep, goat or cattle. These

166 animals were likely confined to the pens themselves and not allowed to roam the courtyard area. The courtyard may have provided space for other domestic tasks, or as a processing area for meat or dung preparation.

Building Complex IV, Phase 4 was beset with at least two destruction events that did not affect Building Complex I. It is unclear whether these events were intentional or accidental, and the actual extent of architecture from Phase 4 is unclear. At least two silos were built during Phase 4 and renovated for the Phase 3 occupation of Building Complex II. There is also evidence of a possible wool-dying industry during Phase 4. It is unclear whether the ash, slag and limonite represent in situ processing activity. It is possible that Building IV was used as a dumping area for the ash following processing elsewhere on site.

An interesting picture of occupation at Tel Tsaf has emerged. Before moving forward with discussion of the social and economic implications of these finds, I offer a caveat to readers. It must be acknowledged that the excavations have not uncovered a single complete building complex. The western boundary wall of Building Complex I remains buried beyond the described excavated area. Similarly, the eastern extends of Building Complexes II and IV has not yet been identified. We can comfortably discuss the evidence as presented, but we do need to remember that crucial data may be missing. The full extent of the courtyards and any additional rooms, silos or features remain unknown. With that in mind, Chapter 5 will explore the potential implications of the finds presented thus far.

167

Chapter 5 Storage, Surplus and Wealth

Chapter 4 presented the evidence for a reconstruction of economy and domestic life at Tel Tsaf during two occupational phases, Phase 3 and Phase 4. Now, I turn to an examination of the socio-economic implications of these reconstructions. As the evidence for Phase 4 at Tel Tsaf is more limited and incomplete compared to Phase 3, I will focus discussion on the Phase 3 occupation and draw in data from Phase 4 where appropriate. I will begin by addressing household archaeology, with special attention to its relevance to my analysis. This will be followed by a discussion of the importance of moving beyond architectural analysis and incorporating microscopic analysis into assessments of socio-economic organization at prehistoric sites. This chapter will close with a discussion of some of the hallmarks of stratified societies or chiefdoms, including surplus, resource control and displays of power, followed by how the evidence at Tel Tsaf fits into such discussions.

5.1 Architecture and Archaeology

There is no denying the importance of architectural analysis in archaeological contexts. Architectural features, where present, provide valuable context for the other material culture excavated from sites. It can also more directly contribute to reconstructions of household size, storage capacities, economic activities and social structures. However, these types of data do not immediately present themselves and must be elucidated from a variety of sources. It may be tempting to identify small buildings as houses and large buildings as public spaces during excavation, but such assumptions lack evidence and may be wrong. Material culture can help shed light on building use in cases where clear patterns of distribution emerge. A hearth within a building combined with a high concentration of domestic ceramics can be a good indicator of a domestic dwelling. But when such clear indicators are lacking, archaeologists need to find alternative methods of analysis with which to frame their discussion. Without additional analytical and theoretical tools, architectural analyses risk becoming a series of unfounded assumptions based on modern expectations or ethnographic analogy.

168

Archaeology has drawn upon social literature that addresses how the built environment influences and reflects behaviour within a society. How humans choose to modify the natural environment through building walls, digging pits or exploiting natural features for their own purpose has strong social implications. In his 1969 book House Form and Culture, Amos Rapoport demonstrated that culture influences house form. Rapoport (1982) in The Meaning of the Built Environment: a nonverbal communication approach went beyond this to outline a new approach to understanding human interaction within built environments. Rapoport suggests that the built environment presents visual cues that inform users of acceptable behaviour by both constraining and enabling behaviour through configuration of walls, doors and other features. The location of such features influences communication and is thus socially meaningful. The features provide cues that inform behaviour seen as appropriate in a given situation (1982: 57). The end result is that the built environment acts on behaviour by providing cues that people use to judge the social context and act accordingly. In other words, “it is the social situation that influences people’s behaviour, but it is the physical environment that provides the cues” (1982: 57).

Such approaches outlined the importance of the built environment in providing cues for behaviour. The social aspects of this approach were further elaborated by Pierre Bourdieu, who outlined Practice Theory and the concept of habitus in 1977, describing the latter as a unique schema of unconsciously internalized dispositions (Bourdieu 1977). Ultimately, material culture and house form, through non-verbal communication, are both structured and structuring (Blanton 1994; Dornan 2002; Verhoeven 1999). These elements determine how the world is perceived, and how people respond to the visual clues around them. They create and enforce specific social practices, unconsciously, upon individuals acting within the built environment (Bourdieu 1990). Giddens's (1984) theory of structuration presents a similar idea of the constraining and enabling natures of social structure. But unlike Bourdieu, Giddens suggests that individual response is not always determined by unconsciously internalised structures. Instead social structures, though manifest and enforced by the built environment, are mutable and influenced by creativity and innovation. Individuals may reflect upon the content and meaning of the social actions enforced by their environment.

169

Archaeological sites that present good preservation of architecture with internal divisions lend themselves well to such analyses. These approaches are further aided by good records of artifact distribution and clear clusters of material culture types within specific areas. One such study was conducted on the “Burnt Village” level of Tell Sabi Abyad, a Late Neolithic site in northern Syria (Verhoeven 1999). In this well-preserved level, artifact distributions were carefully recorded and mapped by the excavators. Combining the architectural data with the material culture distribution allowed the excavators to identify activity, storage and discard areas across the site. Using this data to identify the relationships between architecture, access and use, they were able to instigate cogent discussions of what life may have been like during the village’s occupation (Verhoeven 1999).

The Burnt Village level lent itself particularly well to this type of analysis because it represented the remains of a village destroyed by accidental fire. This meant that many artifacts were found more or less in situ. Most sites do not have such clear preservation, making it difficult to perform such detailed analysis. Tel Tsaf is one such site. There are two destruction events that appear to have taken place, but neither is clearly defined, nor have the levels directly beneath them been extensively investigated. Consequently, further methods of analysis are needed in order to understand the significance of the architectural layout at Tel Tsaf and its social implications. Banning (2010) suggests that Space Syntax and Access Analysis can provide such a framework and are particularly appropriate for discussion of some prehistoric sites that meet certain prerequisites. He argues that there is not a 1:1 relationship between space and social action but there are some connections that we can look to identify.

Access Analysis and Space Syntax provide a quantitative approach to questions of who has access to what spaces across a site. Hillier and Hansen laid out these related, but separate approaches in The Social Logic of Space (1984). Access Analysis relies on topographical relationships both within buildings (Gamma Analysis) and between open spaces and lines of sight across settlements (Alpha Analysis) (Banning 2010). This type of analysis does not always lend itself well to many prehistoric sites in the Southern Levant, including Tel Tsaf, due to the lack of sufficient exposure of architectural remains. At Tel Tsaf, internal divisions within buildings are limited or absent and there has been such limited exposure of the village that Alpha

170

Analysis is not viable. Space Syntax, however, provides a more appropriate tool for analysing and giving meaning to architectural features at Tel Tsaf.

Space Syntax relies upon a structuralist approach to architecture, providing a logic or ‘grammar’ to space (Hillier and Hansen 1984; Banning 2010). Hillier and Hansen (1984) proposed eight elementary Syntaxes (Z1 - Z8)(figure 5.1) that can be used to classify various building or site plans. Each Syntax is classified as either distributed (structure functions independent of other units) or non-distributed (one or more units dictate structure of or access to other units). Within each syntax, x, y and o are used to describe the relationships among closed cells (x), groups of closed cells (xx), open space (y), groups of open spaces (yy) and the act of containment (o) (Hillier and Hansen 1984; Banning 2010)(figure 5.1). This approach allows comparison at the building, community or site level when the physical structure of each element may not be uniform. Architectural forms may differ between two contemporary archaeological sites (round vs. rectangular vs. subterranean), but the syntax of each unit may be the same, suggesting that the underlying intentions regarding access to space is similar. In contrast, two sites may present similar rectangular architecture but the Syntax may differ, implying social relations may have been different at each site. Space Syntax can be employed within sites as well, and is a useful tool for sites, such as Tel Tsaf, that contain more than one building type or house layout.

171

Figure 5.1: Hillier and Hansen’s (1984) eight elementary syntaxes (after Banning 2010). This schematic employs archaeological examples of each of the syntaxes: Byblos (Z1, Z5), Çatalhöyük (Z2), Tel Tsaf (Z4, Z6), Sha’ar Hagolan (Z6), Greek cities, e.g., Olynthus (Z7) and Maasai kraals (Z8). 5.1.1 Building Complex I

The layout of Building Complex I includes a large, bounded courtyard containing one room and multiple silos (figure 5.2). We have already hypothesized that the rectangular Room 70 represents a living space, with several successive mud-rendered floors. The entrance for this broad room appears to be located along the western wall. The northern and western courtyard walls of Building Complex I have not been excavated and, given the lack of any other entrance into the courtyard from outside the complex, it can be assumed that it was entered from one of those directions, likely the west.

172

Figure 5.2: Building Complex I. The Room 70 entrance is circled and potential north or west courtyard entrances are indicated by arrows (after Garfinkel, et al. 2007).

Using Hillier and Hansen’s (1984) syntax classification, Building Complex I falls within the z4 or concentric group. This group is marked by a closed cell within a contained area which is also closed. The location of Room 70 within a walled courtyard strongly suggests its inhabitants’ ownership of all of the features within Building Complex I. The courtyard wall separates Room 70 and the silos from any other households or public space. This stands in contrast to more open site layouts without bounded walls, such as the Golan sites (Epstein 1998) and Munhata 2a (Garfinkel 1992a), and differs from layouts where room walls themselves form courtyards at

173

Teleilat Ghassul (A. Mallon, et al. 1934) and Tel Te’o (Eisenberg, et al. 2001). Despite the containment of the courtyard and Room 70, access to both features appears to be direct and unobstructed. If the courtyard opens on the west, then people arriving would have had direct access to the western entrance of the main feature, Room 70. This provides occupants and guests with easy access to the main room. It also means that guests would not end up walking throughout the courtyard before being greeted or acknowledged by the inhabitants of Building Complex I. There are no walls or other barriers within the main courtyard, but the majority of the silos and roasting pits are east of Room 70. There are no obvious restrictions to the main courtyard, but access to silos and most roasting pits is limited by the placement Room 70. The lack of internal divisions within Room 70 also suggests that controlling access to specific areas within the room was not a priority.

The lack of internal divisions within Room 70 suggests that Building Complex I was inhabited by a nuclear family. A number of ethnographic studies of un-walled, pre-modern settlements in southwest Asia has led to a suggested average family unit of 5-6 people as the base line for population estimates for archaeological houses (Finkelstein 1990; Kramer 1982; Watson 1979). The total living area of Room 70 is about 50 m2, which is more than large enough to house a nuclear family of that size comfortably. Taking the large courtyard and significant storage capacity into account, it is tempting to suggest a larger household unit. It is possible that an extended family inhabited Building Complex I. However, lack of evidence of hearths within Room 70, combined with the lack of internal walls, suggests that a small, close household unit inhabited the area. Extended families are often indicated by demarking space, and it is easy to build additions onto rectangular buildings should a family become too large for an existing building (Flannery 1972). Another possible indicator of multiple generations or families sharing a single house structure is the presence of multiple hearths (Flannery 2002; Garfinkel 2006). However, neither extra rooms nor multiple hearths necessarily dictate extended families, as it is possible for families to share storage facilities but eat meals separately or, vice-versa, store food separately but eat together. However, multiple rooms and hearths can provide archaeologists with some indication that such groups may have inhabited a building. Building Complex I presents evidence for multiple storage units and roasting pits so it is possible that these were each ‘owned’ by members of an extended family group dwelling within the complex. However, given

174 the lack of other evidence, it is a reasonable assumption that a nuclear family inhabited Building Complex I. While it is possible that the household extended beyond Building Complex I, with members seasonally living off-site, such speculation gains no specific support from the available evidence. Based on the above evidence, a picture of Building Complex I as home to a nuclear family partaking in household-level production and consumption emerges as the most plausible scenario.

5.1.2 Building Complex II

The occupation of Building Complex II, Phase 3 presents a syntax that differs slightly from that of Building Complex I. The presence of two internal rooms changes the classification to z6 or Estate. This grouping is marked by a bounded exterior (the courtyard wall) containing at least two interior bounded rooms (Rooms 230 and 263)(figure 5.3). As with Building Complex I, the actual entrance to the courtyard is unknown as the eastern and southern portions were not clearly defined during excavation. The entrances to both Rooms 230 and 263 are oriented to the north. Whether we choose to assume a courtyard entrance to the east or south, there is no direct access to the rooms upon entering the courtyard. People who entered the complex would be required to walk at least part-way around the structures in order to enter them. The occupants of Rooms 230 and 263 would not immediately see or become aware of people who entered Building Complex II. Furthermore, a stranger arriving would not have any immediately visible cues regarding where to enter the rooms. This may represent a level of privacy not present in the exposed portion of Building Complex I. Alternatively, if the rooms were used exclusively to pen animals, indirect alignment of doorways would delay escape from the courtyard should an animal accidentally escape from one of the circular enclosures.

175

Figure 5.3: Building Complex II. The entrances to rooms 230 and 263 are circled and the potential east and south courtyard entrances are indicated with arrows (after Garfinkel, et al. 2007).

Garfinkel, et al. (2007, 2009) and Ben-Shlomo, et al. (2010) argue that Building Complex II represents a second habitation area. They attribute the difference in architecture to economic variation, suggesting that wealth was not distributed equally at Tel Tsaf. Garfinkel, et al. (2009) also acknowledge that the potential storage capacity for Building Complex II exceeds the amount of grain required to feed the nuclear family he assumes inhabits Building Complex II during Phase 3. This creates a model of economic variation in which surplus is still present even in the less economically wealthy household, indicating that surplus does not automatically equate to wealth.

176

Variation in grain storage capacity between building complexes can be a strong indicator of economic variation, but at Tel Tsaf we are uncertain if all silos were used at the same time or for the same thing. Different architectural features could also be an indication of varying household structure. The presence of two small rooms rather than one large room in Building Complex II could be an indication of an extended family household. This could also provide evidence of social and economic variation at Tel Tsaf during the Middle Chalcolithic. However, the silo storage capacity and household structure do not stand in isolation when comparing Building Complex I and Building Complex II. Significant physical and functional variations between the two structures suggest that the layout of each building was based on variation in function rather than variation in economic or social standing (table 5.1). Building Complex I Building Complex II Phases Occupied 4 and 3 3 only Syntax Classification z4 (Concentric) z6 (Estate) Room Architecture Rectangular Circular Household Structure Nuclear Nuclear? or Extended? Building Complex Entrance West (potentially Northwest) East or South Orientation Main Room Entrance West North Orientation Relationship between Building direct access indirect access and Room Entrance Orientation Silo Size (total surface area)(m²) 13.85 / 25.60 10.85 (Phase 4 / Phase 3) (Phase 3) Silo re-use silos retired and new silos Building IV Phase 4 silos built during life of Building renovated/re-used, no new Phase 3 constructions Food Preparation Roasting pits/brick-lined pits no roasting pits, no brick lined pits Burials 2 adult 2 infant Dominant Microfacies MF 1a, MF 1b, MF 2 MF 3a, MF 3b

Table 5.1: Differences between Building Complex I and Building Complex II.

Building Complex II does not have evidence of the roasting pits or ovens that were abundant in Building Complex I. This means that either cooking was performed in a different manner, or that it was done in an area of the courtyard not yet excavated or that there was no cooking done at all. The renovation and re-use of silos may indicate that sanitation was not as important as

177 within Building Complex I, where silos appear to be retired and new ones built within the building complex’s life. The lack of adult burials, but presence of infant burials, also demonstrates differential use of the building complex.

5.1.3 Interpreting the Evidence

Based on architectural form alone, Garfinkel, et al. (2007a, 2009; also see Ben-Shlomo, et al. 2009) suggest that the combination of circular and rectangular structures at Tel Tsaf bears some resemblance to the structures found at Halaf sites in Mesopotamia and northern Syria. In the northern Levant, there is evidence of both rectangular and circular architecture and Garfinkel, et al. (2007a, 2009) specifically cite examples from Tel Sabi Abyad, Arpachiya, Tepe Gawra, Yarim Tepe and Tell Halaf (Akkermans and Verhoeven 1995; Aurenche 1981; Merpert and Munchaev 1987; Verhoeven and Kranendonk 1996). Tel Sabi Abyad is a particularly good example as the late Halaf phase has been well published, with a particularly in-depth analysis of the buildings from the Burnt Village (Verhoeven 1999).

The first major concern that arises from this comparison is the chronological gap between the Halaf sites and Tel Tsaf. For example, the occupation at Tell Sabi Abyad dates between about 6860 and 5760 cal BC (Verhoeven 1999: 4-5), placing it at least 500 years earlier than occupation at Tel Tsaf. Despite the chronological discrepancies, Garfinkel, et al. (2009) go so far as to suggest that Halafian influence or migration to the southern Levant may have contributed to the site plan of Tel Tsaf. The extent of this influence of migration is not made clear, nor will I speculate on how much of a factor the authors think it played in construction and daily life at Tel Tsaf. However, this suggestion represents some of the dangers associated with a purely architecturally-based interpretation of an archaeological site. The immediate analogy between Halafian architecture and that of Tel Tsaf is not unreasonable, as both represent an intriguing mix of two distinct structural forms. However, given the prevalence of broad rooms within the southern Levant during the Late Neolithic, Wadi Rabah and entire Chalcolithic (see chapter 2), the presence of such a structure at Tel Tsaf does not represent a change from local tradition. Circular architecture is not completely unheard of within the region during later prehistory with examples from the Yarmoukian level at Munhatta (1992, 1993b) and ‘Ain Ghazal (Rollefson 1997; Rollefson and Kafafi 1994; Rollefson and Simmons 1987).

178

Further issues with the suggestion of Halafian influences arise when comparing the size, structure and layout of buildings at such sites as Tell Sabi Abyad. The rounded structures, or tholoi, are between 90 m² and 120 m² in area (Verhoeven 1999: 25), making them much larger than the approximately are of 50 m² buildings at Tel Tsaf (Garfinkel, et al. 2007a: 13). At Tell Sabi Abyad, Yarim Tepe and many Halaf sites, both tholoi and rectangular buildings are divided into cells, either creating raised platforms or representing storage chambers (Verhoeven 1999; Verhoeven and Kranendonk 1996). Tel Tsaf lacks these internal divisions. Furthermore, the rectangular buildings at Tell Sabi Abyad are located within the same open space as tholoi, often being directly adjacent to them. At Tel Tsaf, by contrast, there is a very clear and distinct demarcation of space by containment walls around courtyards separating circular and rectangular architectural features, at least insofar as excavated exposures allow us to infer.

A second but related possible explanation for differences in building complex form at Tel Tsaf, as Garfinkel, et al. (2007) hint, is that the complex owners had different regional or cultural backgrounds, again introducing the idea of migration from the North. Garfinkel, et al. (2007) imply that the inhabitants of Building Complex II may have been Ubaid immigrants. This idea is not fully developed in the published literature, and appears to be based on the speculation that a few decorated sherds found at Tel Tsaf are Ubaid in origin. The classification of the sherds as Ubaid is based on the design on the fragments and no ceramic analysis has been reported to determine whether the sherds are of local or foreign origin. The authors do also suggest that presence of the sherds may be due to trade networks with the north (Garfinkel, et al. 2007, 2009, 2014). However, we need not look to external influences to explain the variation in architecture between Building Complex I and Building Complex II. It seems more likely that the differences exhibited in the excavated portion of Building Complex II are due to it having been used differently than Building Complex I. I will further explore this idea in the following sections.

5.2 Surplus and Wealth

Surplus occurs whenever production output of any good surpasses the immediate needs of the production unit. In the archaeology of pre-urban societies, surplus is most frequently discussed in relation to food production, specifically agriculture. The presence of surplus suggests that the labour force and technology of a population has reached a point that allows food to be reserved

179 for later use or to be exchanged with people who do not produce their own (Childe 1950; Sahlins 1958). This represents a significant change in survival strategies from household subsistence production. When a household unit is primarily focused on ensuring that it produces enough food for its survival, there is limited time for full-time production of other goods (Childe 1950). Craft specialists may have existed prior to the availability of surplus, but until enough food could be secured to feed a portion of a population that is not actively contributing to subsistence production, no household or community could commit to full-time non-subsistence specialisation (Childe 1950). However, once food surplus existed, a portion of the population could be freed from subsistence production to be involved in other, non-food-producing tasks, providing a means for redistributing food to support craft specialists.

Traditionally, anthropologists have viewed surplus as a key indicator of the accumulation of wealth and the beginnings of social and economic variation (Childe 1950; Cowgill 1975; Flannery 1969, 1972; Sahlins 1958). This body of literature makes a strong case for such scenarios, but it is important to recognise the distinctions between storage, surplus and wealth. Archaeologically storage can be identified through features such as large jars, bins, pits and silos. These features demonstrate that items were being kept for some time, whether it be a week or a year. As discussed, surplus occurs when the amount of a particular good exceeds the basic short- and long-term needs of a social unit. It does not, however, in itself represent wealth. In order for surplus to become wealth, it must represent a surplus of goods needed by others. In must be desired by others in order to be used to obtain further items that go beyond basic subsistence needs.

The relationship among storage, surplus and wealth is not linear, nor is it limited to material goods. The social aspect of economics plays a significant role in the development of hierarchy within societies. Indeed, economic systems are not driven by solely economic motives, but rather by issues related to social standing, prestige and power (Polanyi 1944). In pre-market societies production and redistribution is managed by reciprocity rather than the quest for material gain or standardisation by a central authority (Polanyi 1944). Within the anthropological literature related to economics, the idea of obligation is dominant and is integral to a social and economic community based on reciprocity. Through the navigation of both

180 material and social aspects of storage and wealth it becomes clear that accumulation and exchange of courtesies, assistance or gifts is aimed at creating a debt owed, as much as gaining physical wealth (Mauss 1967).

The obligation to return gifts or favours comes with the penalty of loss of authority or wealth within many modern and historic societies. Mauss (1967) noted that among Samoan and Maori populations there were well-regulated expectations regarding the giving and returning of gifts, all of which are linked to social standing. In both cases, failure to return gifts or social obligation in the correct manner resulted in a loss of position within that society. In the case of populations in the Andaman Islands, such trade and barter went beyond basic needs, as self-sufficient families still engaged in exchange in a friendly, competitive fashion (Mauss 1967). Attempts to out-do each other through gift-giving and social obligation established and maintained social standings.

In such societies, the aim is prestige and potential power (social wealth) rather than simple material wealth. This concept is clearly highlighted in the competitive feasting of the Potlatch on the northwest coast of North America. The potlatch is a regional feast with complex social meanings. It involves a gift exchange aimed at gaining prestige through expenditure of material goods in a way that creates material and social debt. Participants in the potlatch cannot refuse gifts, ensuring everyone must participate. Some individuals will go so far as to redistribute all their material possessions in order to gain power over debtors (Mauss 1967). In this way, power and prestige have a much higher value than possession of physical goods, but material wealth is required in order to gain social wealth. As a result, the cycle of gift-giving, debt creation and repayment becomes highly competitive.

The interrelationship among physical storage, surplus and wealth, as well as social storage and wealth is very complex, as demonstrated in figure 5.4. The ethnographic examples above highlight the importance of social wealth and obligation in gaining prestige and power. However, the path to social wealth is not solely through material wealth. Having conspicuous storage or surplus can also lead to prestige within a community. Sharing material surplus without immediate remuneration can create a social surplus in the form of obligation, which in turn leads to power over an individual until the debt is repaid. Material storage can lead directly

181 to material wealth if one gains more than one loses in exchange. For example, one could potentially exchange part of subsistence storage of grain for a subsistence amount of alternative food products plus prestige items if the ‘buyer’ is desperate for grain. Alternatively one could acquire material wealth prior to having long-term storage and then be required to build new facilities. These examples are in no way exhaustive, but serve to demonstrate that the relationship among storage, surplus and wealth in both social and economic contexts is not always linear and can be quite complex.

Figure 5.4: The relationship among storage, surplus and wealth. The red arrows represent a linear approach to the concepts in which social storage and wealth are viewed independently of the material counterparts. The purple arrows demonstrate other potential relationships among all the constituents.

The above discussion of material and social storage and wealth are based on ethnographic records and modern studies. Attempting to determine the roles of storage, surplus and wealth in a prehistoric context presents some challenges. In prehistoric farming villages, evidence of grain storage in the form of storage pits, jars or silos encourages speculation regarding the amount of storage available, the surplus represented and how it was shared or controlled (Banning 2010; Flannery 1969; Garfinkel, et al. 2009; Kuijt and Finlayson 2009). Such evidence often forms the basis for discussions regarding economic structure in early villages, especially when the storage

182 capacity is exceptionally large (Banning 2010; Garfinkel, et al. 2009; Kuijt and Finlayson 2009). Sahlins observed that “the degree of stratification is directly related to the surplus output of food producers” (1958: 5). His comments were based on ethnographic observation in Polynesia but the idea has more generally influenced the archaeological study of surplus. It is reasonable, though not conclusive, to assume that large storage capacity suggests surplus, and it is also reasonable to examine how and where such surplus is stored. Indeed, later in this chapter I will do just that for Tel Tsaf.

As mentioned, evidence of exceptionally large storage capacities is often assumed to mean surplus in the archaeological context. However, the scale of surplus is often difficult to establish. In archaeological usage, surplus is often an amount of product that is above and beyond the basic short-term needs of the producers for either consumption by humans and animals or storing seed for future planting. Beyond establishing exactly how much is needed for subsistence, distinguishing whether facilities were used simply to store goods for personal use another day or whether they were used to store items in excess of long-term personal need is difficult to determine. For example, large storage jars could be used to hold grain that will be used to make bread for a family over the coming months or they could be storing the leftover grain that the family does not need and may therefore exchange for something else.

One could argue that large storage features, such as silos, represent substantial investment in resources and time to construct and may therefore represent investment in long-term storage for a consistent surplus. However, such an investment would not be unreasonable if the same structure was to be used for repeated, short-term storage. It should also be considered that not all silos, pits and jars may have held grain or even the same grain. It is plausible that more than one type of grain may have been stored, such as barley and wheat. Other resources, such as lentils, chaff or dung pellets for fuel, may also have been stored by the same family in separate jars, pits or silos at the same time, necessitating multiple storage areas. One should also be consider that some of the stored foodstuffs were for maintaining animals kept within the village. Furthermore, large storage jars may have been used to hold fresh water, milk, yoghurt, honey or other liquids.

If these alternatives are considered, even large storage capacities may not represent a large surplus of grain. There may not be any surplus at all, but rather a comfortable supply of a

183 number of daily-use products. In some cases, these supplies may be topped up daily, weekly or monthly. Being available year-round, items such as dung pellets, water or milk could easily be replenished frequently. Unfortunately, in most cases we have evidence of storage but we do not have evidence of exactly what or how much was stored. Further evidence of these uses is beginning to emerge as residue analysis is becoming more frequent in archaeology (Copley, Berstan, et al. 2005; 2005; Craig, et al. 2005; Evershed, et al. 2008; Gregg, et al. 2009) and botanical analysis can also potentially shed light on plants present within a storage area (Fairbairn, et al. 2007; Graham 2014; Hald and Charles 2008; Twiss, et al. 2009).

One final issue to address before discussing surplus at Tel Tsaf is that not just crops, but also animals can be surplus. As noted, we can often infer surplus in products such as grain by the need for large storage facilities. Despite difficulties in establishing the nature and extent of surplus in these contexts it seems likely that surplus existed when storage facilities greatly exceed the needs of the households with which they are associated. However, with animals it is much more difficult to establish the point at which they become an indicator of surplus. It is likely that cattle and pigs would be kept physically close to the house structure (Epstein 1998), but sheep and goats would most likely be herded off-site. When animals are kept primarily for food, then animals themselves can be viewed as storage on the hoof (Marshall 1986). In such cases, it is possible that surplus is actually undesirable as additional animals would compete with people for food. This is most likely to occur in populations that have limited resources (Marshall 1986).

The main source of evidence of personal ownership of animals one would expect in the archaeological record would be animal pens. Stabling or penning areas have been identified within urban contexts (Albert, et al. 2008; Shahack-Gross, et al. 2005), but such evidence is less common in pre-urban contexts in the southern Levant.

Despite these difficulties, the importance of animals as wealth has been noted in discussions of the southern Levantine Prehistory, but it is also acknowledged that ‘storage’ is difficult to identify (Banning 2010; Barker 2012; Twiss 2008). In the case of feasting, analysis is based on number of animals represented in the faunal collection within the context of consumption or disposal. This informs us about where animals were eaten and bones thrown away but does not

184 provide direct evidence on the location of animal pens on site, and means that ownership cannot be identified to individuals or households. These ideas will be explored more fully in section 5.5.

5.2.1 Tel Tsaf Silos Re-visited

Having addressed concerns regarding surplus in general, I will try to address these issues as we take a closer look at the silos at Tel Tsaf. The number and size of silos strongly points to surplus at the site. The inhabitants of Buildings I, II and IV clearly had a need to construct large silos and these silos were used during both Phases 3 and 4. Garfinkel, et al. (2009) make a case for the silos being built to a height of 2 m and from there calculate an approximate maximum capacity of each silo at Tel Tsaf (table 5.2). Silo Building Phase Diameter Area Estimated Number (Hubbard) (m) (m²) capacity (m³) 286 I 3 3.00 7.05 14.10 339 I 3 3.00 7.05 14.10 415 I 3 2.50 4.90 9.80 53 I 3 2.90 6.60 13.20 288 II 3 2.20 3.80 7.60 272 II 3 3.00 7.05 14.10 415* I 4 2.50 4.90 9.80 633 I 4 1.60 2.00 4.00 66 I 4 1.90 2.80** 5.70** 171 I 4 2.20 3.80 7.60 74 I 4 2.00 3.15 6.30 288* IV 4 1.80 2.55 5.10 272* IV 4 4.00 12.60 25.20

Table 5.2: This table notes the number and dimensions of the silos from Phases 3 and 4 at Tel Tsaf. The estimated capacity is based on the assumption that each silo was 2 m in height (data from Garfinkel, et al. 2009).

*These silos were renamed in Garfinkel, et al. (2009) during Phase 4 as follows: 415 = 603; 288 = 565; 272 = 568. The dimensions vary between phases, but, since the location is unchanged, I have referred to these silos by their Phase 3 designations throughout this dissertation.

** These numbers have been changed from Garfinkel, et al. (2009) as the calculations provided did not correspond with the diameter as published. The above figures are correct based on a diameter of 1.90 m.

It is important to address a few issues with this approach before moving forward. The assumed height of 2 m is a reasonable estimate when compared to the preserved super structures of the Bir

185 el-Abd and Tel Gemme silos. The Bir el-Abd examples have an average diameter of 3.8 m, around 1m larger than most at Tel Tsaf, and the domed shape superstructures are preserved to c. 1.8 m in height (Oren 1973). The preservation does not represent the full height of the structures, so it is likely that the silo heights exceeded 2 m. At Tel Gemme, the dome-shaped portion of the superstructure does not begin until 1.5 m above the base platform and 6 courses of bricks are preserved (van Beek 1972). The full height of the dome is not preserved, but it can once more be assumed the completed structure would stand more than 2 m in height, even if the cylindrical portion was only 1.5m high.

Due to the preservation of these features at Tel Tsaf, we are unable to determine whether all the silos were the same height. Silos with larger bases may also have been taller, meaning there may not be a linear relationship between area and volume. It is also important to keep in mind that, while we can tie certain silos to particular phases, we do not know whether all these silos were used at the same time. They may have been used on a seasonal basis, or even in a rotating system through each phase. The presence of four silos does not necessarily mean that all four were full at any given time. Lastly, while we do have evidence of charred seeds and phytoliths from these structures, we cannot be sure every silo was used to store the same product. They may have been used to store any number of items including (but not limited to) barley, wheat, pulses, chaff or even dung. Each of these would produce similar archaeological evidence, particularly if the silos were cleaned out prior to reuse or abandonment, which is where micro-level evidence becomes important. It is reasonable to assume that silos often fulfilled a role in grain storage but it is not the only explanation for silo use. That being said, the estimates provided by Garfinkel, et al. (2009) do provide a good starting point for discussions regarding the significance of the silos at Tel Tsaf.

Expanding on the potential storage capacity, Garfinkel, et al. (2009) further estimate the mass of grain that could be stored at Tel Tsaf (table 5.3).

186

Estimated Estimated mass of Silo Building Phase Area capacity grain Number Complex (Hubbard) (m²) (m³)* (metric tons)** 286 I 3 7.05 14.10 9.87 339 I 3 7.05 14.10 9.87 415 I 3 4.90 9.80 6.86 53 I 3 6.60 13.20 9.24 288 II 3 3.80 7.60 5.32 272 II 3 7.05 14.10 9.87 Building Complex I - total est. mass of grain (tons) 35.84 Building Complex II - total est. mass of grain (tons) 15.19 *calculated based on 2m silo height **based on 700kg per m³

Table 5.3: Grain storage capacity in metric tons for all Phase 3 silos in Building Complexes I and II.

The numbers presented are based on an estimated density of 700kg/m³. Garfinkel, et al. (2009) do not cite the source of this estimate, but a review of modern figures from three sources suggest that 700kg/m³ is reasonable as an average density of barley and wheat grains (table 5.4).

Walker (2011) A.G.Contractors (2013) LeonCookSilos (2013)

Wheat 600 kg/m³ 620 kg/m³ 620 kg/m³

Barley 780-800 kg/m³ 770 kg/m³ 750 kg/m³

Table 5.4: Modern grain densities per cubic metre. Some fluctuation in grain weight can be expected between modern and prehistoric grains; however, these figures provide a relative guideline for use here.

The potential volume of grain storage presented by Garfinkel, et al. (2009) demonstrates a very significant amount of storage was available to the inhabitants of Tel Tsaf. They further go on to suggest, that assuming it takes about 200 kg of cereal to feed a single person for year (Abramowitch and Gelfet 1944; Broshi 1979; Crawford 1971; Hole 1991; Jardé 1928; Mazar 2001), the silos would not only meet, but greatly exceed, the demands of a single household. Extrapolating the numbers he has provided, the full extent of that potential surplus becomes apparent. Based on the estimated grain capacity in tons and the figure of 200 kg per year per

187 person, Building Complex I could potentially store enough grain for 182 people. Using the same estimates, Building Complex II could have stored enough grain to feed 73 people for a year. As noted above, Building Complex I likely housed a nuclear family. If Building Complex II is considered a house as well, it may have housed an extended family, with each generation having its own circular building. Even in this scenario, the amount of grain storage available far exceeds even the needs of an extended family household. This exercise demonstrates that the silos at Tel Tsaf likely not only represent storage of surplus but potentially a vast surplus of cereal, assuming all silos were filled with grain at the same time.

Having established that the inhabitants of Tel Tsaf had the capacity to store large quantities of surplus, let us revisit Garfinkel, et al. (2009) and their estimates of land required to produce enough grain to fill the silos. The authors suggest that in order to fill the silos of Building I alone, some form of hierarchical administration was likely at play. They go so far as to introduce the idea of potential landlord-tenant relationships or tax collection by elites. Garfinkel, et al. (2009) base this idea on yield calculations of 500 kg of grain per hectare (no source provided). Thus 40 ha of land would be needed to provide 20 tons of grain and a nuclear family would not be able to manage the land required to fill the Building Complex I granaries on its own (Garfinkel, et al. 2009: 232). The reasoning behind using 20 tons of grain is not explicitly outlined by Garfinkel, et al. but it can be assumed it is a rough average of the potential capacities of Building Complex I and Building Complex II. However, based on the estimate of 200 kg per person, even a large nuclear family of 6 people would only need 1.2 metric tons of grain per year, which would only require 2.4 hectares based on their estimated yields.

The amount of land required to support a nuclear family can be further revised by revisiting the estimated yield of 500 kg per hectare that Garfinkel, et al. (2009) use. Their estimate differs from that put forth by Araus, et al. (1999; 2001), who examined potential grain yields in their study of archaeological data from Tell Halula, Syria. The authors first established an exponential relationship between crop yield and carbon isotope discrimination using data from a wide range of modern durum wheat genotypes cultivated in varying environmental conditions in northern Syria and northeast Spain. The same carbon isotope discrimination analysis was performed on charred grains from Tell Halula (9550 – 8630 cal BP) and the modern relationship was applied to

188 the data before adjusting the results to take into account differences in CO₂ levels and between modern and ancient harvest indices. Their results indicate that ancient grain yield estimates based on attempted reproduction of ancient farming or traditional methods (Castro, et al. 1998; Reynolds 1979) are consistent underestimates. Araus, et al. (1999, 2001) suggest that average yields would have been around 1.56 ± 0.28 metric tons per hectare per year in the region.

This is significantly below modern yields (3.83 Mg ha-1)(Araus, et al. 1999), but it is also nearly three times the estimate that Garfinkel, et al. (2009) provide for Tel Tsaf. The yields from Tell Halula cannot be directly assumed for Tel Tsaf, as irrigation, soil quality, type of grain and other unknown factors need to be taken into account. Even so, it seems that the estimates of Garfinkel, et al. (2009) are low and should be reconsidered. Even if we work with their example of 20 metric tons, the application of Araus, et al.’s yield of 1.56 metric tons per ha suggests the amount of land required is only 12.8 ha. This figure is significantly lower than Garfinkel, et al. (2009) suggest. It is possible that 12.8 ha of land would require a labour force greater than a nuclear family, but would certainly be more easily managed than 40 ha.

If we further adjust the grain requirements to reflect a household need of 1.5 metric tons and then apply the Tell Halula yield estimates of 1.56 metric ton per hectare, less than 1 hectare would have been required to produce 200 kg per person annually for those living in Building Complex I. A nuclear family could work the land required with little effort, especially if, as evidence suggests, cattle were being used as draught animals (see section 4.3).

Given the above-mentioned data I believe one could expect higher yields than 500 kg per hectare and plots of sufficient size could have been farmed by the inhabitants of each household. The paleoclimate (Chapter 2.8.1) was quite dry during the Middle Chalcolithic, but proximity to the river likely made farming sustainable. It is premature, with our current data, to begin discussing landlords and serfs, or any type of economically-based inequalities. Instead, we could start considering alternative uses for the storage silos, once again questioning how specific features were used at Tel Tsaf.

Dung clearly served as fuel on site, as seen in the density of dung spherulites within the roasting pit deposits of Building Complex I. Cooking was taking place in Building Complex I and there

189 were silos very near the roasting pits. It makes sense that, if dung was a primary fuel source, its users would store it close to where it was to be used. Most cultures do not use dung straight out of the pen as fuel. In fact, most ethnographic accounts suggest dung is mixed with chaff and then the pellets are dried prior to use to assist in burning (Charles 1996; Sillar 2000). Creating dung pellets could be classified as a specialized skill, given the investment of time and effort necessary to prepare the pellets. Keeping them stored in an enclosed environment for future use then makes sense.

Chaff was likely used in the production of dung pellets for fuel, and we know that it was also used to temper mud-rendered floors and mudbricks, as well as spread out within courtyards. It is therefore reasonable to consider that some of the storage space within Building Complex I was used to store chaff once the grain had been separated. Evidence for the presence of stored chaff from processed grain has been found at the Pre-Pottery Neolithic sites of Jerf el Ahmar and in northern Syria (Willcox and Fornite 1999; Willcox and Stordeur 2012). Flotation at both sites revealed large quantities of charred, winnowed chaff in building material, indicating that chaff was an important resource and stored on-site.

That materials other than grain may have been stored at Tel Tsaf does not contradict the hypothesis of surplus, even of grain. As an example, let us take Building Complex I, Phase 3 as the basis for a model of potential storage use and capacity at Tel Tsaf. Botanical remains from flotation samples taken during excavation indicate that both lentils and chickpeas were present at Tel Tsaf (Graham 2014) making storage of these goods a strong possibility. If we make the very conservative assumption that each silo was 1 m in height, and that a different item was stored in each of the four silos, table 5.5 demonstrates what storage potential could have been at Tel Tsaf.

190

Volume Area potential storage capacity Silo Item stored (1m height) m² (tons) m³ 286 Barley 7.1 7.1 4.4 339 Wheat 7.1 7.1 5.5 53 Chickpeas 6.6 6.6 4.9 415 Lentils 4.9 4.9 3.9

Table 5.5: Estimated silo volume and potential storage capacity for Building I, Phase 3 at Tel Tsaf. The capacities for barley and wheat storage are based on the averages of the density (kg/m³) for each outlined in table 5.4. A.G.Contractors (2013) and LeonCookSilos (2013) estimate the density of chickpeas at 740 kg/m₃. Thomas, et al. (2001) state the density of raw lentils to be 800 kg/m³.

Table 5.5 demonstrates that, even if an ultraconservative estimate of silo height (1 m) is assumed and a different product was stored in each silo, the amount of grain storage available per building complex is still significant. If all 4 silos in Building Complex I were in use, storing different foodstuffs, there is a potential storage capacity of 4.4 tons of barley, 5.5 tons of wheat, 4.9 tons of chickpeas and 3.9 tons of lentils. Based on Garfinkel, et al.’s (2009) statement that a nuclear family requires 1.5 tons of grain (which is actually enough for 7.5 people, based on their own stated requirement of 200 kg/year per person), the silos far exceed the needs of the inhabitants of Building Complex I. The numbers would also provide grain for animals not pastured off-site, such as cattle.

Even if one accepts that the silos were completely full of grain, as Garfinkel, et al. (2009) suggest, there is no evidence of any sort of administrative structure or means of redistribution within Building Complex I. It is reasonable to expect some surplus with such large storage capacity within one complex, but the excess suggested by Garfinkel, et al. (2009) would require some system for measuring, recording and tracking the grain moving into and out of the building.

The above example is not meant as a working model for storage at Tel Tsaf. The use of barley, wheat, chickpeas and lentils as potentially stored items is one possible scenario based on botanical analysis from the site (Graham 2014). Evidence also suggests that chaff and dung pellets were likely stored on site but were excluded from the above example as density estimates were unavailable. The discussion based on Table 5.5 is simply a demonstration of the significant potential storage at the site. The consideration of alternative silo use does not diminish or

191 remove surplus from the discussion. In fact, it demonstrates that a number of different items could be stored and considered surplus and, therefore, potential wealth.

5.2.2 Animals at Tel Tsaf

The faunal analysis from Tel Tsaf, discussed in section 4.3, shows that sheep and goat were primarily raised for meat production. The kill patterns indicate general herd management, with no indication of extensive wool or milk exploitation (Hill 2011). Excavations at the site support this evidence. One would expect to find artifactual evidence such as spindle whorls or churns if wool or milk were being intensively processed. Neither of these items is present, let alone on a scale that would indicate production beyond household needs. This suggests that sheep and goat were consumed or exchanged at a rate that did not produce significant extra animals that could be otherwise exploited. This does not conclusively prove there were no ‘surplus’ sheep or goats at Tel Tsaf, but it does indicate that any extra ungulates were not being used for intensive secondary product exploitation.

Cattle, on the other hand, were being used as draught animals as well as for meat (Hill 2011). This may be an indicator that the number of meat-supplying animals at Tel Tsaf was greater than required for basic subsistence needs. It may alternatively indicate that the need for grain was greater than the need for meat, even if animals were not considered surplus. Ownership of cattle is unclear as we currently have no evidence for whether they were kept on-site or off-site. If they were on-site, it is not clear whether they were kept at the household level, multi-household level, or community level. The use of cattle for manual labour is likely a contributing factor in the production of surplus grain at Tel Tsaf. If grain was being stored at the household level, it is possible draught animals were also maintained at this level. Alternatively, cattle may have been shared among households or have been maintained at a community level.

Faunal evidence suggests that pigs were used primarily for meat, with the majority being slaughtered within the first year of their lives (Hill 2011). It is possible that they were being kept on-site and that their dung was being exploited as fuel. Pigs can produce offspring up to four times a year with litters varying between six and 10 or even 15 piglets, making them a very reliable resource (Rosenberg and Redding 1998; Zeder 1996). A small number of pigs could

192 continually produce dung and offspring year-round, allowing primary and secondary product exploitation. Modern piglets can grow as large as 220 lbs by the time they are six months old, demonstrating how quickly piglets become a viable and abundant source of meat (Zeder 1996). Even assuming lower weights for ancient pigs, one reproductive female pig, maintained in a small pen, can easily provide for a nuclear family for a year (Zeder 1996). It is preferable to pen female pigs independently or in small groups as crowding may increase the spread of disease and increase the chance of injury due to fighting (Zeder 1996). The Tel Tsaf faunal data indicate that only 17% of the pig bone originated from animals older than 3.5 years of age (Hill 2011). Ethnographic data from modern Spain and Portugal show that pig farmers slaughter 80-90% of their piglets by the time they reach one year, meaning only 10-20% of their herds are adult pigs (Hadjikoumis 2012). This puts the adult pig population of Tel Tsaf in line with meat exploitation.

The question of who owned pigs is an important one at Tel Tsaf. As previously mentioned, it is generally assumed that pigs were owned and raised at the household level. However, the presence of animal pens within Building Complex II, Phase 3 may present evidence of pigs being raised outside of a household. Currently there is no evidence of a house or household directly related to Building Complex II, but it could outside the excavated area. It is clearly separate from the household space of Building Complex I. It could be that Building Complex II represents private, but separate storage for pigs. Alternatively, it could represent a communal male pig-penning area as a community would only require a small number of males to impregnate privately owned females. Modern pig farmers of Spain and Portugal maintain a male population that only makes up about 7-17% of the overall reproductive population (Hadjikoumis 2012). Male pigs are territorial and feisty, so keeping them in separate pens within the same area would reduce the risk of injury and maintain a healthy breeding population (Zeder 1998).

5.3 Control of Resources – Public vs. Private

The identification of storage capacity is important, but perhaps more enlightening is where the storage is located. Discussion not only of how much surplus existed, but how it was controlled and distributed, provides clues to social and economic relationships. Storage located in open, public spaces suggests the resources were procured by and distributed among extended families

193 or communities. However, storage areas that are contained within private courtyards or houses represent goods that are owned by a smaller household unit (Banning 2010; B.F. Byrd 1994; Flannery 1972). In populations where resources are scarce, a community approach to hunting and gathering may help reduce risk and ensures everyone contributes to the communal supplies. However, once resource procurement becomes less risky, and surplus is being produced, the strategy may change. A community that can produce significant surplus may not require the entire population to be engaged in activities such as farming and herding. However, rather than having a portion of the population working towards these goals and feeding otherwise ‘idle’ mouths, privatisation can occur. A self-sufficient household that is able to provide its own subsistence needs may then move storage away from public space and into private areas. In order for this next possible economic step to occur, a population needs to have some specialisation, as discussed above. A surplus of grain does not help a household unless they can use it to acquire something they do not have. Alternatively, craft specialists or herders rely on others to ‘pay’ for their goods with grain or products they are not producing. Only when some kind of exchange is taking place can surplus begin to be considered wealth.

Alternatively, the dry and unstable climate during the Tel Tsaf occupation may have influenced the need to store large amounts of grain as a risk mitigation strategy. Over the 600 years of occupation, there were likely periods of drought when grain production was severely affected. While communal storage and sharing of good can be one risk mitigation strategy during long periods of shortage, yearly variability may influence other strategies. Private stores may be used to ensure subsistence needs are met within a household. However, larger storage capacities may also allow a household to trade surplus to those households in need, translating grain stores into wealth or prestige. It is likely both scenarios played out at Tel Tsaf. There would need to be some risk mitigation, but the diameter of the silos are comparable to PPNA and EBII examples from the Levant. Kuijt and Finlayson (2009) have noted that the probably silos at Dhra’ measure about 3 m in diameter, very similar to those at Tel Tsaf. The EBII examples from Tel Beit Yerah and Arad measure c. 3 - 6m (Currid 1985), also in line with Tel Tsaf. Both these periods exhibit wetter climates more conducive to consistent successful agriculture, indicating that large storage capacities existed outside the Middle Chalcolithic. It should be noted that the extent of private or

194 public access to stores at all the above mentioned sites would directly influence the economic role grain played in each case.

The architecture at Tel Tsaf lends itself well to discussions of private and public space. The site appears to be composed of courtyard complexes. Unfortunately neither of the Phase 3 complexes so far discovered has been fully excavated. All discussions have been based on the assumption that the courtyards would have been fully enclosed, representing private space. We will continue with that assumption during this discussion, although future excavation may force us to reconsider.

Building Complex I likely housed a nuclear family, and we have already established that the storage capacity exceeds the needs of a household of that size. The bounded nature of the courtyard and the containment of surplus within the walls indicate that the storage was intended for private purposes. Storing subsistence and surplus grain, dung or other material within controlled, walled structures indicates a desire to protect one’s property (Banning 2010). This action exerts control over who has access to the stores. The desire to protect one’s goods from others suggests that there are others who have less and desire more. It is possible that protection from animals was needed as well. Silos are also designed to keep out unwanted natural elements, such as weather, raiding birds and insects.

Building Complex II presents an interesting study in public and private space. If we work on the assumption that the circular rooms housed pigs and that domestic tasks were not performed in the complex, the question becomes, who owned the pigs and silos? The silos themselves are large enough to have provided more than enough grain to support a nuclear family for a year. In addition, if we assume even just one pig per sty, the two pigs would easily provide sufficient food and dung to support a nuclear family. At present, however, we have no evidence to suggest who filled the silos and raised the pigs. The easiest explanation would be that excavations just did not uncover the living area of the complex, which would have lain to the south or east.

Alternatively, Building Complex II may represent a secondary property owned by a household located elsewhere on the site. Perhaps the inhabitants of Building I or another complex opted to separate their animals from their personal living space. The evidence suggests that grain surplus

195 was very significant on-site, so perhaps the household stored a separate food supply in Building Complex II. This could have been used to house the animals they kept within the same complex. Having animals in a separate building has many advantages. The pigs would have been separated from the daily household activities, including cooking, preventing them from interfering with household activities. In addition, there would be no waste products near these same areas, reducing unpleasant smells and food contamination.

A third potential explanation for Building Complex II is that it was a communal pig-penning area. Each pen may represent a different household’s animal. Or perhaps the community as a whole contributed to the feeding and maintenance of the animals. This scenario is more likely if the pens housed male pigs, as each household would not require its own. Communal stud animals would also help diversify the gene pool, allowing sows to breed with multiple males.

All three of these hypotheses require a lot more evidence than we currently have available. Further excavation of the site would allow us to understand how frequent these structures are and how they relate to other structures, such as Building Complex I. Once complete courtyard structures are uncovered and their relationship to each other understood, we can make more meaningful statements regarding the significance of public and private space at the site. For example, is the level of storage in Building Complex I typical or atypical? Are there multiple penning complexes, or is Building Complex II an oddity?

Despite lacking answers to these questions, we do know that privacy was important at Tel Tsaf. The surplus that is evident was not accessible to the community at large, but rather kept within the confines of Building Complex I. Building Complex II represents storage of grain, as well as livestock. Whether it was private, shared or owned by an individual household is unclear.

5.4 Power, Conspicuous Consumption and Feasting

If surplus can be an indicator of wealth, and control is a form of power, the question becomes how was private storage conveyed to the public if it was kept behind closed doors? How private storage translates into conspicuous consumption is an important issue to address. Equally important is whether or not private surplus was intended as wealth. Control of surplus does

196 represent power over goods, but it may not indicate power within a larger community. In cases where it does indicate wealth and influence on a larger scale, there needs to be a public display of consumption of goods in order for the community to understand the extent and significance of wealth. In addition, public events help to navigate social relationships once private space becomes dominant within a community (Byrd 1994). Ben-Shlomo, et al. (2010) have suggested that feasting was the mechanism used at Tel Tsaf.

Feasting is a much-discussed topic within archaeology these days (Bray 2003; Dietler and Hayden 2001; Goring-Morris and Horwitz 2007; Gumerman 1997; Hamilakis and Konsolaki 2004; Helwing 2003; Pauketat, et al. 2002; Spielmann 2002; Twiss 2008). A hard and fast definition of feasting is difficult as the scale, frequency and purpose of “feasts” varies so greatly. A feast is generally viewed as a meal that is distinguished from everyday meals, often involving more people and special foods (Twiss 2008), but participants can be two people or an entire community. Feasts can also be used to promote social competition or social integration (Dietler and Hayden 2001). They are often associated with surplus because of the amount of food and drink involved. Given the nature of the storage and site layout at Tel Tsaf, it is natural to discuss feasting as a potential means of negotiating or defining changing social and economic roles within the community.

Twiss (2008) addresses the archaeological correlates one might expect in order to identify feasting in prehistory. These include middens with abundant food remains; storage facilities (silos and pens); large-scale cooking facilities; varied cooking and serving wares; rarely eaten or symbolic foods; evidence of alcohol consumption; public feasting sites and ritual behaviors. She stresses that not all these elements need be present in order for feasting to have taken place. Indeed, the scale of feasting would greatly affect the archaeological deposits. Ben-Shlomo, et al. (2010) approached the data from Tel Tsaf addressing some of these indicators.

The first evidence presented is the presence of the roasting pits (earth ovens), mainly located within Building Complex I. Below is a breakdown of roasting pits including faunal data (table 5.6).

197

Number of Identifiable Specimens Dimensions (m) Estimated Bone Sheep/ Wild Pit No. (l x w x d) Lining volume (m3) count Goat Pig Cattle species C46 1.5 x 0.8 x 0.3 clay 0.36 57 17 21 10 8 C96 1x 0.6 x 0.32 clay 0.19 4 2 1 1 C137 1 x 0.5 16 2 11 3 C218 1 x 0.5 x 0.3 0.15 0 C261 0.8 x 0.6 x 0.4 clay 0.19 17 3 10 4 C284 1.3 x 0.7 bricks 26 5 19 2 C289 1.2 x 0.5 bricks, plaster 0 C290 1.6 x 0.8 x0.4 bricks, plaster 0.5 17 16 1 C308 1.8 x 0.7 x 0.4 bricks 0.5 2 1 1 C349 0.8 x 0.6 x 0.3 mud 0.14 4 1 1 2 C417 2.0 x 0.8 x 0.6 bricks 0.9 21 8 4 9 Total NISP 164 39 83 30 11 Percent of total NISP 23.8 50.6 18.3 6.7

Table 5.6: A breakdown of roasting pits within Building I, Phase 3 with physical description and breakdown of faunal remains found within each (data from Ben-Shlomo, et al. 2009; Hill 2010)

The authors present a good discussion regarding the anthropological and archaeological evidence for the use of the pits as cooking instillations. The scale of cooking is assumed based on the large number of pits, as well as their size. The pits are only found within the courtyard of Building Complex I and only during Phase 3. Unfortunately we are not able to identify whether each pit was used once or repeatedly. It is possible, as discussed, that a single pit may have been used and cleaned a number of times. We are also unable to determine whether all the pits were used at once, or whether they were used during separate events. If two or more of the pits represent a single cooking event, it would be strong evidence of large-scale food sharing and feasting. The use of the roasting pits individually as everyday food preparation would support the evidence for Building Complex I as a home, rather than feasting facility.

The faunal data creates a stronger dataset to support the hypothesis of feasting at Tel Tsaf. During the 2005 season of excavation, a pit (174) was uncovered to the south of Building II. It contained nearly 50% pig bones (by NISP n=94, MNI=5), including evidence for at least five pig mandibles (Ben-Shlomo et al 2009; Hill 2010). Of the pig bones, 81% were cranial or axial, and meat-rich upper and lower limb bones were not represented. This deposition was likely the result of a single event, indicating disposal of the partial remains of at least five pigs at one time.

198

Pig bones also represent a significant portion of the faunal remains contained within the Building Complex I cooking pits. Across the site of Tel Tsaf, pig bones only represent 30% (NISP) but they represent about 51% of the specimens found within the roasting pits (Ben-Shlomo et al 2009; Hill 2010). This indicates a general preference for pig during the cooking events within the roasting pits, and within Pit 174. However, in each context, pig bone was mixed with at least one other species, suggesting that, despite the preference, pigs were never cooked to the exclusion of other animals.

The presence of surplus farmed goods, as indicated by silos, combined with the site layout, faunal data and ritual evidence does suggest that Tel Tsaf could have hosted large feasts and feasting probably took place at the site on some scale. The evidence from Pit 174 is compelling but the phase of use and building complex of consumption cannot be confirmed, making it difficult to expand discussion. All previous discussion has indicated that Building Complex I is a house. The evidence presented regarding feasting is not strong enough to re-consider and view it as a large, public feasting area. However, it is possible and likely that the household played host to others in feasts of some scale during their occupation of the area. At present, the nature and scale of these events can only be guessed.

199

Chapter 6 The Great Chiefdom Debate

Most scholars working within the Southern Levant acknowledge that there is good evidence of some form of changing social, religious and economic systems during the Chalcolithic period (Bourke 2002; Gal, et al. 1996; Gopher and Tsuk 1996; A.H. Joffe 2003; Levy 1986a; Rowan and Golden 2009; Schick 1998). To-date, the only suggested theoretical framework to understand these dramatic changes to daily life is the concept of chiefdom level organization (Levy 1983, 1986a, 1986b, 1995). Section 2.7 provided a brief overview of Levy’s assertions, and the theoretical framework on which they are based. This chapter will expand on the evidence that scholars have used to support the chiefdom framework and outline opposing views. The evidence from Tel Tsaf will be explored within the same framework, further testing the validity of the chiefdom model as applied to Levantine Chalcolithic society. I conclude with potential alternative frameworks for understanding the Chalcolithic archaeological record.

6.1 The Case for Chiefdoms in the Southern Levant

As the strongest proponent for using ‘chiefdom’ to describe Late Chalcolithic society in the Southern Levant, Levy relies heavily on Renfrew’s (1973) work on defining chiefdoms in Neolithic Wessex (England) to frame archaeological indicators for chiefdoms in the Levant (1973, Table 2.5). Both Renfrew and Levy acknowledge that the list is a guideline of indicators, and not all of them need be present in order for a chiefdom to exist. Since first introducing the concept of chiefdoms in the Southern Levantine Chalcolithic more than 30 years ago, Levy’s ideas have received limited support from other academics. Table 6.1 outlines the evidence that Levy and his colleagues use to support the case for chiefdoms.

200

Indicators Evidence Sources Distinct Regional Groups ● Increased regional diversity Levy 1986 Greater Population and Density ● Teleilat Ghassul - 20 ha Levy 1986 Centres coordinating social, religious and ● Large 'public' buildings at Teleilat Ghassul and Shiqmim Levy and Alon 1985; Levy economic activity 1986; Levy 1995; Burton and Levy 2011 ● Two-tiered settlement system: Nahal Beersheva; Teleilat Levy 1986; Levy 1995; Levy Ghassul; Coastal Plain et al 2006; ● Formal Cemeteries: Shiqmim; Adeimeh; Coastal Plains; Levy 1986; Levy 1995 Nawamis in Sinai; Nahal Qanah; Peqi'in ● ritual or ceremonial sites: En Gedi, Gilat, Teleilat el-Ghassul; Levy 1986; Levy 1995; Levy open air sanctuaries in the Negev 2006 Craft Specialization ● Metallurgy : Nahal Mishmar hoard; Wadi Faynan; Timna; use of Levy 1986; Levy and Shalev lost-wax casting 1989; Levy 1995; Levy 2006; Burton and Levy 2011 Lithic Production: workshops in Northern Negev, Ceramics : introduction of slow wheel; increase in quality and variety More Defined Territorial Boundaries ● Formal Cemeteries: Shiqmim; Adeimeh; Coastal Plains; Levy 1986; Levy 2006 Nawamis in Sinai; Peqi'in Organization/Deployment of Public Labour ● Formal Cemeteries: Shiqmim; Adeimeh; Coastal Plains; Levy 1986; Levy 2006 Nawamis in Sinai; Peqi'in Greater Productivity ● Irrigation: Nahal Beersheva Alon and Levy 1980; Levy 1986; ● Secondary Production exploitation Levy 1986; Levy 1995 ● Subterranean storage in Beersheva valley Burton and Levy 2011 Distinctive dress or Ornament indicating ● Nahal Mishmar: maceheads, standards, crowns Levy 1986; High Status Competition and Warfare ● Shiqmim: Cranial trauma on intramural burial of adolescent Dawson et al 2003; Burton male; violent imagery evoked by maceheads and Levy 2011 ● Destruction levels' at Shiqmim Burton and Levy 2011

Table 6.1: Overview of the proposed evidence for chiefdom level organization in the Late Chalcolithic of the Southern Levant.

201

6.1.1 Distinct Regional Groups

The first archaeological feature that Levy notes as indicative of chiefdom-level social organization is the presence of distinct regional groups. However, regional diversity was also a feature of the Earlier Chalcolithic phases. As discussed in Chapter 2, the Ghassulian is the most widely discussed cultural entity from the period. However, a good deal of data from the Beersheva region demonstrates similarities to, as well as some variation from, the Ghassulian sites (Alon and Levy 1980; T. E. Levy, et al. 2006; Levy and Alon 1979, 1982, 1985a, 1985b, 1985c, 1987a, b; Levy, et al. 1994; 1991). For example, site layout and size vary from the Ghassulian type-site of Teleilat Ghassul, the most apparent difference being the prominence of subterranean chambers in the Beersheva settlements. To the north, the Golan sites present yet another regional variation. The material culture shows some similarities to typical Ghassulian objects such as v-shaped bowls, but there do not appear to be any large villages in the Golan, rather, settlements consist of small clusters of broadroom houses , sometimes in chain-like arrangements (Epstein 1978a, 1985, 1988, 1998). Similarities, such as broadroom buildings and ceramic forms, across all three regions suggest that even if there was not a single homogenous cultural entity, there certainly was some degree of interaction among groups occupying these sites.

6.1.2 Greater Population and Density

Increases in overall population in the region and population density in larger habitation centres can be difficult to determine because both site identification and excavation can be biased. For example, without fully excavating every site it can be difficult to determine the extent of the settlement, or to determine the number and nature of phases of occupation present. Our best evidence for site size and extent comes from Teleilat Ghassul where extensive, but incomplete, excavations revealed a site more than 25 ha in area (Bourke 2002). This represents a marked increase in size from Early Chalcolithic sites in the southern Levant. An increase in population and population density may also be inferred from the large number of sites detected or excavated in the Beersheva region, including the substantial settlement at Shiqmim (T. E. Levy, et al. 2006). In addition to an increase in known sites and site sizes, many Chalcolithic sites are found in areas previously uninhabited, further suggesting population growth (Finkelstein and Gophna 1993; Gophna and Portugali 1988; Gophna and Tsuk 2005; Khalaily and Marder 2010).

202

Although use of this evidence requires caution, it does at least suggest that there were increases in the size and density of population during the Chalcolithic.

6.1.3 Centres for Coordinating Social, Religious and Economic Activity

6.1.3.1 Economic Centres

Levy (1986a, 1995, 2006) has suggested that a two-tiered settlement system may be an economic indicator of chiefdoms in this period. Levy’s assertion of such organization is based primarily on his work in the Negev of southern Israel. It is clear from his excavations that not all sites within the Beersheva Valley are the same size, but artifact assemblages are similar at both large and small sites (T. E. Levy, et al. 2006). Levy suggests that the presence of large central villages surrounded by smaller hamlets represents new economic relationships, as well as social differentiation and varied integration (Levy 1995). He further suggests that the same types of relationships may have existed in the East Jordan Valley, where we find similar variation in site size. Bourke (2002) has noted that sites in the East Jordan Valley region range in size from a few over 10 ha and many 5 ha or less in size. Even with this evidence, Bourke states that ascribing rank-size distributions or sub-regional economic relationships is difficult given the limited archaeological evidence. He suggests that Teleilat Ghassul may have served as a regional cultural and trade centre, but the hinterland was likely autonomous (Bourke 2002).

Recent spatial analyses of the Negev sites have presented conclusions similar to Bourke’s (Fletcher 2008; Winter-Livneh et al. 2010). Fletcher (2008) employs Average Nearest Neighbour Distance analysis, Getis-Ord General G and Moran’s I tests to the site data from the region in order to establish whether site distribution was random or clustered. He concludes that, while sites clustered around wadis, presumably as a water source, there is no detectable pattern for sites within 20km2 of each other. He further concludes that such patterns do not reflect the type of distribution expected when dealing with two-tiered chiefdom societies. Winter-Livneh et al. (2010) have revisited Fletcher’s method, but use newly available excavation data and apply Ripley’s K function to the analysis. Their aim was to find more appropriate methods for answering questions regarding preferred locations and where and how the landscape was used. Their results indicate clustering on both local and regional scales (Winter-Livneh et al. 2010). However, they suggest this pattern is related to environmental factors, most notably the volume or flow of water available in particular areas along individual wadis. Despite their opposing

203 results regarding site distribution patterns, both Fletcher (2008) and Winter-Livneh et al. (2010) conclude that social and economic issues did not strongly influence site relationships within the Negev region. These methods have yet to be applied outside the Negev or to the southern Levant as a whole.

6.1.3.2 Social and Ritual Centralization

Key indicators of probable centralized social or ritual activities include caches of metal items, as at Nahal Mishmar, the presence of ‘religious’ or ‘cultic’ centres such as and Gilat, and formal cemeteries. Copper is rare at Chalcolithic sites, having been found at only 25 sites, and the Nahal Mishmar cache represents the largest collection of copper items found to date, with 417 copper artifacts (Shugar and Gohm 2011). By contrast, the number of copper artifacts documented from all other Chalcolithic sites taken together is quite small, with only 142 present across 25 sites (Table 6.2)(Garfinkel, et al. 2014; Shugar and Gohm 2011). The fact that nearly 75% of copper items were found at Nahal Mishmar and only six of the remaining 25 copper- bearing sites yielded more than 10 items, suggests that copper was not a common, daily-use material during the Chalcolithic. The only other metal found in Chalcolithic context is only found at Nahal Qanah cave. Here two gold and six electrum ring-shaped ingots were discovered, all with similar dimensions (Gopher and Tsuk 1996).

The majority of copper forms discovered tend towards symbolic, rather than functional purposes. Axes, chisels and awls are the most widespread artifact type, found at 19 sites in the Southern Levant, but they only represent about 16% of the entire copper collection (Shugar and Gohm 2011). The remaining artifacts fall into the categories of maceheads, standards or crowns, and some unspecified miscellaneous items. The axes, chisels and awls generally seem to be cast in open molds, most often with almost pure copper. The more symbolic forms were cast using the lost-wax technique, frequently using a copper alloy containing arsenic, antimony and nickel (Levy and Shalev 1989; Shalev 1991, 1995, 1999; S. Shalev and P.J. Northover 1987; Shugar and Gohm 2011).

204

Axe / Site Awl Macehead Standard Crown Misc. TOTAL Chisel Abu Hamid 2 1 3 Abu Matar 1 1 4 1 4 11 Arad 1 1 Azor 2 2 Bir es-Safadi 3 1 1 5 Gilat 1 1 2 Giv'at ha-Oranim 6 2 4 3 2 2 19 Horvat Beter 2 2 Ketef Jericho 2 1 3 Makuch 5 1 1 7 Meser 5 1 6 Nahal Ashan 1 1 Nahal Lahat 1 1 Nahal Mishmar 16 1 256 118 10 16 417 Nahal Qanah 1 1 1 8 11 Nahal Ze'elim 1 3 4 Nevatim 1 1 Neve Noy 2 3 2 1 2 10 Palmachim 1 3 4 Peqi'in 2 2 3 7 Shiqmim 5 12 3 2 1 2 25 Shoham 1 1 Tall al-Magass 2 1 3 Teleilat Ghassul 3 8 1 12 Umm Qatafa 1 1 Tel Tsaf 1 1 TOTAL 55 38 274 131 15 47 560

Table 6.2: A summary of copper items found at Chalcolithic sites in the southern Levant (Shugar and Gohm 2011, Garfinkel et al. 2014).

Table 5.1

The site of Ein Gedi is located on a high promontory to the west of the Dead Sea. Architecturally, it consists of two broadroom buildings and a gate within an enclosing wall (Bourke 2002; Ottosson 1980; D. Ussishkin 1980). There are no obvious domestic dwellings in the immediate vicinity, and its location high in the hills overlooking the Dead Sea suggests further that the site had a special purpose. A number of pits within the larger broadroom contained burnt horn cores, as well as fragments of pedestalled and fenestrated ceramic bowls. A zoomorphic figurine was also found on a semi-circular altar-like feature (Ussishkin 1980). It is widely accepted that Ein Gedi is an atypical Late Chalcolithic site with heavy ritual use, but the specific scale and nature of this use is elusive (Rowan and Golden 2009).

205

The site of Gilat, located just north of the Beersheva sites, also presents evidence of ritual practices (Alon and Levy 1989; Levy 2006). Most notable are the anthropomorphic and zoomorphic ceramic figurines (Figure 6.2). One figure, called the Gilat Lady, depicts a seated female with engraved genitalia holding a churn on her head (Commenge, et al. 2006b). There is speculation that this figure, unparalleled elsewhere, may represent a fertility goddess (Amiran 1989; Fox 1995; Joffe, et al. 2001; Weippert 1998). A figurine depicting a ram with three cornets was also found at the site, and is similar to one of a laden donkey found at Ein Gedi (Rowen and Golden 2009). Also of note at Gilat are stelae, violin-shaped figurines of stone (Commenge et al. 2006b), distinctive ground-stone vessels (Rowan, et al. 2006) and a dog burial (T.E. Levy, et al. 2006). All these finds were located in a complex composed of large buildings and a large open area. As with Ein Gedi, the scale and nature of ritual activities is unclear, but some form of large-scale activity likely took place there.

Ein Gedi and Gilat represent the best candidates for sanctuary or shrine sites during the Late Chalcolithic, but there is also evidence of a potential temple precinct at Teleilat Ghassul (J. B. Hennessy 1982). Building 78 is a broadroom complex with multiple rooms. Non-domestic evidence uncovered includes two wall paintings, offering pits, over 30 cornets similar to those from Ein Gedi and Gilat, and an “altar” (Koeppel, Mallon, et al. 1940; Koeppel, Senès, et al. 1940; A. Mallon, et al. 1934).

Burial practices in the Late Chalcolithic were not uniform, but most mortuary sites fall into one of two broad categories; constructed off-site cemeteries or burial caves. The best examples of burial structures come from Shiqmim, where over 100 cist graves and grave circles have been excavated (Levy and Alon 1982, 1985a, 1985b, 1987a). The cists generally lack mortuary remains, but sometimes contain v-shaped bowls (Levy and Alon 1982; 1987a). The grave circles tend to contain secondary remains of several individuals and limited grave goods, most commonly v-shaped bowls, shell beads and fan-scrapers (Levy and Along 1982; 1987a). Bar- Yosef (1986; 1977) and Levy and Alon (1985a; 1985b) speculated that the Shiqmim grave circles may be related to the nawamis near ‘Ein Huderah and Gebel Gunna in the Sinai, owing to their similar construction and the presence similar lambis shell bracelets at both locations. Other potential examples of burial structures are less complete. It has been suggested that Adeimah, near Teleilat Ghassul, served as the community’s cemetery (Neuville 1930; Stekelis 1935). The

206 site also contains and tumuli, but these may post-date the Late Chalcolithic period (Rowen and Golden 2009).

Burial caves contain some of the richest mortuary data from the period. They are prominent along the coastal plains and in the northern highlands of Israel. They are in contrast to the built structures that tend to be located inland. Peqi’in in northern Israel is a burial cave that consists of three archaeological units within a natural karstic cave. All the mortuary remains contained in the cave are secondary deposits, some randomly placed while others appear intentionally lain out (Gal, et al. 1997). The cave contained many ceramic ossuaries and large jars, one ivory figurine, violin-shaped figures, high-footed fenestrated v-shaped bowls, as well as other ceramics (Gal, et al. 1997). The ossuaries from Peqi’in and coastal caves, such as Azor, often appear in the form of house-like boxes (Ben-Tor 1975; Perrot, et al. 1980)(figure 2.5). These ossuaries exhibit details like cross-beams and windows that parallel houses or other structures. Others are more jar-like, often displaying facial features (Ben-Tor 1975). Various ossuary forms are often found together, but do not show any spatial patterns in their co-occurrences.

The introduction of formal burial grounds and the inclusion of ivory, shell or bronze items in graves during the Late Chalcolithic are potential indicators of increased community ritual and the creation of public centres in which to carry out these rituals. Burial ritual was by no means a new invention in the southern Levant, which has a rich history of mortuary ritual since the Palaeolithic. However, the reappearance of formalised cemeteries and burial caves for the first time since the Natufian period, combined with regional ritual sites appearing at Ein Gedi and Gilat, suggests that the underlying belief system may have extended beyond the local community and was practiced at a regional scale.

6.1.4 Defined Territorial Boundaries

Saxe (1970) and Goldstein (1981) have suggested that extramural cemeteries and burial sites may be indicators of territorial boundaries and Levy (1986; 2006) draws upon their ideas in his discussion of territorial boundaries. Central to the chiefdom debate is the concept that tombs could be used not only to house the dead, but to lay claim to regional resources by announcing permanency and establishing historical ties to the area, especially when they were highly visible on the landscape (Chapman 1981; Renfrew 1973). While we can attribute many extramural burial sites to the Chalcolithic period, we can associate few of these with any confidence to

207 habitation sites. Furthermore, many of the off-site burials are in caves, especially along the central coast (Chapter 2.5.2), and not as prominent on the landscape as the other built grave sites. Winter-Livneh, et al. (2012) explored the question of burial caves as territorial markers by employing spatial analysis (Kernel density) and view-shed analysis. Through the analysis of 24 burial sites along the Israeli coast, and their nearest-neighbour habitation sites, the authors established that the burial caves greatly increased the area of land that could be monitored. The limited overlap of visible areas suggests that the caves may have marked a new form of land tenure in the region (Winter-Livneh, et al. 2012). Winter-Livneh, et al. (2012) emphasize that this likely represents cues for intra-community relations, rather than a safeguard against foreign attackers. They also suggest that this pattern represents negotiation of boundaries among local tribes or clans rather than indicating developed chiefdoms.

Another possible indication of territoriality is warfare. We do not have a great deal of evidence of this, as towns remained un-walled and we find little or no evidence of violent destruction of Late Chalcolithic sites. The prevalence of maceheads may suggest that warfare or violence was a concern. Examination of a skeleton from an intramural burial at Shiqmim revealed cranial damage consistent with blunt-object trauma, possibly a macehead (Dawson, et al. 2003). Thus far, this example is exceptional, and, if anything, suggests a local feud or interpersonal conflict rather than full-scale warfare. Also, the ceremonial context of most copper maceheads suggests that they had greater symbolic value than practical use. The possibility of their prominence in hoards may be related to the importance of the weapons in society, but more evidence is required to support the idea that their main role was in inter-territorial warfare.

6.1.5 Greater Productivity

Levy argues that greater productivity is also a hallmark of chiefdoms, and the Late Chalcolithic presents plenty of evidence to support this. One of the best indicators of increased production in prehistoric societies is greater visible capacity for storage, most notably of grains and foodstuffs. Storage facilities take on many forms at Late Chalcolithic sites, including subterranean stone- lined silos, small rooms, clay- or stone-lined silos, paved platforms and bins (Banning 2011). These features often appear together at numerous sites such as Abu Hamid (Dollfus and Kafafi 1993; 1988), Teleilat Ghassul (Blackham 1999; Bourke 2001; Koeppel, Mallon, et al. 1940), Gilat (Levy 2006), Sahab (Ibrahim 1984), Tel Ali (Garfinkel 1993a), Abu Sneshleh (Kerner, et

208 al. 1992; Lehmann, et al. 1991), the Golan sites (Epstein 1998), Shiqmim (Levy and Alon 1985b), Tel Te’o (Eisenberg, et al. 2001) and Neve Yam (Galili, Eshed, et al. 2009). As discussed in context of the Tel Tsaf silos, it is difficult to determine the capacity of storage facilities, what was stored in them, or whether they represent surplus. In this case, the prevalence of identifiable storage facilities across Late Chalcolithic sites is the important issue in identifying increased production.

6.2 All Hail the Chief at Tel Tsaf?

Tel Tsaf and the Middle Chalcolithic immediately proceed the Late Chalcolithic period chronologically, and we might expect at least some of the features of Levy’s Chalcolithic chiefdoms to be evident to some degree at the site. Such evidence could be used to support the case for the local emergence of chiefdoms. Table 6.3 revisits Levy’s criteria for identifying chiefdoms in the Levant, with evidence from Tel Tsaf noted.

Tel Tsaf? Chiefdom Indicators Evidence Yes Maybe No Distinct Regional Groups Greater Population and Density extensive storage capacity (silos) Centres for coordinating social, religious On-site feasting evident, scale unclear and economic activity Craft Specialization copper awl; ostrich egg shell beads; figurines More Defined Territorial Boundaries Organization/Deployment of Public Labour Greater Productivity increased grain storage (silos) cattle as draft animals Dress or Ornament indicating High Status burial with copper awl and ostrich egg shell beads Competition and Warfare

Table 6.3: Evidence for a chiefdom societal structure at Tel Tsaf based on the indicators outlined in Table 6.1

The presence of numerous silos with large storage capacities at Tel Tsaf could be an indicator of an increase in population. However, the fact that the silos are associated with walled, nuclear- family complexes makes this inconclusive. More grain would be required to support a growing population, but the suggested private use of the foodstuffs stored at Tel Tsaf raises some questions about population increase, as the stored goods belonged to a household unit. The silos are a much better indicator of increased productivity from earlier periods. As discussed in Chapter 5, filling the silos in both complexes would require a large amount of grain or other foodstuffs. Combined with evidence of cattle used for traction, the production capacity at Tel Tsaf would have been far greater than in earlier periods.

209

The only evidence of potential central social or ceremonial activity comes from the site level at Tel Tsaf, with regional data currently lacking. There is evidence of feasting in Building Complex I (Ben-Shlomo, et al. 2009), but its scale is unknown. It may be that the residents of the complex held some central position of prestige within the community. The presence of two burials in Building Complex I offer further support. The courtyard burial is loosely associated with ritual evidence in the form of figurines, loose ostrich egg shell beads and cached gazelle horn cores. More convincing evidence of individual status based on dress comes from the burial in Silo 339. The location of the ostrich eggshell beads in in situ rows across the hips indicate that they likely formed a belt worn by the individual rather than a deposit of grave goods. The copper awl was found in the fill associated with the burial. It is the only evidence of copper from the site, and the first evidence of metal-working in the region prior to the Late Chalcolithic. In the Middle Chalcolithic context, both these factors are prime indicators of an individual with higher status. It is arguably the best evidence of the veneration of one single individual in the entire Chalcolithic period, as much of the later evidence comes from caches of artifacts unrelated to individual human remains, or from burial caves with numerous burials, such as Nahal Qanah.

Unfortunately, the lack of secure radiometric dates directly associated with the Building Complex I burials is an issue. Based on a Bayesian analysis of radiocarbon dates (Chapter 2) the occupation of Tel Tsaf likely spans about 600 years between c. 5040 and 4440 cal BC. It is possible the burials date to the very end use of this timespan, or potentially even after abandonment of this part of the site. Garfinkel et al. (2014) acknowledge that the mudbrick paving of the Silo 339 floor is not present in the area directly above the burial (figure 6.1). This indicates that the burial was dug through the floor, removing the mudbricks, signifying that the silo was no longer in use when the burial was deposited.

210

Figure 6.1: Diagram of Silo 339. The mudbrick paving is absent in the central portion of Silo 339 (squared area), however, there were a several large stones over that area (circled) (after Garfinkel, et al. 2014).

Another potential issue with dating the burial is chemical composition of the copper awl. The original metal has completed corroded away, but analysis of the corroded material by portable XRF indicates that it contained approximately 6% tin (Garfinkel, et al. 2014). This confirms that it was not crafted from native copper. The authors note that copper containing tin is not found in Late Chalcolithic or Early Bronze Age contexts, only appearing during the Middle Bronze Age. The lack of a secure date for the burial, combined with the surprising copper composition raises some serious questions. However, the presence of ostrich eggshell beads related to the courtyard burial strongly suggests it is contemporaneous with the silo burial. As previously discussed, the gazelle horn core stashes and the figurines from the courtyard also have parallels in Chalcolithic material culture, and the association of mortuary remains with silos is seen at Early Chalcolithic sites. It seems highly likely that the burials are Chalcolithic in date, despite the chronological uncertainties. But whether they date to Tel Tsaf’s Middle Chalcolithic occupation, or are the result of ceremonial re-use of the site by Late Chalcolithic populations, remains to be determined with certainty.

The limited number of Middle Chalcolithic sites and even more limited data from Abu Hamid II make it impossible to discuss regional groups during the period. There is also no evidence of any territorial boundary markers in the Tel Tsaf or Abu Hamid regions. None of the

211 architectural evidence suggests the organisation and deployment of large labour forces in construction. Garfinkel and his colleagues (2009) have suggested that the grain storage capacity at Tel Tsaf may have necessitated an organised labour force but, as discussed in section 5.2.2, this claim may be overstated. Lastly, there is no evidence of regional competition for resources or warfare during the Middle Chalcolithic.

The evidence for Tel Tsaf meets some the requirements that Levy has suggested signify chiefdom-level organisation in the region, but I believe it is too limited to justify classifying the site as part of a chiefdom society. The data does enrich our understanding of the origins of several Late Chalcolithic archaeological features and, thus, social practices. The largest contributions that Tel Tsaf makes to the chiefdom debate are the evidence for increased production of foodstuffs, accumulation of surplus and, if the burials are Middle Chalcolithic in date, evidence of individuals with presumed rank or status. The evidence of increased production and draft animals in the Middle Chalcolithic appear earlier at Tel Tsaf than previously assumed from evidence elsewhere and could suggest that the use of animal traction was well established and perhaps even widespread by the Late Chalcolithic. As discussed, the burial of an individual with shell beads and a copper awl is good evidence of a person of potential status at Tel Tsaf. The specific correlation of these goods with an individual, as well as similar items found in proximity to the courtyard burial, indicates that those members of society were important to the community, but whether they had ascribed or achieved status remains unknown.

Tel Tsaf highlights some of the issues with applying the chiefdom model to the Levantine Late Chalcolithic. The strongest correlates to Levy’s chiefdoms at Tel Tsaf are the same features considered strong, uncontested evidence for chiefdoms in the Late Chalcolithic. Increased population, ceremonial centres, craft specialisation, increased productivity and dress or goods indicating high-status are elements agreed upon by several scholars, though they may debate their scale and significance. Tel Tsaf lacks evidence for territorial boundaries but does have some indications of site-level privatization and divisions. There is no evidence for deployment of organized labour forces and regional competition or warfare at Tell Tsaf, all features which scholars show less agreement on during the Late Chalcolithic. Our understanding of all these elements is just too sparse to reasonably apply Levy’s assertion of the chiefdom model of social and economic organization in the region.

212

6.3 Alternative Frameworks

Surprisingly, despite the hesitancy of the academic community to accept Levy’s chiefdom model, there have been very few attempts to re-frame the discussion. The exception is Isaac Gilead, who argues instead for the classification of the Chalcolithic as a group of bounded regional cultural complexes in the southern Levant, each distinct and independent (Gilead 1988, 2011). He argues that archaeological period boundaries are biased without acknowledging a basic culture history of the time span (2011), championing culture as the primary unit for social and economic analysis Chalcolithic communities. He defines his cultures on the basis of typo- technological observations and radiocarbon dates. Within his framework, Gilead advocates the existence of clusters of early and late cultures during the Chalcolithic period (figure 6.2).

His early phase is dated to ca. 4600 cal BC, around the end of the Wadi Rabah culture, which he classifies at Late Neolithic (2011). He views the earlier cultures as pre-Ghassulian and representing a Neolithic-Chalcolithic transitional phase. He then dates the later phase to ca. 4200 cal BC and refers to it as the Chalcolithic period. While there is value in acknowledging these regional differences, the broad categories that Gilead promotes suggest greater cohesion of material culture and settlement structures than actually exists. Notably, Gilead lumps the Beersheva sites within the Ghassulian cultural entity, despite differences in settlement patterns, site structure and material culture. In contrast, he bases a Tel Tsaf cultural entity on material from a single site. The assignment of broad cultural entities does not take into account variation between regions and communities in a meaningful way, especially when the resulting entities are broad. Nor does this approach account for the interaction that likely took place between these groups.

213

Figure 6.2: Regional cultural entities as suggested by Gilead (2011). Gilead refers to the ca. 4600 cal BC phase as Neolithic – Chalcolithic Transition and the ca. 4200 cal BC phase as the Chalcolithic. The sites labelled are not specifically mentioned by Gilead, but represent known sites from the general time periods (based on Gilead 2011).

Rowan and Lovell (2011) note that, despite failure of Levy’s model to gain full acceptance, beyond Gilead there have been no real attempts to replace it. Some authors have accepted it with their own modified definitions (Gal, et al. 1999; Gopher and Tsuk 1996), but most researchers still work within a culture-history framework, but less formally than Gilead. Some have suggested treating regional cultures as dynamic and fluid communities instead of bounded cultures could be more effective, or even that such cultures are simply artificial and arbitrary (Bernbeck 2008; Campbell 2007; Philip 2011).

Interpretations of community construction with flexible and changing boundaries have been successful in New World archaeology (Abbott 2000; Sassaman and Rudolphi 2001; Varien and Wilshusen 2002; Wernke 2007; Wills and Leonard 1994) and recent attempts have been made to investigate similar relationships within the Late Neolithic of the Levant (Banning and Gibbs 2010; Gibbs 2008b) and elsewhere in the Near East (Bernbeck 1999, 2008; Campbell 2007).

214

These approaches, largely based on material-culture analysis, focus on understanding the interaction between and transmission of style among communities within a region. By acknowledging that even occasional or indirect interaction can influence how communities viewed themselves and each other, these studies help create a more realistic understanding of the reproduction of culture through material remains. The data for the Middle Chalcolithic period is too sparse at the moment for such an approach, but the framework offers potential for changing how the Late Chalcolithic is discussed.

Another potential, yet unexplored framework for the chalcolithic data is the concept of heterarchy. In 1945, biophysicist McCullough identified that human neuro-structures were not ordered hierarchically, but were a collective of unranked interactions. The most immediate implications for this came in the fields of artificial intelligence and computer design work, but Crumley (1995) introduced the idea as a useful tool for archaeological analysis. She argued that many cultural structures and influences cannot be ranked, or can be ranked a variety of ways. She provides the example of a world renowned spiritual leader who may have no influence over a locally renowned community business leader. In this case, global ranking is very different than local ranking for the individual. She suggests that heterarchical systems are dynamic and can form or reform both stable and unstable power relationships within a society. Factors influencing these shifts can be social, climatic, administrative or commercial, for example.

Levy’s arguments for chiefdom level organization during the Chalcolithic depends on identifying and accepting hierarchy both within sites and between sites. As discussed, the evidence he uses to support his argument is not always interpreted in the same way, and many scholars acknowledge the level of ranking posited by Levy. Crumley’s concept of heterarchy has potential to create a more fluid framework for scholars to work in. Heterarchy does not exclude some level of hierarchy within societies, but rather views such structures as being potentially unstable, or localised. Some aspects of Levy’s chiefdom societies do have compelling evidence, but these aspects need to be seen as indicative of static hierarchy, but rather pieces of a dynamic system of potentially emerging economic, religious or administrative sub-sets of society.

215

Chapter 7 Conclusion

Both the original data provided by micromorphology and bulk sediment analysis from Tel Tsaf, and the detailed discussion of the significance of other data and analyses from the site have contributed significantly to our understanding of the Middle Chalcolithic of the southern Levant. This chapter provides an overview of those contributions beginning with clarifications related to site formation and phase relationships at Tel Tsaf. I will then directly address the three research goals stated in Chapter 1:

1. What are the implications of varied architectural forms during a single phase of occupation?

2. With evidence of silos and animal pens at Tel Tsaf, is it possible to identify meaningful surplus? If so, what are its social and economic implications?

3. How can the results of microscopic sediment analysis provide data that directly contribute to “big-picture” debates regarding the possibility of chiefdoms during the Chalcolithic?

The final section will put the evidence from Tel Tsaf and the Middle Chalcolithic in its greater southern Levantine context during the Chalcolithic period. This section also builds the case for greater acknowledgement of the Early and Middle Chalcolithic developments by Chalcolithic scholars.

7.1 Clarification of Site-Formation Processes

The geoarchaeological assessment presented in Section 3.1 clarifies some of the formation processes during Phases 3 and 4 at Tel Tsaf. The most significant contribution is the identification of disturbed courtyard floors between the mudbrick spill and thick burn levels in the eastern courtyard of Building Complex IV. Garfinkel et al. (2007, 2009) present Room 612 as a defined structure during Phase 4a, and both Rooms 612 and 662 as defined during Phase 4b. The micromorphology evidence demonstrates that the full extent of the two rooms cannot be determined with certainty as samples from the excavated limits of each contained fragments of disturbed courtyard floors within general fill material. Micromorphology was also able to

216 demonstrate that the southern portion of Building Complex IV and the lowest courtyard levels of Building Complex I were composed of well-sorted, water-lain clay and silt. This indicates that later levels were constructed over naturally deposited sediment rather than anthropogenic fill.

7.2 What are the implications of varied architectural forms during a single phase of occupation?

This dissertation presents a detailed discussion of the significance of varied architectural forms found at Tel Tsaf. Most notably, the Phase 3 site layout includes two walled courtyard complexes, one with a single broadroom structure and the other with two circular structures. Building Complex I contained a broadroom building, Room 70, as well as a number of silo features and roasting pits. Its excavators have concluded that this was a nuclear-family residence on the basis of these features. Micromorphological data from Room 70 confirm the presence of multiple floors and show evidence of re-surfacing these levels. That the fill material between floors in Room 70 did not contain significant micro-evidence of grain grinding, cooking activities or lithic production supports its use as indoor domestic space. On the other hand, the courtyard samples highlight at least two levels of charred material over rough-lain floors, as well as scattered flint flakes. Thus, the courtyard samples contain evidence consistent with outdoor activities such as cooking, food processing and, possibly, lithic production and use or retooling.

Building Complex II was also enclosed by a wall around the courtyard, but had two round rooms (230 and 263) along with silos. It did not contain any roasting pits, and had fewer silo features than Building Complex I. Some researchers at Tel Tsaf have concluded that either Building Complex II is a second residential feature (Garfinkel, et al. 2009; Garfinkel and Ben Shlomo 2007; Garfinkel, et al. 2007) or have shied away from discussing its possible function (Ben- Shlomo, et al. 2009). The micromorphological evidence presented here from the interiors of both rooms indicates that they were used to house animals. The lack of evidence for the outdoor activities evident in Building Complex I, such as cooking, suggests a different and probably non- domestic purpose for Building Complex II.

These results indicate a strong correlation between architectural form and function. Room 70 is a traditional broadroom house, similar to those found at Early and Late Chalcolithic sites. Both macro- and micro-scale evidence supports its use as a residential space, rather than as a purely storage or processing area. The presence of silos, ovens, rough-lain floor surfaces and rake-out

217 of charred material indicates that most domestic processing took place in the courtyard. In contrast, the circular structures of Rooms 230 and 263 in Building Complex II have no parallels in southern Levantine village architecture. There is no other specific indication of on-site animal pens during the Neolithic or Early Chalcolithic periods in the region. The appearance of circular buildings acting as animal stabling areas suggests a change in the architectural structuring of communities during or shortly before the Middle Chalcolithic.

The presence of varied architectural forms with specific functions also suggests that the site’s inhabitants made deliberate choices in structuring the village of Tel Tsaf. This is supported by the separation of forms by courtyard walls. Thus far, both broadroom and circular structures have not been found in the same complex. Other differences related to access to rooms, outdoor activities and storage capacity discussed in detail in this dissertation also indicate that Building Complex I served a different function than Building Complex II. The presence of animal pens outside of a residential context suggests the possibility that there may have been a diversified economy at Tel Tsaf, with some community members specialising in animal rearing, and others focused on grain production. A better understanding of the distribution of rectangular and circular features at Tel Tsaf would help us identify specific functional uses of space and the roles they played in community function.

7.3 With evidence of silos and animal pens at Tel Tsaf, is it possible to identify meaningful surplus? If so, what are its social and economic implications?

Garfinkel et al. (2009) outline the potential storage capacity at Tel Tsaf in their article about the silos at the site. Their analysis, including estimated silo volumes, was extensively discussed in Chapter 5. In contrast to their approach, in this dissertation I have explored the possibility that storage of items other than grain, such as lentils, chickpeas or dung-fuel, was a prominent feature of this site. Even with a very modest assumed height of 1 m for each silo, it was evident that the storage capacities in Building Complex I and Building Complex II far exceed the subsistence needs of a nuclear or even extended family. Even if only one silo was actively used for storage at any one time, the capacity is sufficient to accommodate surplus. Whether or not we accept the estimates by Garfinkel et al. (2009), there can be very little doubt that production of storable goods was sufficient to allow households to stockpile more than they needed for their own use.

218

Meanwhile, the animal pens present solid evidence of animal storage on-site at Tel Tsaf, and here I have proposed that they housed pigs. As discussed in Chapter 5, it is very difficult to identify animals as surplus in the archaeological record. It would require much more information than we have at present to identify the details of animal economy at the site. The evidence that cattle served as draft animals may suggest that cattle were no longer as important as a primary food source, perhaps because of a dietary preference for other animals or, more likely, because the production of grain was more important than the production of beef. At this point, the identification of physical storage of animals on site does not allow us to assume that this is evidence of surplus, even though we know that a focus on pigs can greatly increase meat production.

Section 5.3 engages in an in-depth discussion of the relationships among storage, surplus and wealth. The relationship cannot be assumed to be linear, and should not be viewed in strictly material terms. Social storage plays heavily into the transformation of physical surplus into wealth (see section 5.3). If there was no demand within a community for the surplus goods of an individual or household, it may simply have gone to waste. A market economy cannot begin to emerge until individuals or households accumulate surplus that they can exchange for other goods or services within the community. The separation of animal pens from the residential context suggests that there may have been some specialised production at Tel Tsaf. If the pens housed male pigs as stud animals, the owners could have been esteemed within the community for having high-quality stock. Alternatively, the pens may represent specialized pig production and the animals may have constituted physical wealth. Surplus can also create social prestige through mechanisms such as feasting. If surplus food could not be used for obtaining other goods, one could still use it to host other members of the community. The sharing and demonstration of surplus could have raised one’s social status and potentially have created social debt.

There is some evidence of feasting at Tel Tsaf, and it is possible that it played a role in creating social wealth within the community (Ben-Shlomo, et al. 2009). There is also some evidence of ritual goods at Tel Tsaf, but very limited evidence of rare goods such as obsidian and copper (Garfinkel, et al. 2009; Garfinkel and Ben Shlomo 2007; Garfinkel, et al. 2007; Garfinkel, et al. 2014). The Silo 339 burial contains a copper awl, which may be a wealth item or symbol of status, but the context of the burial is unclear. It is likely that the surplus at Tel Tsaf did play an

219 important role in community relations at Tel Tsaf, but, whether it was a social or economic role, or both, remains unclear.

Overall, the evidence from Tel Tsaf presents solid evidence of grain storage on a scale that represents surplus. The evidence of animal pens in a private courtyard also indicates a changing role for animals and animal products during the Middle Chalcolithic. The nature of the social and economic structures at Tel Tsaf are still unclear, but the evidence suggests that relationships between households were based on private ownership of surplus that was likely used to gain either wealth or prestige.

7.4 How can the results of microscopic sediment analysis provide data that directly contributes to big-picture debates regarding the possibility of chiefdoms during the Chalcolithic?

The largest contribution that the sediment analysis makes to the debate surrounding social organization during the Chalcolithic period is its identification of animal pens at the site. As noted, the presence of circular rooms at Tel Tsaf is a new phenomenon in the region. The identification of Rooms 230 and 263 as animal storage areas presents a new social and economic dimension to the Chalcolithic period. It is justifiably assumed that animals play an important role in more diversified economies, which began to appear during the Chalcolithic. Evidence from the Early Chalcolithic suggests that storage was more moderate and potentially shared, suggesting communities with little economic differentiation. The large silos and animal pens at Tel Tsaf, by contrast, suggest that the economy may have been more focused on specialised production. On the other hand, the current lack of evidence for communal buildings, central storage or distribution systems or regional site hierarchy at Tel Tsaf also differentiates it from some of the larger Late Chalcolithic sites.

Bulk-sediment analysis at Tel Tsaf identifies dung as the fuel for the roasting pits in Building Complex I. This, combined with evidence of cattle as draft animals (Hill 2011), indicates that animals were being used for secondary products. Sheep and goat were being exploited for wool and dairy products during the Early Chalcolithic, but the expansion of animal product use during the Chalcolithic is further evidence of specialized economies. Both dung as fuel and animals for traction are phenomena usually associated with evidence emerging during the Late Chalcolithic.

220

Once more, the evidence from Tel Tsaf has demonstrated that these changes were taking place earlier than originally expected.

7.5 Tel Tsaf and the Middle Chalcolithic in Context

As noted in Chapter 2, some researchers suggest that the Early Chalcolithic should be classified as Late Neolithic and some (Bourke 2007; Levy 2007; Lovell 2001, 2007 ) even include Tel Tsaf in the Neolithic on the grounds that its site structure and material culture allegedly have more in common with the Neolithic than with the Late Chalcolithic. Based on the results and discussions of this dissertation, I believe it is more appropriate to view Tel Tsaf and the Middle Chalcolithic as an important phase of Chalcolithic development in the southern Levant. Table 6.3 provides an overview of the evidence from Tel Tsaf in certain key aspects and comparisons it to the Early Chalcolithic and Late Chalcolithic record discussed in Chapters 2 and 6.

221

Early Chalcolithic Tel Tsaf Late Chalcolithic extensive regional variability in most extensively excavated site regional variation in site type, material culture in region; Abu Hamid II likely, size and layout; evidence for Regional context but not conclusively large and small sites in close contemporary proximity (Negev)

habitation - villages; farmsteads habitation - village habitation - villages, hamlets?; cemeteries; communal tombs; Site types cave stashes; ceremonial or cultic sites Rectangular broadroom; Bounded courtyard complexes; Rectangular broadroom possible courtyard enclosures rectangular broadroom; circular associated with (but not Domestic Architecture rooms bounded by) courtyards; chain houses; multi-room complexes; subterranean dwellings no clear courtyard walls but is abutting courtyard complexes streets and alley formed by likely, possibly shared courtyard walls of houses Site organization space between a number of houses Household size nuclear likely nuclear nuclear and extended Private or semi-private? Private - within courtyard Private - within houses modest - bins, pits and jars large silos, capacity for surplus. silos; store jars; storage rooms; Storage associated with houses, no Storage rooms, plaster-coated lined pits and bins; clear courtyard walls, but pits and basins and deep subterranean? potentially shared courtyard cylindrical pits found at Abu space Hamid II animal pens sheep/goat dominant; pig, sheep/goat dominant; pig sheep/goat dominant; pig cattle, fishbone present significant; cattle; limited wild significant (numbers vary by Faunal game region); cattle; limited wild game

wool exploited? (spindle whorls dung as fuel; cattle for traction; dung as fuel; cattle for traction; Secondary Product use and loom weights) likely small scale wool and dairy large-scale wool and dairy; exploitation transport? adult (primary) sub-floor, 2 adult (primary) - near or in predominantly secondary; formal articulated (Ein el Jarba) infant silos cemeteries; communal tombs Burials jar burials , children in storage 2 infant (primary) - one jar in (caves); ossuaries; infant jar cists and feature at Tabaqat al- courtyard, 1 in silo burials Bûma; cemetery (Neve Yam) stone maceheads?; applied burnt horn core cache; violin- copper items (maceheads, figure on jars (Ein el Jarba) shaped figurines; ostrich shell scepters, crowns etc.); gold Ceremonial/Prestige items beads, copper awl ingots; violin-shaped figurines; ivory figures; basalt bowls; burnt horn core cache

Table 7.1: An overview of the archaeological evidence from the Early Chalcolithic, Tel Tsaf and the Late Chalcolithic within the Southern Levant.

In general, village sizes increased from the Early Chalcolithic through to the Late Chalcolithic, with Tel Tsaf fitting into this trend. The nuclear-family households of the Early Chalcolithic sites and Tel Tsaf are still evident in the Late Chalcolithic, but extended family households are suggested by the large multi-room complexes excavated at Teleilat Ghassul (Bourke 2001). It should be noted that there is evidence to support extended-family households at the Yarmoukian site of Sha’ar Hagolan (Garfinkel 2006), so it is not a wholly new phenomenon in the southern Levant, just one that was absent during the Early and Middle Chalcolithic periods.

222

Architecturally, Tel Tsaf shows continued use of broadroom buildings from the Early Chalcolithic as the primary residential architectural style. In the Early and Middle both Chalcolithic phases, these buildings are associated with courtyards containing storage and food- processing activity spaces. In the Early Chalcolithic, the courtyards were likely shared among two or more houses (Banning 2010; Kadowaki 2007), but at Tel Tsaf it appears that the courtyards were part of private walled complexes. The circular rooms found within Building Complex II at Tel Tsaf are a departure from the Early Chalcolithic. Rectangular broadroom houses in bounded courtyards were still common in the Late Chalcolithic, with good examples from Teleilat Ghassul (Koeppel, Mallon, et al. 1940; R. Mallon, et al. 1934; North 1961) and Abu Hamid III (Lovell, et al. 2007), but new domestic forms also appeared. Building complexes with multiple-family dwellings bordering a single courtyard have been uncovered at Tel Te’o (Eisenberg, et al. 2001) and Shiqmim (Levy 1987). The chain houses of the Golan sites (Epstein 1998) represent a unique architectural phenomenon in the region and may have housed extended families. Subterranean domestic space is also new during the Late Chalcolithic, with prime examples found at Giv’at Ha-Oranim (Bourke, et al. 2003), Tel Abu Matar (Perrot 1955) and Bir as-Safadi (Perrot 1984). It is unclear if these were houses, as it has been suggested that they were better suited as storage or livestock pens (Gilead 1987).

Tel Tsaf presents some of the best prehistoric evidence for storage in the Southern Levant. The number of silos and their size demonstrate significant storage potential at the site. Their location within bounded courtyard complexes also tells us that storage was a private, controlled affair. Early Chalcolithic sites exhibit evidence of more modest household-level storage. Paved platforms at Wadi Rabah and al-Basatîn (Kadowaki, et al. 2008) were likely household storage areas. Two stone cists, a clay-lined pit and another platform at Tabaqat al-Bûma also indicate storage on a small scale (Banning 2010). In contrast, Late Chalcolithic storage was generally much more significant. The presence of multiple types of storage in individual houses suggests that often more than one type of material was being stored. Store jars may have held liquids while silos may have contained grains. This indicates that by the Late Chalcolithic households were able to or needed to stock up on more items than previously in the Levant. At Teleilat Ghassul, storage features are most frequently located inside house walls, indicating private control of goods (Koeppel, Mallon, et al. 1940). It has been suggested that, in some instances, there may have been differential access to storage and particular resources. A subterranean

223 storage area directly above a large building at the edge of the site of Shiqmim may indicate controlled access to the area (Rowan and Golden 2009). The same may be true at other sites with substantial subterranean storage, as there would be limited access to goods stored in these areas.

There is no conclusive evidence of animal pens at Early Chalcolithic sites, but animals may have been kept within courtyards or houses. Tel Tsaf presents well-defined animal penning areas, as identify by the micromorphology work presented in this dissertation, apparently separate from actual dwellings. The Late Chalcolithic lacks the same evidence of independent animal penning complexes, but larger courtyards likely provided protection for animals (Banning 2010). The evidence for storage facilities at Tel Tsaf is clear and on a large, private scale. This is a change from small-scale, possibly communal storage on Early Chalcolithic sites. The storage capacity of Tel Tsaf is more in line with that of Late Chalcolithic sites, which largely present private or controlled storage facilities. The number and size of the Tel Tsaf silos and the presence of animal pens separate from family dwellings are features unique to the site.

Faunal collections from the entire Chalcolithic period are dominated by domestic species, with the proportion of wild game decreasing from Early to Late phases. In assemblages of all periods, sheep and goat are dominant, with cattle and pigs forming a significant proportion of assemblages. The mean relative abundance of domestic species in the Chalcolithic shows 47% sheep/goat, 28% cattle and 16% pig (Hill 2010). This figure, based on published faunal data, includes data from Munhata, which is a Late Neolithic and Early Chalcolithic site, and is not broken down by Early, Middle or Late Chalcolithic. Despite this, it provides a useful framework for putting Tel Tsaf in context.

The faunal collection from Tel Tsaf varies significantly from the mean relative abundance for the Chalcolithic. Sheep and goat are still dominant (47%), but cattle only represent 13% in contrast to 28% for the Chalcolithic. Pigs also stray from the mean relative abundance, with 30% as opposed to 16%. The prevalence of pigs may be due to the geographic location, as there is a trend of declining abundance of fauna at sites from the northern Levant to the south (Hill 2010). This is correlated with the amount of rainfall in the region, as pigs require approximately 350 mm of annual rainfall and are rather and are intolerant of arid conditions (Grigson 1995). In fact, pigs are almost entirely absent from faunal assemblages from Shiqmim (Grigson 1987;

224

Whitcher, et al. 1998) and Bir es Safadi (Josien 1955). The only other Chalcolithic site with similar abundances in pig remains is Abu Hamid (c. 24%), also a northern Middle Chalcolithic site in a well-watered environment.

Evidence for secondary product use increases throughout the Chalcolithic period. The presence of loom weights and spindle whorls at Nahal Zehora (Gopher 1993) and al-Basatîn (Kadowaki 2005; Kadowaki, et al. 2008) indicates that wool was exploited during the Early Chalcolithic. Similar evidence of spinning and weaving was found at the Late Chalcolithic site of Gilat (Levy 2006). Faunal analysis also suggests kill-off patterns that are consistent with intensive wool exploitation (Grigson 2006). More generally, Grigson (1987) and Levy (1992) suggest that Late Chalcolithic kill patterns as a whole indicate exploitation of caprines for both wool and milk. Exploitation of animals for dairy products is supported by the appearance and prevalence of churns in Late Chalcolithic collections (Grigson 1998), and the presence of fatty acids indicative of milk discovered through Gas Chromatography and Mass Spectrometry (Burton 2004). Similar evidence was identified at the Early Chalcolithic site of Basatîn when residue from ceramics was analysed (Gregg, et al. 2009). Pathologies on cattle bones from Shiqmim (Grigson 1987, 1995) and Gilat (Levy 2006) suggest that cattle were used as draft animals. There is some iconography to suggest that animals were used for transport during the Late Chalcolithic as well. Animal-shaped churns with cornets on their backs, including a sheep or ram from Gilat and a bull from En Gedi (David Ussishkin 1980), may be representations of animal transporting goods (Grigson 1995).

Despite the prevalence of spindle whorls and loom weights during the Early and Late Chalcolithic, no such evidence has been uncovered at Tel Tsaf. Animal kill patterns do not strongly indicate herd management for wool or milk exploitation of sheep or goat, although it is likely that they were used on a small-scale for these products (Hill 2010). The best potential evidence of wool exploitation is the fragment of limonite from Building Complex IV that may have been used to dye textiles. Tel Tsaf presents good evidence that cattle were used for plowing fields, with indicative pathologies on phalanges (Hill 2010). Previously, it was assumed animals were not used for traction until the Late Chalcolithic, but this evidence pushes the assumed origin of traction back to the Middle Chalcolithic. A final secondary product at Tel Tsaf was animal dung for fuel. The Late Chalcolithic site of Grar presents similar evidence (Katz, et al. 2007). It is likely that dung was exploited as fuel prior the Middle Chalcolithic, but at this

225 point there is no direct evidence. Overall, secondary product use at Tel Tsaf is in line with the Late Chalcolithic, while we lack similar evidence from the Early Chalcolithic.

The burial data from the Early Chalcolithic period is largely restricted to infants. Infant jar burials have been found at Ein Jarba (Kaplan 1969a) and Tel Te’o (Eisenberg, et al. 2001), while at Tabaqat al-Bûma, two stone cists and a clay-lined bin that originally served as storage features ultimately became the burial locations of children (Banning 2010). The submerged site of Neve Yam presents compelling evidence of a cemetery related to but separate from the habitation site (Galili, Rosen, et al. 2009). A concentration of ten stone-lined and pit graves are located south of the main habitation area. Hearths with charred seeds are associated with the cemetery, suggesting some ritual associated with the burials. Mortuary remains within the six excavated graves were articulated, denoting primary burial.

No off-site cemetery has been identified for any potential Middle Chalcolithic site, but Tel Tsaf had four burials located within residential contexts. Two adult burials were excavated in Building Complex I, one located next to Silo 74 in the southwest corner of the complex, and the other within Silo 339 (Garfinkel, et al. 2009; Garfinkel and Ben Shlomo 2007; Garfinkel, et al. 2007)(also see section 2.8.4). No grave goods were directly associated with the burial adjacent to Silo 75. The Silo 339 burial had an ostrich shell belt and copper awl (Garfinkel, et al. 2009; Garfinkel, et al. 2014). Two infant burials were uncovered in Building Complex II (Garfinkel, et al. 2009). An infant jar burial was found in the courtyard to the west of Silo 272 and a flexed, articulated infant was interred in a lower level of the mudbrick of Silo 272.

A fuller discussion of Late Chalcolithic burial practices is found in Chapter 6. As noted there, burial practices were diverse and highly regional, but primarily include multiple secondary interments containing limited grave goods. The mortuary remains at Tel Tsaf are more in line with the Early Chalcolithic practice of primary burials and infant jar burials. The location of all Tel Tsaf burials either in close proximity to or directly within silos or storage features is similar to the situation at Tabaqat al-Bûma. All Tel Tsaf burials are located within the habitation area, a feature also noted at Early Chalcolithic sites, with the exception of Neve Yam. It should be noted that off-site burials are likely to have existed in the Early and Middle Chalcolithic, as the number of burials found on-site are too limited to represent the village population. The presence

226 of non-local ostrich shell beads and a copper awl in the Silo 339 burial at Tel Tsaf are outliers for both Early and Late Chalcolithic burial practices.

Evidence of ritual items or activity is sparse during the Early Chalcolithic. A jar with a figurine appliqué was found at Ein el-Jarba (Kaplan 1969a), but no other figurines have been identified. It now appears possible that maceheads, hitherto considered a hallmark of Late Chalcolithic cultures, had their origin during the Late Neolithic and Early Chalcolithic (Rosenberg 2010). Based on morphology and technology, Rosenberg argues that ground stone artifacts from at least ten Early Chalcolithic sites are actually maceheads rather than loom weights. The maceheads discussed are generally made of white limestone, with a few examples of basalt and other materials. Rosenberg notes that the shaft holes tend to be narrow and wooden handles inserted into them would not be robust enough to effectively support the macehead while striking an object, suggesting a ceremonial role in the vein of Late Chalcolithic copper maceheads.

Most of the evidence of ritual or ceremonial activity in the Middle Chalcolithic comes from the southwest corner of Building Complex I at Tel Tsaf. Two female figurines were found near the burial in that area of the courtyard (Garfinkel, et al. 2007) and a single cache of 32 gazelle horn cores was near this area (Hill 2010). In addition, further horn core caches occurred along the southern wall of the complex. The horn cores were predominantly from male gazelles and often burnt (Hill 2010). Their abundance, combined with a low percentage of other gazelle remains in the faunal collection suggests that these horn cores played some kind of ritual role in the complex. Indeed, the concentration of beads, horn cores, figurines and burial in the same general region of the courtyard indicates that something beyond normal household tasks took place in that area at some point.

The presence of cave caches with abundant items of potential wealth, such as copper, ivory and gold, demonstrates a magnitude of ritual or ceremonial activity unparalleled by the household or site-level evidence from the Early and Middle Chalcolithic periods. This phenomenon is further supported by the ritual sites at Ein Gedi and Gilat. Figurines are more abundant, and some parallels may be drawn between the female figurines at Tel Tsaf and the violin-shaped figurines from Gilat, Peqi’in, Ghassul, Abu Hamid and Pella (Alon and Levy 1990; Commenge, et al. 2006b; Gal, et al. 2011; 1997; Garfinkel, et al. 2007). Gazelle horn core caches similar to those at Tel Tsaf are also present at Gilat, strengthening the argument for their ritualistic nature as they

227 are found in context of other presumed ritual material. As mentioned above, maceheads of stone and copper are considered a hallmark feature of the Late Chalcolithic period, and are predominantly found in residential contexts, with the exception of the caches at Ein Gedi and Gilat (Rosenberg 2010). No maceheads were found at Tel Tsaf.

Based on the above discussion, Tel Tsaf and, therefore, the Middle Chalcolithic as a whole, appear to be distinct from both the Early and Late Chalcolithic. The site exhibits features characteristic of both Early and Late phases, highlighting its position as a transitional site. However, Tel Tsaf demonstrates that changing human and animal relations that we would usually associate with the economic specialization and secondary product use of the Late Chalcolithic co-exist with other elements more characteristic of Late Neolithic communities. This makes a strong case for acknowledging that Tel Tsaf should be considered a Chalcolithic community and academic discussion of the important changes that took place in the southern Levant during the fifth millennium cal B.C. should move away from Ghassulo-centrism .

228

229

References

A.G.Contractors 2013 (webpage). accessed 15/01/2013, www.agcontracting.org.au/pdf/bdcg/pdf.

Abbott, D.R. 2000 Ceramics and Community Organization among the Hohokam. University of Arizona Press, Tucson.

Abramowitch, F. and I. Gelfet 1944 The Arab Economy. Hakkibutz Hameuchad (hebrew), En Harod.

Akkermans, P. M. M. G. and M. Verhoeven 1995 An Image of Complexity - the Burnt Village at Late Neolithic Sabi-Abyad, Syria. American Journal of Archaeology 99(1):5-32.

Al-Weshah, R.A. 2000 Optimal use of irrigation water in the Jordan Valley: a case study. Water Resources Management 14(5):327-338.

Albert, R.M. and D.O. Henry 2004 Herding and agricultural activities in the early neolithic site of Ayn Abu Nukhayla (Wadi Rum, Jordan). The results of phytolith and spherulite analyses. Paléorient 30(2):81-92.

Albert, R.M., R. Shahack-Gross, D. Cabanes, A. Gilboa, S. Lev-Yadun, M. Portillo, I. Sharon, E. Boaretto and S. Weiner 2008 Phytolith-rich layers from the Late Bronze and Iron Ages at Tel Dor (Israel): mode of formation and archaeological significance. Journal of Archaeological Science 35(1):57-75.

Alley, R.B., J. Marotzke, W.D. Nordhaus, J.T. Overpeck, D.M. Peteet, R.A. Pielke Jr., R.T. Pierrehumbert, P.B. Rhines, T.F. Stocker, L.D. Talley, and J.M. Wallace 2003 Abrupt Climate Change. Science 299:2005-2010.

Alon, D. and T.E. Levy 1980 Preliminary Note on the Distribution of Chalcolithic Sites on the Wadi Beer- sheba—Lower Wadi Besor Drainage System. Israel Exploration Journal 30(3/4):140- 147.

1989 The archaeology of cult and the Chalcolithic sanctuary at Gilat. Journal of Mediterranean Archaeology 2(2):163-221.

1990 The Gilat Sanctuary: Its Centrality and Influence in the Southern Levant during the Late 5th-Early 4th Millennium BCE. (Hebrew), 101*-102 (English summary). Eretz- Israel 21((Hebrew)):101-102 (English summary).

230

Amiran, R. 1978 Early Arad: The Chalcolithic Settlement and Early Bronze Age City. Israel Exploration Society, Jerusalem.

1989 The Gilat goddess and the temples of Gilat, En-Gedi and Ai. In L'Urbanisation de la Palestine a l'Age du Bronze Ancien, Bilan et Perspectives des Recherches Actuelles, edited by P. de Miroschedji, pp. 59-71. BAR International Series 527. British Archaeological Reports, Oxford.

Anfinset , N., H. Taha, al-Zawahra and J Yasine 2011 Societies in Transition: Contextualizing Tell el-Mafjar, Jericho. In Culture, Chronology and the Chalcolithic: Theory and Transition, edited by J. Lovell and A. J. Rowan, pp. 97-113. Levant Supplementary Series. vol. 9. Oxbow Books, Oxford.

Araus, J.L., G.A. Slafer and I. Romagosa 1999 Durum wheat and barley yields in antiquity estimated from 13C discrimination of archaeological grains: a case study from the Western Mediterranean Basin. Australian Journal of Plant Physiology 26:345-352.

Araus, J.L., G.A. Slafer, I. Romagosa and M. Molist 2001 Estimated wheat yields during the emergence of agriculture based on the carbon isotope discrimination of grains: evidence from a 10th millennium BP site on the Euphrates. Journal of Archaeological Science 28(4):341-350.

Aurenche, O. 1981 La Maison Orientale. L'architecture du Proche-Orient Ancien des Origines au Milieu du Quatrième Millenaire. Geuthner, Paris.

B., Mark, Kevin F. and David L. 1998 Tell Fendi, a late Chalcolithic settlement in the lower Wadi Ziqlab, Jordan. Annual of the Department of Antiquities of Jordan 42:161-177.

Baird, D. and G. Philip 1992 Preliminary Report on the First (1991) Season of Excavations at Tell ash-Shuna North. Annual of the Department of Antiquities of Jordan 36:71-88.

1994 Preliminary Report on the Third (1993) Season of Excavations at Tell esh-Shuna North. Levant 26:111-133.

Ball, T.B., J.D. Brotherson and J.S. Gardner 1993 A typologic and morphometric study of variation in phytoliths from einkorn wheat (Triticum monococcum). Canadian Journal of Botany 71(9):1182-1192.

231

Ball, T.B., J.S. Gardner and J.D. Brotherson 1996 Identifying Phytoliths Produced by the Inflorescence Bracts of Three Species of Wheat (Triticum monococcum L., T. dicoccon Schrank., and T. aestivum L.) Using Computer-Assisted Image and Statistical Analyses. Journal of archaeological science 23(4):619-632.

Banning, E. B. 2011 Houses, Households, and Changing Society in the Late Neolithic and Chalcolithic of the Southern Levant. Paleorient 36(1):49-87.

Banning, E. B., Mark Blackham and David Lasby 1998 Excavations at WZ 121, a Chalcolithic site at Tubna, in Wadi Ziqlab. Annual of the Department of Antiquities of Jordan 42:141-159.

Banning, E. B. and I. Köhler-Rollefson 1992 Ethnographic lessons for the pastoral past: camp locations and material remains near Beidha, southern Jordan. In Pastoralism in the Levant, Archaeological Materials in Anthropological Perspectives, edited by O. Bar-Yosef and A. Khazanov, pp. 181-201. Monographs in World Archaeology No. 10. Prehistory Press, Madison.

Banning, E.B. 1983 Wadi Ziqlab Survey. In Archaeological Survey in the Mediterranean Area, edited by D. R. Keller and D. W. Rupp. vol. 155. BAR International Series, Oxford.

1992 1992 Excavations at Wadi Ziqlab, al-Kura, Jordan. Canadian Archaeological Association Newsletter 12(2):3-4.

1995 Tubna in Wadi Ziqlab. ACOR Newsletter 7(2):5-6.

1996 Tubna in Wadi Ziqlab IN Archaeology in Jordan. P.M. Bikai and V. Egan (eds.). American Journal of Archaeology 100:517-518.

1997 Excavations at Tubna, Wadi Ziqlab, Jordan. Echos du Monde Classique/Classical Views XLI-n.s. 16(1):75-80.

2001a Book Review--Neolithic and Chalcolithic Pottery of the Southern Levant BY Yosef Garfinkel. Bulletin of the American Schools of Oriental Research 322:79-81.

2002 Consensus and Debate on the Late Neolithic and Chalcolithic of the Southern Levant. Paléorient 28(2):148-155.

2007a Time and Tradition in the Transition from Late Neolithic to Chalcolithic: Summary and Conclusions. Paléorient 33(1):137-142.

2007b Wadi Rabah and Related Assemblages in the Southern Levant: Interpreting the Radiocarbon Evidence. Paléorient 33(1):77-102.

232

2010 Houses, households, and changing society in the Late Neolithic and Chalcolithic of the Southern Levant. Paléorient 36(1):49-87.

Banning, E.B., R. Dods, John J. Field, Ian Kuijt, J. McCorriston, Julian Siggers, H. Ta'ani and J. Triggs 1992 Tabaqat al-Buma: 1990 Excavations at a Kebaran and Late Neolithic Site in Wadi Ziqlab. Annual of the Department of Antiquities of Jordan 36:43-69.

Banning, E.B., R. Dods, John J. Field, S. L. Maltby, J. McCorriston, Steve Monckton, R. Rubenstein and P. Sheppard 1989 Wadi Ziqlab Project 1987: A Preliminary Report. Annual of the Department of Antiquities of Jordan 33:43-58.

Banning, E.B., R. Dods, J. McCorriston, Steve Monckton and P. Sheppard 1987 Report on the Wadi Ziqlab Project 1986 Season of Excavations. Annual of the Department of Antiquities of Jordan 31:321-342.

Banning, E.B. and K. Gibbs 2010 Networking in the Neolithic: Wadi Ziqlab in its regional context. Adumatu (22):7- 14.

Banning, E.B., K. Gibbs, Michael Gregg, Seiji Kadowaki and Steve Rhodes 2002 Excavations at al-Basatîn, Wadi Ziqlab, Jordan. Neo-Lithics 2:13-14.

Banning, E.B., K. Gibbs and S. Kadowaki 2011 Changes in Material Culture at Late Neolithic Tabaqat al-Bûma, in Wadi Ziqlab, northern Jordan. In Culture, Chronology, and the Chalcolithic: Theory and Transition, edited by J. L. Lovell and Y. Rowan, pp. 36-60. Oxbow Books and the Council for British Research in the Levant, Oxford.

Banning, E.B., K. Gibbs and Seiji Kadowaki 2005 Excavations at Late Neolithic al-Basatîn, in Wadi Ziqlab, northern Jordan. Annual of the Department of Antiquities of Jordan 49:229-243.

Banning, E.B. and Mohammad Najjar 1999 Excavations at Tell Rakan, A Neolithic Site in Wadi Ziqlab, Jordan. Neo-Lithics 99(2):1-3.

Bar-Matthews, M., and A. Ayalon 2003 Climatic Conditions in the Eastern Mediterranean During the Last Glacial (60-10 ky bp) and Their Relations to the Upper Palaeolithic in the Levant: Oxygen and Carbon Isotope Systematics of Cave Deposits. In More Than Meets the Eye: Studies on Upper Palaeolithic Diversity in the Near East, edited by N. G.-M. a. A. Belfer-Cohen, pp. 13- 18. Oxbow Books, Oxford.

233

Bar-Matthews, M., A. Ayalon and A. Kaufman 1997 Late Quaternary Paleoclimate in the Eastern Mediterranean Region from Stable Isotope Analysis of Speleothems at Soreq Cave, Israel. Quaternary Research 47:155-168.

Bar-Matthews, M., A. Ayalon, A. Kaufman and G. Wasserburg 1999 The Eastern Mediterranean Paleoclimate as a Reflection of Regional Events, Soreq Cave, Israel. Earth and Planetary Science Letters 166:85-95.

Bar-Yosef, O., A. Belfer-Cohen, A. Goren, I. Hershkovitz, O. Ilan, H.K. Mienis and B Sass 1986 Nawamis and habitation sites near Gebel Gunna, southern Sinai. Israel Exploration Journal:121-167.

Bar-Yosef, O., A. Belfer, A. Goren and P. Smith 1977 The Nawamis near'Ein Ḥuderah (Eastern Sinai). Israel Exploration Journal:65- 88.

Bar Yosef, O. and A. Gopher 1997 An Early Neolithic Village in the Jordan Valley. Part I The Archaeology of Netiv Hagdud. American School of Prehistoric Research, Peabody Museum, Cambridge MA.

Barker, G. 2012 The desert and the sown: Nomad–farmer interactions in the Wadi Faynan, southern Jordan. Journal of Arid Environments 86:82-96.

Bartov, Y., M. Stein, Y. Enzel, A. Agnon, and Z. Reches 2002 Lake Levels and Sequence Stratigraphy of Lake Lisan, the Late Pleistocene Precursor of the Dead Sea. Quaternary Research 57:9-21.

Bartov, Yuval, Mordechai Stein, Yehouda Enzel, Amotz Agnon and Ze'ev Reches 2002 Lake Levels and Sequence Stratigraphy of Lake Lisan, the Late Pleistocene Precursor of the Dead Sea. Quaternary Research 57(1):9-21.

Baruch, U 1999 A New Pollen Diagram from Lake Hula. Vegetational, Climatic and Anthropogenic implications. In Ancient Lakes. Their Cultural and Biological Diversity, edited by H. Kawanabe, G. W. Coulter and A. C. Roosevelt, pp. 75 - 86, G. Rijksuniversiteit, general editor. Kenobi Productions.

Begin, Z. B., A. Ehrlich and Y. Nathan 1974 Lake Lisan, the Pleistocene precursor of the Dead Sea. Geological Survey of Israel Bulletin 63:1-30.

Begin, Z. B., Y. Nathan and A. Ehrlich 1980 Stratigraphy and facies distribution in the Lisan Formation: New evidence from the area south of the Dead Sea, Isreal. Israel Journal of Earth Sciences 29(3):182-189.

234

Ben-Shlomo, D., A.C. Hill and Y. Garfinkel 2009 Feasting Between the Revolutions: Evidence from Chalcolithic Tel Tsaf, Israel. Journal of Mediterranean Archaeology 22:129-150.

Ben-Tor, A. 1975 Two Burial Caves of the Proto-Urban Period at Azor, 1971: The First Season of Excavations at Tell-Yarmuth, 1970 1. Institute of Archaeology, Hebrew University of Jerusalem.

Bernbeck, R. 1999 Structure Strikes Back: Intuitive Meanings of Ceramics from Qale Rostam, Iran. In Material Symbols: Culture and Economy in Prehistory, edited by J. E. Robb, pp. 90- 111. Center for Archaeological Investigations, Southern Illinois University, Carbondale.

2008 An Archaeology of Multisited Communities. In The Archaeology of Mobility: Old World and New World Nomadism, edited by H. Barnard and W. Wendrich, pp. 43-77. Cotsen Institute of Archaeology, University of California, Los Angeles.

Bienert, H.-D. and D. Vieweger 1999 Archaeological Excavations at the Late Neolithic Site of Ash-Shalaf: A Preliminary Report on the 1998 Season. Annual of the Department of Antiquities of Jordan 43:49-67.

Blackham, M. 1999 Teleilat Ghassul: An Appraisal of Robert North's Excavations (1959-1960). Levant 31:19-64.

2002 Modeling Time and Transition in Prehistory: The Jordan Valley Chalcolithic (5500-3500 BC). BAR International Series 1027. Archaeopress, Oxford.

Blanton, R.E. 1994 Houses and households: a comparative study. Springer.

Boserup, E. 1965 The Conditions of Agricultural Growth. Aldine Publishing Company, Chicago.

1981 Population and Technological Change. The University of Chicago Press, Chicago.

Bourdieu, P. 1977 Outline of a Theory of Practice. Cambridge University Press, Cambridge.

Bourke, S.J. 1997a The Urbanisation Process in the South Jordan Valley: Renewed Excavations at Teleilat Ghassul 1994/1995. Studies in the History and Archaeology of Jordan 6:249- 259.

235

1997b The 'pre-Ghassulian' Sequence at Teleilat Ghassul. In The Prehistory of Jordan II: Perspectives from 1997, edited by H. G. K. Gebel, Z. Kafafi and G. Rollefson, pp. 395- 417. ex oriente, Berlin.

2001 The Chalcolithic Period. In The Archaeology of Jordan, edited by B. MacDonald, R. B. Adams and P. Bienkowski, pp. 107-163. Sheffield Academic Press, Sheffield.

2002 The Origins of Social Complexity in the Southern Levant: New Evidence from Teleilat Ghassul, Jordan. Palestine Exploration Quarterly 134(1):2-27.

2007 The Late Neolithic/Early Chalcolithic Transition at Teleilat Ghassul: Context, Chronology and Culture. Paléorient 33(1):15-32.

Bourke, S.J., E. Lawson and U. Zoppi 2004 The Chronology of the Late Neolithic and Early Chalcolithic Periods in the North Jordan Valley: New C14 Determinations from Pella in Jordan. In Radiocarbon and Archaeology: Fourth International Symposium, edited by T. Higham, C. B. Ramsey and C. Owen, pp. 119-125. Oxford University School of Archaeology, Oxford.

Bourke, S.J., Ewan Lawson, J.L. Lovell, Quan Hua, Ugo Zoppi and Michael Barbetti 2001 The Chronology of the Ghassulian Chalcolithic Period in the Southern Levant: New 14C Determinations from Teleilat Ghassul, Jordan. Radiocarbon 43:1217-1222.

Bourke, S.J., J.L. Lovell, R Sparks, P Seaton, L Mairs and J Meadows 2000 Preliminary report on a second and third season of renewed excavations at Teleilat Ghassul by the University of Sydney, 1995/1997. ADAJ 44:37-89.

Bourke, S.J., P. Seaton, J.L. Lovell and L. Mairs 1994 Preliminary Report on a First Season of Renewed Excavations at Teleilat Ghassul by the University of Sydney. Annual of the Department of Antiquities of Jordan 39:31- 64.

Bourke, S.J., R. Sparks, B. McLaren, B. Sawada, L. Mairs, J. Meadows, T. Hikade and W. Reade 2003 Preliminary Report on the University of Sydney's Eighteenth and Nineteenth Seasons of Excavations at Pella (Tabaqat Fahl) in 1996/97. Annual of the Department of Antiquities of Jordan 47:335-388.

Bradley, Raymond S. 1999 Paleoclimatology: Reconstructing Climates of the Quaternary. 2nd ed. Academic Press, San Diego.

Braun, E. 1989 The problem of the apsidal house: new aspects of Early Bronze Age I domestic architecture in Israel, Jordan, and Lebanon. Palestine Exploration Quarterly 121:1-43.

236

2008 Deconstructing Y. Garfinkel's 'Beth Shean Ware' and 'Middle Chalcolithic': a realistic look at two archaeological chimeras. Journal of the Serbian Archaeological Society 24:81-108.

Bray, T.L. 2003 The archaeology and politics of food and feasting in early states and empires. Springer.

Brochier, J.E., P. Villa, M. Giacomarra and A. Tagliacozzo 1992 Shepherds and sediments: geo-ethnoarchaeology of pastoral sites. Journal of Anthropological Archaeology 11(1):47-102.

Broshi, M. 1979 The population of western Palestine in the Roman-Byzantine period. Bulletin of the American Schools of Oriental Research 236:1-10.

Brouchier, J.E., Paola V. and Mario G. 1992 Shepherds and Sediments: Geo-ethnoarchaeology of Pastoral Sites. Journal of Anthropological Archaeology 11:47-102.

Buck, C.E., J.A. Christen and G.N. James 1999 BCal: an On-line Bayesian Radiocarbon calibration tool. Internet Archaeology 7.

Bull, I.D., M.M. Elhmmali, P. Perret, W. Matthews, D.J. Roberts and R.P. Evershed 2005 Biomarker evidence of faecal deposition in archaeological sediments at Çatalhöyük. Inhabiting Çatalhöyük: Reports from the 1995–1999 seasons. McDonald Institute for Archeological Research/British Institute for Archeology at Ankara Monograph:415-420.

Bull, I.D., M.J. Lockheart, M.M. Elhmmali, D.J. Roberts and R.P. Evershed 2002 The origin of faeces by means of biomarker detection. Environment international 27(8):647-654.

Burton, M. 2004 Collapse, continuity, and transformation: Tracking protohistoric social change through ceramic analysis case studies of late 5th–early 4th millennium societies in the southern Levant. , University of California San Diego.

Burton, M. and T.E. Levy 2001 The Chalcolithic radiocarbon record and its use in southern Levantine archaeology. Radiocarbon 43(3):1223-1246.

2011a The end of the Chalcolithic period (4500–3600 BCE) in the northern Negev desert, Israel. Culture, Chronology and the Chalcolithic: Theory and Transition, ed. JL Lovell and YM Rowan. Oxford: Oxbow.

237

2011b The End of the Chalcolithic Period (4500 - 3600 BC) in the Northern Negev Desert, Israel. In Culture, Chronology and the Chalcolithic: Theory and Transition, edited by J. L. Lovell and Y. M. Rowan, pp. 178-191. Levant Supplement Series. Oxbow Books, Oxford.

Byrd, B. F. 1994 Public and Private, Domestic and Corporate - the Emergence of the Southwest Asian Village. American Antiquity 59(4):639-666.

Byrd, B.F. 1994 Public and private, domestic and corporate: The emergence of the Southwest Asian village. American Antiquity 49(4):639-666.

Campbell, S. 2007 Rethinking Halaf Chronologies. Paléorient 33(1):103-136.

Canti, M. G. 1997 An investigation of microscopic calcareous spheruites from herbivore dungs. Journal of Archaeological Science 24(3):219-231.

Canti, M.G. 1997 An Investigation of Microscopic Calcareous Spherulites from Herbivore Dungs. Journal of Archaeological Science 24:219-231.

1998 The Micromorphological Identification of Faecal Spherulites form Archaeological and Modern Materials. Journal of Archaeological Science 25(? ):435-444.

1999 The production and preservation of faecal spherulites: animals, environment and taphonomy. Journal of Archaeological Science 26(3):251-258.

Carneiro, R. L. 1998 What happened at the flashpoint? Conjectures on chiefdom formation at the very moment of conception. In Chiefdoms and Chieftaincy in the Americas, edited by E. M. Redmond, pp. 18-42. University Press of Florida.

Castro, P.V., R.W. Chapman, S. Gili, V. Llull, R. Mico, C. Rihuete, R. Risch and M.E. Sanahuja 1998 EUR 18036—Aguas Project. Palaeoclimatic Reconstruction and the Dynamics of Human Settlement and Land-use in the Area of the Middle Aguas (Almeria), in the South- east of the Iberian Peninsula. . Office for Official Publications of the European Communities., Luxembourg.

Chapman, R. 1981 The emergence of formal disposal areas and the'problem'of megalithic tombs in prehistoric Europe. In The archaeology of death, edited by R. Chapman, I. Kinnes and K. Randsborg, pp. 71-81. Cambridge University Press, Cambridge.

238

Charles, M. 1996 Fodder from dung: the recognition and interpretation of dung-derived plant material from archaeological sites. Environmental Archaeology 1(1):111-122.

Childe, V.G. 1950 The urban revolution. Town Planning Review 21(1):3.

Commenge, C. 2005 The Late Chalcolithic Pottery. Shoham (North), Lod Valley, Israel. Excavations of three Late Chalcolithic Burial Caves:51-97.

Commenge, C., D. Alon, T.E. Levy and E. Kansa 2006a Gilat ceramics: Cognitive dimensions of pottery production. In Archaeology, anthropology and cult: The sanctuary at Gilat (Israel), edited by T. E. Levy, pp. 394- 506. Equinox, London.

Commenge, C., T.E. Levy, D. Alon and E. Kansa 2006b Gilat's figurines: Exploring the social and symbolic dimensions of representation. In Archaeology, anthropology and cult: The sanctuary at Gilat (Israel), edited by T. E. Levy, pp. 730-830. Equinox, London.

Copley, M.S., R. Berstan, A.J. Mukherjee, S.N. Dudd, V. Straker, S. Payne and R.P. Evershed 2005 Dairying in antiquity. III. Evidence from absorbed lipid residues dating to the British Neolithic. Journal of Archaeological Science 32(4):523-546.

Copley, M.S., H.A. Bland, P. Rose, M. Horton and R.P. Evershed 2005 Gas chromatographic, mass spectrometric and stable carbon isotopic investigations of organic residues of plant oils and animal fats employed as illuminants in archaeological lamps from Egypt. Analyst 130(6):860-871.

Courty, M. A., P. Goldberg and R. Macphail 1989 Soils and Micromorphology in Archaeology. Cambridge Manuals in Archaeology. Cambridge University Press, Cambridge.

Courty, M.A., R. I. Macphail and J. Wattez 1991 Soil Micromorphological indicators of pastoralism: with special reference to Arene Candide, Finale Ligure, Italy. Rivista di Studi Liguri 57(1-4):127-150.

Cowgill, G.L. 1975 On Causes and Consequences of Ancient and Modern Population Changes. American Anthropologist 77(3):505-525.

Craig, O.E., J. Chapman, C. Heron, L.H. Willis, L. Bartosiewicz, G. Taylor, A. Whittle and M. Collins 2005 Did the first farmers of central and produce dairy foods? Antiquity 79(306).

239

Crawford, D.J. 1971 Kerkeosiris; an Egyptian Village in the Ptolemaic Period. . Cambridge University, Cambridge.

Cribb, R.L. 1984 Greener pastures: mobility, migration and the pastoral mode of subsistence. Production Pastorale et Societe 14:11-46.

Currid, J.D. 1985 The beehive granaries of ancient Palestine. Zeitschrift des Deutschen Palâstina- Vereins 101(2):97-110.

Das, H. A. and J. Zonderhuis 1964 The Analysis of Electrum Coins. Archaeometry 7(1):90-97.

Dawson, L., T.E. Levy and P. Smith 2003 Evidence of interpersonal violence at the Chalcolithic village of Shiqmim (Israel). International Journal of Osteoarchaeology 13(3):115-119. de Contenson, H. 1966 Notes on the Chronology of Near Eastern Neolithic. Bulletin of the American Schools of Oriental Research:2-6.

De Cupere, B., A. Lentacker, W. Van Neer, M. Waelkens and L. Verslype 2000 Osteological evidence for the draught exploitation of cattle: first applications of a new methodology. International Journal of Osteoarchaeology 10(4):254-267. de Vaux, R. 1970 Palestine during the Neolithic and Chalcolithic Periods. In The Cambridge Ancient History, edited by I. E. S. G. Edwards, C.J. and Hammond, N.G.L., pp. 498-537. Third ed. vol. 1:1. University of Cambridge Press, Cambridge.

Diener, P., E.E. Robkin, E.N. Anderson Jr, H.B. Barclay, P.J. Bertocci, K.L. Brown, P. Centlivres, A. Ghani, L.E. Grivetti and M. Harris 1978 Ecology, Evolution, and the Search for Cultural Origins: The Question of Islamic Pig Prohibition [and Comments and Reply]. Current Anthropology:493-540.

Dietler, M. and B. Hayden (editors) 2001 Feasts: Archaeological and Ethnographic Perpspectives on Food, Politics, and Power. Smithsonian Institution Press, Washington.

Dollfus, G. and Z. Kafafi 1993 Recent Researches at Abu Hamid. Annual of the Department of Antiquities of Jordan 37:241-255.

240

Dollfus, G., Z. Kafafi, J. Rewerski, N. Vailland, E. Coqueugniot, J. Desse and R. Neef 1988 Abu Hamid, an early fourth millennium site in the Jordan Valley. In The Prehistory of Jordan, The State of Research in 1986, edited by A. N. Garrard and H. G. Gebel, pp. 567-601. BAR International Series. vol. 396 (ii). British Archaeological Reports, Oxford.

Dornan, J.L. 2002 Agency and Archaeology: Past, Present, and Future Directions. Journal of Archaeological Method and Theory 9(4):303-329.

Dunand, M. 1973 Fouilled de Byblos V: L'Architecture, ses tombes, le material domestique, des origines a l'avenement urbain. Adrien Maisonneuve, Paris.

Earle, T.K. 1987 Chiefdoms in archaeological and ethnohistorical perspective. Annual review of Anthropology:279-308.

1991 The evolution of chiefdoms. In Chiefdoms: Power, Economy, and Ideology, pp. 1- 15. Cambridge University Press, New York.

Eisenberg, E., A. Gopher and R. Greenberg 2001 Tel Te'o: A Neolithic, Chalcolithic, and Early Bronze Age Site in the Hula Valley. IAA Reports 13. Israel Antiquities Authority, Jerusalem.

Eitam, D. 2009 Cereals in the Ghassulian Culture in Central Israel: Grinding Installations as a Case Study. Israel Exploration Journal:63-79.

Epstein, C. 1977 The Chalcolithic culture of the Golan. The Biblical Archaeologist:57-62.

1978a Golan 1977. Israel Exploration Journal 28:116-119.

1978b A new aspect of Chalcolithic culture. Bulletin of the American Schools of Oriental Research:27-45.

1985 Laden animal figurines from the Chalcolithic period in Palestine. Bulletin of the American Schools of Oriental Research:53-62.

1988 Basalt Pillar Figures from the Golan and the Ḥuleh Region. Israel Exploration Journal:205-223.

1993 Oil production in the Golan Heights during the Chalcolithic Period. Tel Aviv 20:133-146.

1998 The Chalcolithic culture of the Golan. Israel Antiquities Authority.

241

Evershed, R.P., S. Payne, A.G. Sherratt, M.S. Copley, J. Coolidge, D. Urem-Kotsu, K. Kotsakis, M. Özdoğan, A.E. Özdoğan and O. Nieuwenhuyse 2008 Earliest date for milk use in the Near East and southeastern Europe linked to cattle herding. Nature 455(7212):528-531.

Faegri, K and J1 Iversen 1989 Textbook of pollen analysis (4th edition). 4th ed. Wiley, Chichester.

Fahn, A., E. Werker and P. Baas 1986 Wood anatomy and identification of trees and shrubs from Israel and adjacent regions. Israel Academy of Sciences and Humanities.

Fairbairn, A., D. Martinoli, A. Butler and G. Hillman 2007 Wild plant seed storage at Neolithic Çatalhöyük East, Turkey. Vegetation History and Archaeobotany 16(6):467-479.

Feinman, G.M. 1991 Demography, surplus, and inequality: early political formations in highland Mesoamerica. In Chiefdoms: Power, economy and ideology, edited by T. Earle, pp. 229- 262. Cambridge University Press, New York.

Finkelstein, I. and R. Gophna 1993 Settlement, demographic, and economic patterns in the highlands of Palestine in the Chalcolithic and Early Bronze periods and the beginning of urbanism. Bulletin of the American Schools of Oriental Research 289:1-22.

Finkelstein, Israel 1990 A few notes on demographic data from recent generations and ethnoarchaeology. Palestine Exploration Quarterly 122(1):47-52.

Fitzgerald, G.M. 1934 Excavations at Beth-Shan in 1933. Palestine Exploration Fund Quarterly Statement 66:123-134.

1935 The Earliest Pottery of Beth-Shan. The Museum Journal 24:6-7.

Fitzpatrick, E.A. 1993 Soil Microscopy and Micromorphology. John Wiley and Sons, Toronto.

Flannery, K.V. 1969 Origins and ecological effects of early domestication in Iran and the Near East. In The domestication and exploitation of plants and animals, edited by P. Ucko and G. Dimbleby, pp. 73-100. Duckworth, London.

242

1972 The origins of the village as a settlement type in Mesoamerica and the Near East: a comparative study. In Man, Settlement and Urbanism, edited by P. Ucko, R. Tringham and G. Dimbleby, pp. 23-53. Duckworth, London.

1983 Early pig domestication in the fertile crescent: a retrospective look. In The Hilly Flanks and Beyond: Essays on the Prehistory of Southwestern Asia, edited by P. E. L. Smith and T. C. Young, Jr., pp. 163-188. University of Chicago Press, Chicago.

2002 The origins of the village revisited: From nuclear to extended households. American Antiquity 67(3):417-433.

Fox, N.S. 1995 The striped Goddess from Gilat: Implications for the Chalcolithic cult. Israel Exploration Journal:212-225.

Fried, M.H. 1967 The Evolution of Political Society. Random House, New York.

Frumkin, A. 1997 The holocene history of Dead Sea levels. OXFORD MONOGRAPHS ON GEOLOGY AND GEOPHYSICS 36:237-248.

Frumkin, A., I. Carmi, I. Zak and M. Magaritz 1994 Middle Holocene Environmental Change Determined from the Salt Caves of Mount Sedom, Israel. In Late Quaternary Chronology and Paleoclimates of the Eastern Mediterranean, edited by O. Bar-Yosef and R. S. Kra, pp. 315-332. Radiocarbon, Tuscon.

Gal, Z., D. Shalem and H. Smithline 2011 The Peqi ‘in Cave: A Chalcolithic Cemetery in Upper Galilee, Israel. Near Eastern Archaeology 74(4):196-206.

Gal, Z., H. Smithline and D. Shalem 1996 A Chalcolithic burial cave in Peqi'in. Qadmoniot 111:19-24.

1997 A Chalcolithic burial cave in Peqi'in, Upper Galilee. Israel Exploration Journal:145-154.

1999 New iconographic aspects of Chalcolithic art: Preliminary observations on finds from the Peqi'in cave. ‘Atiqot 37:1-16.

Galili, E., V. Eshed, B. Rosen, M.E. Kislev, O. Simchoni, I. Hershkovitz and A. Gopher 2009 Evidence for a Separate Burial Ground at the Submerged Pottery Neolithic Site of Neve-Yam, Israel. Paléorient 35(1):31-46.

243

Galili, E., Baruch Rosen, Vered Eshed, Israel Hershkovitz, Mordechai E. Kislev, Orit Simchoni and A. Gopher 2009 Evidence for a Separate Burial Ground at the Submerged Pottery Neolithic Site of Neve-Yam, Israel. Paléorient 35(1):31-46.

Galili, E., J. Sharvit and A. Shifrony 1998 Neve Yam: Underwater Survey. Excavations and Surveys in Israel 18:35-36.

Galili, E., D. J. Stanley, J. Sharvit and M. WeinsteinEvron 1997 Evidence for earliest olive-oil production in submerged settlements off the Carmel coast, Israel. Journal of Archaeological Science 24(12):1141-1150.

Garfinkel, Y. 1992 The Pottery Assemblages of the Sha'ar Hagolan and Rabah Stages of Munhata (Israel). Cahiers du Centre de Recherche Français de Jérusalem No. 6. Association Paléorient, Paris.

1992a The Pottery Assemblages of the Sha'ar Hagolan and Rabah Stages of Munhata (Israel). Cahiers de Recherche Français de Jerusalem, no. 6. Association Paléorient, Paris.

1993a 'Ali, Tel. In The New Encyclopedia of Archaeological Excavations in the Holy Land, pp. 53-55.

1993b The in Israel. Paléorient 19(1):115-134.

1995 Human and Animal Figurines of Munhata (Israel). Les Cahiers des Missions Archéologiques Françaises en Israel no. 8. Paléorient, Paris.

1999 Neolithic and Chalcolithic Pottery of the Southern Levant. QEDEM 39. The Hebrew University of Jerusalem, Jerusalem.

2004 The Goddess of Sha'ar Hagolan: Excavations at a Neolithic Site in Israel. Israel Exploration Society, Jerusalem.

2006 The Social Organization at Neolithic Sha'ar Hagolan: the Nuclear Family, the Extended Family and the Community. In Domesticating Space: Construction, Community, and Cosmology in the Late Prehistoric Near East, edited by E. B. Banning and M. Chazan, pp. 103-111. ex oriente, Berlin.

Garfinkel, Y., D. Ben-Shlomo and T. Kuperman 2009 Large-scale storage of grain surplus in the sixth millennium BC: the silos of Tel Tsaf. Antiquity 83(320):309-325.

Garfinkel, Y. and D. Ben Shlomo 2007 Tel Tsaf. Israel Exploration Journal 57(2):236-241.

244

Garfinkel, Y., D. Ben Shlomo, M. Freikeman and A. Vered 2007 Tel Tsaf: The 2004-2006 Excavation Seasons. Israel Exploration Journal 57:1- 33.

Garfinkel, Y., I. Carme and J. C. Vogel 1987 Dating of horsebean and lentil seeds from the Pre-Pottery Neolithic B village of Yiftah'el. Israel Exploration Journal 37:40-42.

Garfinkel, Y., F. Klimscha, S. Shalev and D. Rosenberg 2014 The Beginning of Metallurgy in the Southern Levant: A Late 6th Millennium CalBC Copper Awl from Tel Tsaf, Israel. PLoS ONE 9(3):e92591.

Garfinkel, Y. and M. Miller 1999 Sha'ar-ha-Golan, 1998 - Excavations and surveys. Israel Exploration Journal 49(1-2):136-138.

Garfinkel, Y. and M. A. Miller 2002 Sha'ar Hagolan Volume 1: Neolithic Art in Context. Oxbow, Oxford.

Gibbs, K. 2008a Pierced Clay Disks and Late Neolithic Textile Production. In Proceedings of the 5th International Congress on the Archaeology of the , edited by J. M. Córdoba, M. Molist, M. C. Pérez, I. Rubio and S. Martínez, pp. 89-96. vol. II. UAM Ediciones, Madrid.

2008b Understanding Community: A Comparison of Three Late Neolithic Pottery Assemblages from Wadi Ziqlab, Jordan, Unpublished PhD dissertation, Department of Anthropology, University of Toronto, Toronto.

Gibbs, K., S. Kadowaki and E.B. Banning 2006 The Late Neolithic at al-Basatîn in Wadi Ziqlab, Northern Jordan. Antiquity 80(310):http://www.antiquity.ac.uk/ProjGall/gibbs/index.html.

2010 Excavations at al-Basatîn, a Late Neolithic and Early Bronze I Site in Wadi Ziqlab, Northern Jordan. Annual of the Department of Antiquities of Jordan 54:461-476.

Giddens, A. 1984 The Constitution of Society: Outline of the Theory of Structuration. Polity Press, Cambridge.

Gilead, I 2009 The Neolithic–Chalcolithic transition in the Southern Levant: Late sixth-fifth millennium culture history. Transitions in Prehistory: Essays in Honor of O. Bar- Yosef:339-359.

245

Gilead, I and D. Alon 1988 Excavations of protohistoric sites in the Nahal Besor and the Late Neolithc of the northern Negev and Sinai. Mitekufat Ha'even: Journal of the Israel Prehistoric Society 21:109-130.

Gilead, I, O Marder, H Khalaily, P Fabian, Y Abadi and Y Yisrael 2004 The Chalcolithic flint workshop in Beer Sheva: a preliminary report. Mitekufat Ha’even: Journal of the Israel Prehistoric Society 34:245-263.

Gilead, I. 1987 A new look at Chalcolithic Beer-Sheba. The Biblical archaeologist 50(2):110- 117.

1988 The Chalcolithic period in the Levant. Journal of World Prehistory 2:397-443.

1989 Grar: A Chalcolithic Site in the Northern Negev, Israel. Journal of Field Archaeology 16:377-394.

1990 The Neolithic-Chalcolithic transition and the Qatifian of the northern Negev and Sinai. Levant 22:47-63.

1994 The history of the Chalcolithic settlement in the Nahal Beer Sheva Area: the radiocarbon aspect. Bulletin of the American Schools of Oriental Research:1-13.

2007 The Besorian: A Pre-Ghassulian Cultural Entity. Paléorient 33(1):33-50.

2011 Chalcolithic culture history: Ghassulian and other entities in the southern Levant. Culture, Chronology and the Chalcolithic. Theory and Transition. Oxbow Books, Oxford:12-24.

Gilead, I., D.E. Bar-Yosef, A. Ben-Tor, P. Fabian, Y. Goren, C. Grigson, D. Hershman and O. Marder 1995 Grar: A chalcolithic site in the Northern Negev. Ben-Gurion University of the Negev Press.

Goldberg, P. 1983 Applications of Micromorphology in Archaeology. In Soil Micromorpohology, edited by P. B. a. C. P. Murphy, pp. 139-150. A.B. Academic Publishers, Berkhamsted.

1987 The geology and stratigraphy of Shiqmim. In Shiqmim I: Studies Concerning Chalcolithic Societies in the Northern Negev Desert, Israel (1982-1984), edited by T. E. Levy, pp. 35-43. BAR International Series 356. British Archaeological Reports, Oxford.

1992 Micromorphology, Soils, and Archaeological Sites. In Soils in Archaeology: Landscape Evolution and Human Occupation, edited by V. T. Holliday, pp. 145-167. Smithsonian Institution Press, Washington.

246

Goldberg, P. and O. Bar-Yosef 1995 Sedimentary environments of prehistoric sites in Israel and the Southern Levant. Ancient Peoples and Landscapes:29-49.

Goldberg, P. and F. Berna 2010 Micromorphology and context. Quaternary International 214(1-2):56-62.

Goldberg, P., F. Berna and R.I. Macphail 2009 Comment on “DNA from Pre-Clovis Human Coprolites in Oregon, North America”. Science 325(5937):148.

Goldberg, P. and R.I. Macphail 2006 Practical and Theoretical Geoarchaeology. Blackwell Science, Malden.

Goldberg, P., C.E. Miller, S. Schiegl, B. Ligouis, F. Berna, N.J. Conard and L. Wadley 2009 Bedding, hearths, and site maintenance in the Middle of Sibudu Cave, KwaZulu-Natal, South . Archaeological Anthropological Science 1(2):95-22.

Goldberg, Paul 1994 Interpreting Late Quaternary continental sequences in Israel. In Late Quaternary Chronology and Paleoclimates of the Eastern Mediterranean, edited by O. Bar-Yosef and R. S. Kra, pp. 89-102. Department of Geosciences, Univeristy of Arizona, Tucson.

Golden, J. 2009 New light on the development of Chalcolithic metal technology in the Southern Levant. Journal of World Prehistory 22(3):283-300.

Golden, J., T.E. Levy and A. Hauptmann 2001 Recent discoveries concerning chalcolithic metallurgy at Shiqmim, Israel. Journal of Archaeological Science 28(9):951-963.

Goldstein, L. 1981 One-dimensional archaeology and multi-dimensional people: spatial organization and mortuary analysis. The archaeology of death 53:69.

Gopher, A. 1989 The Flint Assemblages of Munhata. Association Paléorient, Paris.

1993 Sixth-Fifth Millennia B.C. Settlements in the Coastal Plain, Israel. Paléorient 19(1):55-63.

1994 Nahal Zehora I and II-1990. Excavations and Surveys in Israel 12:27-28.

1998 Early pottery bearing groups in Israel - The Pottery Neolithic period. In The archaeology of society in the Holy Land, edited by T. E. Levy, pp. 205-225. Leicester University Press, London.

247

Gopher, A., Estelle Orrelle, Noga Blockman, Ran Barkai, Ruth Eyal and Dan Naveh 1996 The Nahal Zehora Village Sites: Pottery Neolithic in the Menashe Hills Israel. In XIII UISPP Congress Proceedings, pp. 465-467. A.B.A.C.O. Edizioni, Forli'.

Gopher, A. and T. Tsuk 1991 Ancient Gold: Rare Finds from the Nahal Qanah Cave. The Israel Museum, Jerusalem.

1996 The Nahal Qanah Cave: Earliest Gold in the Southern Levant. Tel Aviv University, Tel Aviv.

Gopher, Avi and Ram Gophna 1993 Cultures of the Eighth and Seventh Millennia BP in the Southern Levant: A Review for the 1990s. Journal of World Prehistory 7(3):297-353.

Gophna, R. and J. Portugali 1988 Settlement and demographic processes in Israel's coastal plain from the Chalcolithic to the Middle Bronze Age. Bulletin of the American Schools of Oriental Research 269:11-28.

Gophna, R. and S. Sadeh 1988/1989 Excavations at Tel Tsaf: An Early Chalcolithic Site in the Jordan Valley. Tel Aviv 15/16(1):3-36.

Gophna, R. and T. Tsuk 2005 Chalcolithic settlements in the western Samaria foothills. Tel Aviv 2005(1):3-19.

Goren, Y. and P. Fabian 2002 Kissufim Road: A Chalcolithic Mortuary Site. IAA Reports 16. Israel Antiquities Authority, Jerusalem.

Goren, Y. and A. Gopher 1995 The Beginnings of Pottery Production in the Southern Levant: A Model. In The Ceramics Cultural Heritage, edited by P. Vincenzini, pp. 21-28. Monographs in Materials and Society, 2. Techna, Faenza.

Goring-Morris, N. and L. K. Horwitz 2007 Funerals and feasts during the pre-pottery Neolithic B of the Near East. Antiquity 81(314):902-919.

Graham, P. 2014 Archaeobotanical remains from late 6th/early 5th millennium BC Tel Tsaf, Israel. Journal of Archaeological Science 43:105-110.

248

Greenfield, H.J. 1999 The origins of metallurgy: distinguishing stone from metal cut-marks on bones from archaeological sites. Journal of Archaeological Science 26(7):797-808.

2000 The origins of metallurgy in the central Balkans based on the analysis of cut marks on animal bones. . Environmental Archaeology 5(1):93-106.

2013 “The Fall of the House of Flint”: A Zooarchaeological Perspective on the Decline of Chipped Stone Tools for Butchering Animals in the Bronze and Iron Ages of the Southern Levant. Lithic Technology 38(3):161-178.

Gregg, M. W., E.B. Banning, K. Gibbs and G. F. Slater 2009 Subsistence practices and pottery use in Neolithic Jordan: molecular and isotopic evidence. Journal of Archaeological Science 36(4):937-946.

Grigson, C. 1987 Shiqmim: pastoralism and other aspects of animal management in the Chalcolithic of the northern Negev. In Shiqmim I: studies concerning Chalcolithic societies in the northern Negev Desert, Israel (1982-1984), edited by T. E. Levy, pp. 1982-1984. vol. 356. BAR, Oxford.

1995 Plough and pasture in the early economy of the southern Levant. In The Archaeology of Society in the Holy Land, edited by T. E. Levy, pp. 245-268. Facts on File, New York.

1998 Plough and pasture in the early economy of the southern Levant. . In The Archaeology of society in the Holy Land, edited by T. E. Levy, pp. 245-268. Leicester University Press, London.

2006 Farming? Feasting? Herding? Large mammals from the Chalcolithic of Gilat. In Archaeology, Anthropology, and Cult: The Gilat sanctuary., edited by T. E. Levy and D. Alon. Leicester University Press, London.

2007 Culture, ecology, and pigs from the 5th to the 3rd millennium bc around the Fertile Crescent. In Pigs and humans: 10,000 years of interaction, edited by U. Albarella, K. Dobney, A. Ervynck and P. Rowley-Conwy, pp. 83-108. Oxford University Press, New York.

Gumerman, I.G. 1997 Food and complex societies. Journal of Archaeological Method and Theory 4(2):105-139.

Gur-Arieh, S., E. Mintz, E. Boaretto and R. Shahack-Gross 2013 An ethnoarchaeological study of cooking installations in rural Uzbekistan: development of a new method for identification of fuel sources. Journal of Archaeological Science 40(12):4331-4347.

249

Gustavson-Gaube, C. 1985 Tell esh-Shuna North 1984: a preliminary report. Annual of the Department of Antiquities of Jordan 29:43-88.

1986 Tell Esh-Shuna North 1985: A Preliminary Report. Annual of the Department of Antiquities of Jordan 30:69-113.

Haas, J. 1982 The evolution of the prehistoric state. Columbia University Press New York.

Haber, A. and T. Dayan 2004 Analyzing the Process of Domestication: Hagoshrim as a Case Study. Journal of Archaeological Science 31:1587-1601.

Hadjikoumis, A. 2012 Traditional pig herding practices in southwest Iberia: Questions of scale and zooarchaeological implications. Journal of Anthropological Archaeology 31(3):353-364.

Hald, M.M. and M. Charles 2008 Storage of crops during the fourth and third millennia BC at the settlement mound of Tell Brak, northeast Syria. Vegetation history and archaeobotany 17(1):35-41.

Hamilakis, Y. and E. Konsolaki 2004 Pigs for the gods: Burnt animal sacrifices as embodied rituals at a Mycenaean sanctuary. Oxford Journal of Archaeology 23(2):135-151.

Harvey, E.L. and D.Q. Fuller 2005 Investigating crop processing using phytolith analysis: the example of rice and millets. Journal of Archaeological Science 32(5):739-752.

Hauptmann, A. 2007 The archaeometallurgy of copper: Evidence from Faynan, Jordan. Springer.

Hayden, B. 1995 The Emergence of Prestige Technologies and Pottery. In The Emergence of Pottery: Technology and Innovation in Ancient Societies, edited by W. K. Barnett and J. W. Hoopes, pp. 257-265. Smithsonian Institution Press, Washington.

Helwing, B. 2003 Feasts as a Social Dynamic in Prehistoric : Three Case Studies from Syria and Anatolia. Paléorient 29(2):63-85.

Hennessy, J. B. 1982 Teleilat Ghassul: Its place in the archaeology of Jordan. In Studies in the History and Archaeology of Jordan, edited by A. Hadidi. vol. I. Department of Antiquities of Jordan, Amman.

250

Hennessy, J.B. 1969 Preliminary Report on a First Season of Excavations at Teleilat Ghassul. Levant 1:1-24.

1982 Teleilat Ghassul and its Place in the Archaeology of Jordan. Studies in the History and Archaeology of Jordan 1:55-58.

1989 Ghassul, Teleilat. In Archaeology of Jordan 2, edited by D. Homes-Fredericq, pp. 230-241. Peeters, Leuven.

Hesse, B. 1990 Pig lovers and pig haters: Patterns of Palestinian pork production. Journal of Ethnobiology 10(2):195-225.

Hesse, B. and P. Wapnish 1997 Can Pig Remains Be Used for Ethnic Diagnosis in the Ancient Near East? In The : constructing the past, interpreting the present., edited by N. A. Silberman and D. Small, pp. 238-270. Sheffield Academic Press, Sheffield.

Hill, A.C. 2011 Specialized Pastoralism and Social Stratification---Analysis of the Fauna from Chalcolithic Tel Tsaf, Israel, Anthropology, University of Connecticut.

Hillier, B. and J. Hanson 1984 The social logic of space 1. Cambridge university press Cambridge.

Hole, F. 1991 Middle Khabur settlement and agriculture in the Ninevite 5 period. Bulletin of the Canadian Society for Mesopotamian Studies 21:17-29.

Hubbard, E.M. 2010 Livestock and people in a Middle Chalcolithic settlement: a micromorphological investigation from Tel Tsaf, Israel. Antiquity 84:1123-1134.

Ibrahim, M.M. 1984 Sahab. Archiv fur Orientforschung 29:256-260.

Jardé, A. 1928 Les céréales dans l'antiquié grecque. de Brocard, Paris.

Joffe, A. H. 2003 Identity/Crisis. Archaeological Dialogues 10(1):77-95.

Joffe, A. H. and J. P. Dessel 1995a Redefining Chronology and Terminology for the Chalcolithic of the Southern Levant. Current Anthropology 36(3):507-518.

251

1995b Redefining Chronology and Terminology for the Southern Levant. Current Anthropology 36(3):507-518.

Joffe, A.H. 2003 Slouching toward Beersheva: Chalcolithic mortuary practices in local and regional context. The Near East in the southwest: Essays in Honor of William G. Dever, Annual of the American Schools of Oriental Research 58:45-67.

Joffe, A.H., J.P. Dessel and R.S. Hallote 2001 The "Gilat Woman" - Female iconography, chalcolithic cult, and the end of southern Levantine prehistory. Near Eastern Archaeology 64(1-2):8-23.

Josien, T. 1955 La faune chalcolithique des gisements palestiniens de Bir es-Safadi et Bir Abou Matar. Israel Exploration Journal 5(4):246-256.

Kadowaki, S. 2005 Designs and Production Technology of Sickle Elements in Late Neolithic Wadi Ziqlab, Northern Jordan. Paléorient 31(2):69-85.

2007 Changing Community Life at a Late Neolithic Farmstead: Built Environments and the Use of Space at Tabaqat al-Bûma in Wadi Ziqlab, Northern Jordan. Unpublished PhD Dissertation, University of Toronto.

Kadowaki, S., K. Gibbs, A. Allentuck and E.B. Banning 2008 Late Neolithic Settlement in Wadi Ziqlab, Jordan: al-Basatîn. Paléorient 34(1):105-129.

Kafafi, Z. 2001 Chronological problems of the sixth and fifth millennia B.C. in Jordan. In Beyond Tools: Redefining the PPN Lithic Assemblages of the Levant, edited by C. I., L. C., Z. D. and B. P., pp. 353-361. Ex oriente, Berlin.

2011 Ghrubba: Ware or Culture? In Culture, Chronology and the Chalcolithic: Theory and Transition, edited by J. Lovell and Y. M. Rowan, pp. 12-24. Levant Supplementary Series. Oxbow Books, Oxford.

Kaplan, J. 1958 Excavations at Wadi Rabah. Israel Exploration Journal 8:149-160.

1958b Excavation at Teluliot Batashi in the Vale of Sorek. Eretz-Israel 5:83-84.

1959 The Neolithic Pottery of Palestine. Bulletin of the American Schools of Oriental Research 156:15-22.

1969a 'Ein el Jarba: Chalcolithic Remains in the Plain of Esdraelon. Bulletin of the American Schools of Oriental Research 194:2-39.

252

1969b 'Ein el Jarba: Chalcolithic Remains in the Plain of Esdraelon. Bulletin of the American Schools of Oriental Research (194):2-39.

Katz, O, I Gilead, P. Bar and R. Shahack-Gross 2007 Chalcolithic agricultural life at Grar, northern Negev, Israel: dry farmed cereals and dung-fueled fires. Paléorient 33(2):101-116.

Kenyon, K.M. and T.A. Holland 1960 Excavations at Jericho 1. British school of archaeology in Jerusalem.

Kerner, S. 2001 Das Chalkolithikum in der südlichen Levante. Orient-Archäologie Band 8. VML, Rahden.

2010 Craft specialisation and its relation with social organisation in the late 6th to early 4th millennium BCE of the southern Levant. Paléorient 36(1):179-198.

Kerner, S., R. Bernbeck, R. Lamprichs and G. Lehmann 1992 Excavations in Abu Snesleh: Middle Bronze Age and Chalcolithic architecture in central Jordan. The Near East in Antiquity III:43-54.

Khalaily, H and O Marder 2010 Prehistoric investigations along the cross-Israel highway: state of research. Atiqot 64:5-23.

Klinger, Y., J.P. Avouac, D. Bourles and N. Tisnerat 2003 Alluvial deposition and lake-level fluctuations forced by Late Quaternary climate change: the Dead Sea case example. Sedimentary Geology 162(1):119-139.

Koeppel, R., R. Mallon and R. Neuville 1940 Teleilat Ghassul II. Pontifical Biblical Institute, Rome.

Koeppel, R., H. Senès, J. W. Murphy and G. S. Mahan 1940 Teleilat Ghassul I: Compte rendu des fouilles de l'Institut biblique pontifical 1929-1932. Scripta Pontificii Instituti. Pontifical Biblical Institute, Rome.

Kramer, C. 1982 Village Ethnoarchaeology. Academic Press, New York.

Kristiansen, K. 1991 Chiefdoms, states, and systems of social evolution. Cambridge University Press.

Kuijt, I. 2008 Demography and storage systems during the southern Levantine Neolithic demographic transition. In The Neolithic demographic transition and its consequences, edited by J. P. Bocquet-Appel and O. Bar Yosef, pp. 287-313. Springer.

253

Kuijt, I. and B. Finlayson 2009 Evidence for food storage and predomestication granaries 11,000 years ago in the Jordan Valley. Proceedings of the National Academy of Sciences 106(27):10966-10970.

Lancelotti, C. and M. Madella 2012 The ‘invisible’product: developing markers for identifying dung in archaeological contexts. Journal of Archaeological Science 39(4):953-963.

Larson, G., K. Dobney, U. Albarella, M. Y. Fang, E. Matisoo-Smith, J. Robins, S. Lowden, H. Finlayson, T. Brand, E. Willerslev, P. Rowley-Conwy, L. Andersson and A. Cooper 2005 Worldwide phylogeography of wild boar reveals multiple centers of pig domestication. Science 307(5715):1618-1621.

Lehmann, G., R. Lamprichs, S. Kerner and R. Bernbeck 1991 The 1990 excavations at Abu Snesleh: preliminary report. In Annual of the Department of Antiquities of Jordan.

LeonCookSilos 2013 Typical grain bulk densities and angles of repose (webpage). accessed on 15/01/13, http://www.leoncooksilos.com.au/.

Levy, T. E., M. M. Burton and Y. M. Rowan 2006 Chalcolithic hamlet excavations near Shiqmim, Negev Desert, Israel. Journal of Field Archaeology 31(1):41-60.

Levy, T.E. 1983 The Emergence of Specialized Pastoralism in the Southern Levant. World Archaeology 15(1):15-36.

1986a Social archaeology and the Chalcolithic period: explaining social organizational change during the 4th millennium in Israel. Michmanim 3:5-20.

1986b The Chalcolithic period: archaeological sources for the . Biblical Archaeologist June:82-108.

1987 Shiqmim I: studies concerning Chalcolithic societies in the northern Negev Desert, Israel (1982-1984). 356. BAR.

1992 Transhumance, Subsistence, and Social Evolution in the Northern Negev Desert. In Pastoralism in the Levant: Archaeological Materials in Anthropological Perspectives, edited by O. Bar-Yosef and A. Khazanov. Monographs in World Archaeology vol. 10. Prehistory Press, Madison Wisconsin.

1993 Production, Space and Social Change in Protohistoric Palestine. In Spatial Boundaries and Social Dynamics, edited by A. Holl and T. E. Levy, pp. 63-81. Ethnoarchaeological Series. International Monographs in Prehistory, Ann Arbor.

254

1995 Cult, Metallurgy and Rank Societies--Chalcolithic Period (ca. 4500-3500 BCE). In The Archaeology of Society in the Holy Land, edited by T. E. Levy, pp. 226-243. Leicester University Press, London.

1998 Cult, metallurgy and rank societies - Chalcolithic period (ca. 4500-3500 BCE). In The archaeoloyg of society in the Holy Land, edited by T. E. Levy, pp. 226-244. Leicester University Press, London.

2006 Archaeology, Anthropology and Cult: the Sactuary at Gilat, Israel. Equinox, London.

Levy, T.E. and D. Alon 1979 A preliminary note on the Chalcolithic cemeteries at Shiqmim, northern Negev, .117-114:מתקופת האבן .Israel

1982 The Chalcolithic mortuary site near Mezad Aluf, northern Negev Desert: A preliminary study. Bulletin of the American Schools of Oriental Research:37-59.

1985a The Chalcolithic Mortuary Site near Meẕad Aluf, Northern Negev Desert: Third Preliminary Report, 1982 Season. Bulletin of the American Schools of Oriental Research. Supplementary Studies:121-135.

1985b Shiqmim: A Chalcolithic village and mortuary centre in the northern Negev. Paléorient:71-83.

1985c An Anthropomorphic Statuette from Shiqmim. Atiqot 17:187-189.

1987a Excavations in Shiqmim Cemetery 3: Final report on the 1982 excavations. In Shiqmim I, edited by T. E. Levy, pp. 333-355. BAR International Series Archaeopress, Oxford.

1987b Excavations in the Shiqmim village. In Shiqmim I: Studies Concerning Chalcolithic Societies in the Northern Negev Desert, Israel (1982-1984), edited by T. E. Levy, pp. 153-218. BAR International Series 356. British Archaeological Reports, Oxford.

1987c Settlement patterns along the Nahal Beersheva–lower Nahal Besor: Models of subsistence in the northern Negev. Shiqmim I, BAR International Series 356:45-138.

Levy, T.E., D. Alon, P. Goldberg, C. Grigson, P Smith and Jane E Buikstra 1994 Protohistoric investigations at the Shiqmim Chalcolithic village and cemetery: Interim report on the 1988 season. Annual of the American Schools of Oriental Research 51:87-106.

255

Levy, T.E., M. Burton and Y.M. Rowan 2006 Chalcolithic hamlet excavations near Shiqmim, Negev desert, Israel. Journal of field archaeology 31(1):41-60.

Levy, T.E., C. Grigson, J.E. Buikstra, D. Alon, P. Smith, S. Shatev, P. Goldberg, A. Holl, S.A. Rosen and S.B. Itzhak 1991 Protohistoric investigations at the Shiqmim Chalcolithic village and cemetery: Interim report on the 1987 season. Bulletin of the American Schools of Oriental Research. Supplementary Studies:29-46.

Levy, T.E. and S. Shalev 1989 Prehistoric Metalworking in the Southern Levant - Archaeometallurgical and Social Perspectives. World Archaeology 20(3):352-372.

Litt, T., C. Ohlwein, F. Neumann, A. Hense and M. Stein 2012 Holocene climate variability in the Levant from the Dead Sea pollen record. Quaternary Science Reviews 49:95-105.

Lloyd, S and F. Safar 1945 Tell Hassuna. Journal of Near Eastern Studies 4:255-289.

Love, S. 2012 The Geoarchaeology of Mudbricks in Architecture: A Methodological Study from Çatalhöyük, Turkey. Geoarchaeology 27(2):140-156.

Lovell, J.L. 2001a The Late Neolithic and Chalcolithic Periods in the Southern Levant: New Data from the Site of Teleilat Ghassul, Jordan. Bar International Series 974. Archaeopress, Oxford.

2001b Review of Yosef Garfinkel "Neolithic and Chalcolithic Pottery of the Southern Levant". Paléorient 26(1):162-163.

2001c An Indigeneous Development: Basil Hennessy and the Late Neolithic at Teleilat Ghassul. In Australians in Jordan: 50 Years of Excavation and Survey, edited by A. Walmsley, pp. 31-42. University of Sydney, Sydney.

2010 Community is Cult, Cult is Community: Weaving the Web of Meanings for the Chalcolithic. Paléorient 36(1):103-122.

Lovell, J.L., G. Dollfus and Z. Kafafi 2004 The Middle Phases at Abu Hamid and the Wadi Rabah Horizon. Studies in the History and Archaeology of Jordan 8:263-274.

2007 The Ceramics of the Late Neolithic and Chalcolithic: Abu Hamid and the Burnished Tradition. Paléorient 33(1):51-76.

256

Lovell, J.L., Z. Kafafi and G. Dollfus 1997 A Preliminary Note on the Ceramics from the Basal Levels of Abu Hamid. In The Prehistory of Jordan, II. Perspectives from 1997, edited by H. G. K. Gebel, Z. Kafafi and G. O. Rollefson, pp. 361-370. Ex Oriente, Berlin.

Macphail, R.I., M.A. Courty, J. Hather and J. Wattez 1997 The soil micromorphological evidence of domestic occupation and stabling activities. In Arene Candide: a Functional and Environmental Assessment of the Holocene Sequence, edited by R. Maggi, pp. 53-88. Instituto Italiano di Paleontologia Umana, Roma.

Madella, M. 2003 Investigating agriculture and environment in South Asia: present and future contributions from opal phytoliths. 2003) Indus Ethnobiology: New Perspectives from the Field. Lanham: Lexington Books:199-249.

Maher, L., E. B. Banning and M. Chazan 2011 Oasis or Mirage? Assessing the Role of Abrupt Climate Change in the Prehistory of the Southern Levant. Cambridge Archaeological Journal 21(1):1-29.

Mallon, A., R. Koeppel and R. Neuville 1934 Teleilat Ghassul I, 1929-32. Pontifical Biblical Institute, Rome.

Mallon, R., R. Koeppel and R. Neuville 1934 Teleilat Ghassul I. Pontifical Biblical Institute, Rome.

Manning, S.W. 2007a Beyond Dates to Chronology: Rethinking the Neolithic-Chalcolithic Levant. Paléorient 33(1):5-10.

2007b Why radiocarbon dating 1200 BC is difficult: a sidelight on dating the end of the Late Bronze Age and the contrarian contribution. Scripta Mediterranea 27(28):53-80.

Marshall, F. B. 1986 Aspects of the Advent of Pastoral Economies in East Africa. Ph.D., University of California, Berkeley.

Matthews, W. 1995 Micromorphological characterisation and interpretation of occupation deposits and microstratigraphic sequences at Abu Salabikh, Southern Iraq. Archaeological Sediments and Soils: Analysis, Interpretation and Management. Institute of Archaeology, University College London:41-74.

1996 Microstratigraphy, micromorphology and sampling report.

257

2003 Microstratigraphic sequences: Indications of uses and concepts of space. In Excavations at Tel Brak: Exploring an upper Mesopotamian regional centre, edited by R. J. Matthews, pp. 377-388. vol. 4. McDonald Institute for Archaeological Research and British School of Archaeology in Iraq, Cambridge.

2005 Micromorphological and micrstratigraphic traces of uses and concepts of space, edited by I. Hodder. The Çatalhöyük Project. McDonal Institute for Archaeological Research and British Institute of Archaeology at Ankara, Cambridge.

2010 Geoarchaeology and taphonomy of plant remains and microarchaeological residues in early urban environments in the Ancient Near East. Quaternary International 214(1–2):98-113.

Matthews, W., C. A. I. French, T. Lawrence, D. F. Cutler and M. K. Jones 1997 Microstratigraphic traces of site formation processes and human activities. World Archaeology 29(2):281-308.

Matthews, W., J.N. Postgate, S. Pane, M.P. Charles and K. Dobney 1994 The imprint of living in a Mesopoatamian city: questions and answers. In Whither environmental archaeology?, edited by R. Luff and P. R. Conwy, pp. 171-212. Oxbow Books, Oxford.

Matthews, W., J. Wiles and M.J. Almond 2006 Micromorphology and microanalysis of architectural surface materials and and residues: investigation of source materials and the life-cycle of buildings.

Mauss, M. 1967 The Gift (1925) Trans. Ian Cunnison. Norton, New York.

Mazar, A. 2001 The significance of the granary building at Beth Yerah. In Studies in the archaeology of Israel and the neighboring lands in memory of Douglas L. Esse edited by S. R. Wolff, pp. 447-464. vol. 59. Oriental Institute of the University of Chicago, Chicago

McDonnell, G. 1983 Tap slags and hearth bottoms or how to identify slags. Current archaeology no. 86 8(3):81-83.

McNicoll, A., P.C. Edwards and J.W. Hanbury-Tenison 1992 Pella in Jordan 2: The Second Interim Report of the Joint University of Sydney and College of Wooster Excavations at Pella 1982-1985 1. Meditarch.

McNicoll, A. and R.H. Smith 1982 Pella in Jordan 1: an interim report on the joint University of Sydney and the College of Wooster excavations at Pella 1979-1981 1. Australian National Gallery.

258

Mellaart, J. 1962 Preliminary Report of the Archaeological Survey in the Yarmouk and Jordan Valley. Annual of the Department of Antiquities of Jordan 6-7:126-157.

1966 The Chalcolithic and Early Bronze Ages in the Near East and Anatolia. Khayats, Beirut.

Merpert, N.Y. and R.M. Munchaev 1987 The earliest levels at Yarim Tepe I and Yarim Tepe II in northern Iraq. Iraq:1-36.

Milevski, I., E. Braun and D. Varga 2012 A newly-discovered Early Bronze Age settlement and silo complex at Amaziya, Israel. Antiquity 86(331).

Moore, P.D., J. A. Webb and M.E. Collison 1991 Pollen analysis. Blackwell scientific publications.

Neev, D., and K.O. Emery 1967 The Dead Sea: depositional processes and environments of evaporites. Geological Survey of Israel Bulletin 41:1-147.

Neev, D. and JK Hall 1977 Climatic fluctuations during the Holocene as reflected by the Dead Sea levels. Proceedings of the International conference on terminal lakes.

Nelson, B.A. 1995 Complexity, hierarchy, and scale: a controlled comparison between Chaco Canyon, New Mexico, and La Quemada, Zacatecas. American Antiquity:597-618.

Nelson, S.M. 1998 Introduction: pigs in prehistory. In Ancestors for the pigs: pigs in prehistory, edited by S. M. Nelson. Penn Press, Philadelphia.

Neuville, R. 1930 La necropole megalithique d'el-'Adeimeh (Transjordanie). Biblica:249-265.

North, R. 1961 Ghassul 1960 Excavation Report. Pontifical Biblical Institute, Rome.

Oren, D. 1973 Notes and News: Bir el-Abd (Northern Sinai. Israel Exploration Journal 23:112- 113.

Orni, E. and E. Efrat 1971 Geography of Israel. Geography of Israel. New revised edition.

259

Ottosson, M. 1980 Uppsala Studies in Ancient Mediterranean and Near Eastern Civilizations. Universitet.

Palmer, C, H Smith and P Daly 2007 Ethnoarchaeology. In Archaeology and Desertification. The Wadi Faynan Landscape Survey, Southern Jordan., edited by G. Bakes, pp. 369-395. Oxford Books, Oxford.

Pauketat, Timothy R., L.S. Kelly, G.J. Fritz, N.H. Lopinot, S. Elias and E. Hargrave 2002 The residues of feasting and public ritual at Early Cahokia. American Antiquity 67(2):257-279.

Payne, S. and G. Bull 1988 Components of variation in measurements of pig bones and teeth, and the use of measurements to distinguish wild from domestic pig remains. Archaeozoologia 2(1):2.

Pearsall, D.M. 2000 Paleoethnobotany. A handbook of procedures. San Diego, etc.: Second Edition. Academic Press.

Peebles, C.S. and S.M. Kus 1977 Some archaeological correlates of ranked societies. American Antiquity:421-448.

Perrot, J. 1955 The Excavations at Tell Abu Matar, near . Israel Exploration Journal 5(3):167-189.

1959 Bir es-Safadi. Israel Exploration Journal 9:141-142.

1963 Hazorea, Tell Turmus, Ben Shemen. Revue Biblique 70:559.

1966 Munhata (Kh. Minha). Israel Exploration Journal 16:269-271.

1967 Munhata. Bible et Terre Sainte 93:4-16.

1968 Palestine-Syria-Cilicia. In Courses Toward Urban Life, edited by R. J. a. W. Braidwood, G.R., pp. 147-164. Viking Fund Publications in Anthropology, S. Tax, general editor. Aldine Publishing Company, Chicago.

1984 Structures d'habitat, mode de vie et environnement. Les villages souterrains des pasteurs de Beersheva, dans le sud d'Israel, au IVe millenaire avant l'ere chretienne. Paléorient 10(1):75-96.

Perrot, J., D. Ladiray and D. Ussishkin 1980 Tombes à ossuaires de la région côtière palestinienne au IVe millénaire avant l'ère chrétienne 1. Association Paléorient.

260

Philip, G. 2011 The Later Prehistory of the Southern Levant: Issues of Practice and Context. In Culture, Chronology and the Chalcolithic: Theory and Transition, edited by J. L. a. R. Lovell, Y.M., pp. 192-209. Levant Supplementary Series. Oxbow Books, Oxford.

Philip, G. and D. Baird 1993 Preliminary Report on the Second (1992) Season of Excavations at Tell esh- Shuna North. Levant 25:13-36.

Piperno, D.R. 1988 Phytolith Analysis: An Archaeological and Geological Perspective. Academic Press, San Diego.

2006 Phytoliths: a comprehensive guide for archaeologists and paleoecologists. Rowman Altamira.

Polanyi, K. 1944 The great transformation: Economic and political origins of our time. Rinehart, New York.

Powers-Jones, A.H. 1994 The use of phytolith analysis in the interpretation of archaeological deposits: an Outer Hebridean example. Whither environmental archaeology 38:41-49.

Prausnitz, M.W. 1977 The Pottery at Newe Yam. Eretz-Israel 13:272-276.

Rambeau, C. 2010 Palaeoenvironmental Reconstruction in the Southern Levant: synthesis, challenges, recent developments and perspectives. Philosophical Transactions of the Royal Society 368:5225-5248.

Rambeau, C., B. Finlayson, S. Smith, S. Black, R.. Inglis and S. Robinson 2011 16 Palaeoenvironmental reconstruction at Beidha, southern Jordan (c. 18,000– 8,500 BP): Implications for human occupation during the Natufian and Pre-Pottery Neolithic. In Water, Life and Civilisation: Climate, Environment and Society in the Jordan Valley, edited by S. Mithen, pp. 245.

Rapoport, A. 1969 House Form and Culture. Prentice-Hall, New Jersey.

1982 The meaning of the built environment: A nonverbal communication approach. University of Arizona Press.

261

Rasmussen, M., L.S. Cummings, M.T.P. Gilbert, V. Bryant, C. Smith, D.L. Jenkins and E. Willerslev 2009 Response to Comment by Goldberg et al. on “DNA from Pre-Clovis Human Coprolites in Oregon, North America”. Science 325(5937):148.

Reimer, P. J., M. G. L. Baillie, E. Bard, A. Bayliss, J. W. Beck, C. J. H. Bertrand, P. G. Blackwell, C. E. Buck, G. S. Burr, K. B. Cutler, P. E. Damon, R. L. Edwards, R. G. Fairbanks, M. Friedrich, T. P. Guilderson, A. G. Hogg, K. A. Hughen, B. Kromer, G. McCormac, S. Manning, C. B. Ramsey, R. W. Reimer, S. Remmele, J. R. Southon, M. Stuiver, S. Talamo, F. W. Taylor, J. van der Plicht and C. E. Weyhenmeyer 2004 IntCal04 terrestrial radiocarbon age calibration 0-26 kyr BP. Radiocarbon 46(3):1029-1058.

Reimer, P.J., Mike GL Baillie, Edouard Bard, Alex Bayliss, J Warren Beck, Paul G Blackwell, Christopher Bronk Ramsey, Caitlin E Buck, George S Burr and R Lawrence Edwards 2009 IntCal09 and Marine09 radiocarbon age calibration curves, 0-50,000 years cal BP.

Renfrew, C. 1973 Monuments, Mobilization and Social Organization in Neolithic Wessex. In The Explanation of Culture Change: Models in Prehistory, edited by C. Renfrew, pp. 539- 558. Gerald Duckworth and Company Ltd, London.

Reynolds, P.J. 1979 Farm: The Buster Experiment. , London.

Robinson, S.A., S. Black, B.W Sellwood and P.J. Valdes 2006 A review of palaeoclimates and palaeoenvironments in the Levant and Eastern Mediterranean from 25,000 to 5000 years BP: setting the environmental background for the evolution of human civilisation. Quaternary Science Reviews 25(13):1517-1541.

Rollefson, G.O. 1997 Changes in Architecture and Social Organization at 'Ain Ghazal. In The Prehistory of Jordan II: Perspectives from 1997, edited by H. G. Gebel, Z. Kafafi and G. O. Rollefson, pp. 287-308. ex oriente, Berlin.

Rollefson, G.O. and Z. Kafafi 1994 The 1993 Season at 'Ain Ghazal: Preliminary Report. Annual of the Department of Antiquities of Jordan 38(11-32).

Rollefson, G.O. and A.H. Simmons 1987 The Life and Death of 'Ain Ghazal. Archaeology 40(6):38-45.

Rosen, A.M. 1986a Alluvial Stratigraphy of the Shephela and its Paleoclimatic Implications. Geological Society of Israel.

262

1986b Environmental change and settlement at Tel Lachish, Israel. Bulletin of the American Schools of Oriental Research:55-60.

1991 Early Bronze Age Tel Erani: an environmental perspective. Tel Aviv 1991(2):192- 04.

1992 Preliminary identification of silica skeletons from Near Eastern archaeological sites: an anatomical approach. In Phytolith Systematics, edited by G. Rapp Jr. and S. C. Mulholland, pp. 129-147. Plenum Press, New York.

1997 Environmental change and human adaptational failure at the end of the Early Bronze Age in the Southern Levant. In Third millennium BC climate change and Old World collapse, edited by H. N. Dalfes, G. J. Kukla and H. Weiss, pp. 25-38. Springer.

1999 Phytolith analysis in Near Eastern Archaeology.

Rosen, A.M. and S. Weiner 1994 Identifying ancient irrigation: a new method using opaline phytoliths from emmer wheat. Journal of Archaeological Science 21(1):125-132.

Rosen, A.R. 1993 Phytolith Evidence for Early Cereal Exploitation in the Levant. In Current research in Phytolith Analysis: Applications in Archaeology and Paleoecology, edited by D. M. Pearsall and D. R. Piperno, pp. 161-171. MASCA Research Papers in Science and Archaeology. vol. 10. University of Pennsilvania Press, Philadelphia.

Rosen, Arlene M. 2007 Civilizing Climate: Social Responses to Climate Change in the Ancient Near East. AltaMira Press, Lanham.

Rosen, S.A. 1997 Lithics after the Stone Age: A Handbook of Stone Tools from the Levant. Altamira Press, Walnut Creek CA.

Rosenberg, D. 2010 Early Maceheads in the Southern Levant: a "Chalcolithic" Hallmark in Neolithic Context. Journal of Field Archaeology 35(2):204-216.

Rosenberg, M. and R.W. Redding 1998 Early pig husbandry in southwestern Asia and its implications for modeling the origins of food production. MASCA research papers in science and archaeology 15:55- 64.

Roshwalb, A. 1981 Protohistory in the Wadi Ghazzeh: A typological and technological study based on the Macdonald excavations., Institute of Archaeology, University College London.

263

Rossignol-Strick, Martine 1993 Late Quaternary climate in the eastern Mediterranean region. Paléorient 19(1):135-152.

1999 The Holocene climatic optimum and pollen records of sapropel 1 in the eastern Mediterranean, 9000–6000 BP. Quaternary Science Reviews 18(4–5):515-530.

Roux, V. and M-A. Courty 2005 Identifying Social Entities at a Macro-Regional Level: Chalcolithic Ceramics of South Levant as a Case Study. In Pottery Manufacturing Processes: Reconstitution and Interpretation, edited by A. Livingstone Smith, D. Bosquet and R. Martineau, pp. 201- 214. British Archaeological Reports, Oxford.

Roux, Valentine, M.-A. Courty and J.L. Lovell 2011 A Techno-petrographic Approach for Defining Cultural phases and Communities: Explaining the Variability of Abu Hamid (Jordan Valley) Early 5th Millennium cal. BC Ceramic Assemblage. In Culture, Chronology, and the Chalcolithic, edited by J. L. Lovell and Y. Rowan. CBRL Monographs. Oxbow, Oxford.

Rowan, Y. M. and J. Golden 2009 The Chalcolithic Period of the Southern Levant: A Synthetic Review. Journal of World Prehistory 22(1):1-92.

Rowan, Y. M. and D. Ilan 2013 The Subterranean Landscape of the Southern Levant during the Chalcolithic Period. In Sacred Darkness: A Global Perspective on the Ritual Use of Cave, edited by H. Moyes, pp. 87-107. University Press of Colorado.

Rowan, Y. M. and J.L. Lovell 2011 Introduction: Culture, Chronology and the Chalcolithic. In Culture, Chronology and the Chalcolithic: Theory and Transition, edited by J. L. Lovell and Y. M. Rowan, pp. 1-11. Levant Supplementary Series. Oxbow Books, Oxford.

Rowan, Y.M. and M.M. Kersel 2014 New perspectives on the Chalcolithic Period in the Galilee: investigations at the site of Marj Rabba. In Material Culture Matters: Essays on the Archaeology of the Southern Levant in Honor of Seymour Gitin, edited by J. R. Spencer, R. A. Mullins and A. J. Brody, pp. 221-238. Eisenbrauns, Winona Lake, Indiana.

Rowan, Y.M., T.E. Levy, D. Alon and Y. Goren 2006 Gilat's Ground Stone Assemblage: Stone Fenestrated Stands, Bowls, Palettes, and Related Artifacts. In Archaeology, anthropology and cult: The sanctuary at Gilat, Israel, edited by T. E. Levy, pp. 575-684. Equinox.

Sahlins, M. D. 1958 Social stratification in Polynesia 29. University of Washington Press Seattle.

264

1960 Evolution: Specific and General. In Evolution and Culture, edited by M. Sahlins and e. al. University of Michigan Press, Ann Arbor.

Salāmah, I. and H. Bannayan 1993 Water resources of Jordan: Present status and future potentials. Friedrich Ebert Stiftung.

Sassaman, K. E. and W. Rudolphi 2001 Communities of practice in the early pottery traditions of the American southeast. Journal of Anthropological Research 57(4):407-425.

Saxe, A. 1970 Social Dimensions of Mortuary Practices, Department of Anthropology, University of Michigan, Michigan.

Schick, T. 1998 The Cave of the Warrior: A Fourth Millennium Burial in the Judean Desert. IAA Reports 5. Israel Antiquities Authority, Jerusalem.

Schwab, M.J., F. Neumann, T. Litt, J. Negendank and M. Stein 2004 Holocene palaeoecology of the Golan Heights (Near East): investigation of lacustrine sediments from Birkat Ram crater lake. Quaternary Science Reviews 23(16):1723-1731.

Seaton, P. 2008 Chalcolithic Cult and Risk Management at Teleilat Ghassul. BAR: Monographs of the Sydney University Teleilat Ghassul Project 2. Archaeopress, Oxford.

Service, E.R. 1962 Primitive Social Organisation: an evolutionary perspective. Random House, New York.

Severinghaus, J.P., and E.J. Brook 1999 Abrupt Climate Change at the End of the Last Glacial Period Inferred from Trapped Air in Polar Ice. Science 286(5441):930-934.

Severinghaus, J.P., T.Sowers, E.J. Brook, R.B. Alley, and M.L. Bender 1998 Timing of abrupt climate change at the end of the Younger Dryas interval from thermally fractionated gases in polar ice. Nature 391:141-146.

Shahack-Gross, R. 2011 Herbivorous livestock dung: formation, taphonomy, methods for identification, and archaeological significance. Journal of Archaeological Science 38(2):205-218.

265

Shahack-Gross, R., R.A. Albert, A. Gilboa, O. Nagar-Hilman, I. Sharon and S. Weiner 2005 Geoarchaeology in an urban context: the uses of space in a Phoenician monumental building at Tel Dor (Israel). Journal of Archaeological Science 32(9):1417- 1431.

Shahack-Gross, R. and I. Finkelstein 2008 Subsistence practices in an arid environment: a geoarchaeological investigation in an Iron Age site, the Negev Highlands, Israel. Journal of Archaeological Science 35(4):965-982.

Shahack-Gross, R., F. Marshall and S. Weiner 2003 Geo-ethnoarchaeology of pastoral sites: the identification of livestock enclosures in abandoned Maasai settlements. Journal of Archaeological Science 30(4):439-459.

Shalev, S. 1991 Two different copper industries in the Chalcolithic culture of Israel. Decouverte du metal:413-424.

1995 Metals in Ancient Israel: Archaeological Interperetation of Chemical Analysis. Israel journal of chemistry 35(2):109-116.

1999 Recasting the Nahal Mishmar hoard: experimental archaeology and metallurgy. In The Beginnings of Metallurgy, edited by A. Hauptmann, E. Pernicka, T. Rehren and U. Yalcin. Dt. Bergbau Museum, Bochum.

Shalev, S. and P. J. Northover 1987 Chalcolithic metal and metalworking from Shiqmim. In Shiqmim I: Studies Concerning Chalcolithic Societies in the Northern Negev Desert, Israel (1982-1984), edited by T. E. Levy, pp. 357-371. BAR International Series 356. British Archaeological Reports, Oxford.

Shalev, S. and P.J. Northover 1987 Chalcolithic metal and metalwroking from Shiqmim. In Siqmim I: Studies Concerning Chalcolithic Societies in the Northern Negev Desert, Israel (1982-1984), edited by T. E. Levy. BAR International Series. BAR, Oxford.

Shennan, S. 2011 Property and wealth inequality as cultural niche construction. Philosophical Transactions of the Royal Society B: Biological Sciences 366(1566):918-926.

Shennan, S.J. 1993 Settlement and social change in Central Europe, 3500–1500 BC. Journal of world prehistory 7(2):121-161.

Shillito, L-M., I.D. Bull, W. Matthews, M.J. Almond, J.M. Williams and R.P. Evershed 2011 Biomolecular and micromorphological analysis of suspected faecal deposits at Neolithic Çatalhöyük, Turkey. Journal of Archaeological Science 38(8):1869-1877.

266

Shillito, M., W. Matthews, M. Almond and I. Bull 2011 The microstratigraphy of middens: capturing daily routine in rubbish at Neolithic Catalhoyuk, Turkey. Antiquity 85:1024-1038.

Shugar, A.N. 2001 Archaeometallurgical investigation of the Chalcolithic site of Abu Matar, Israel: A reassessment of technology and its implications for the Ghassulian Culture. Unpublished Ph. D. dissertation, Institute of Archaeology, University College London.

Shugar, A.N. and C. Gohm 2011 Developmental Trends in Chalcolithic Copper Metallurgy: A radiometric perspective changed the world. In Culture, Chronology and the Chalcolithic: Theory and transition, edited by J. Lovell and Y. M. Rowan. Levant Supplementary Series Oxbow Books, Oxford.

Sillar, B. 2000 Dung by Preference: The choice of fuel as an example of how andean potter production is embedded within wider technical, social and economic practices. Archaeometry 42(1):43-60.

Silver, I.A. 1969 The ageing of domestic animals. In Science in Archaeology, edited by D. R. Brothwell and E. S. Higgs, pp. 283-302. 2 ed. Praeger, New York.

Smart, T.L. and E.S. Hoffman 1988 Environmental interpretation of archaeological charcoal. Current paleoethnobotany:167-205.

Sneh, A., Y. Bartov, T. Weissbrod and M. Rosensaft 1998 Geological Map of Israel, 1:200,000, edited by I. G. Survey, pp. 4 sheets.

Spielmann, K.A. 2002 Feasting, Craft Specialization, and the Ritual Mode of Production in Small‐Scale Societies. American Anthropologist 104(1):195-207.

Stein, G. J. 1998 Heterogeneity, power, and political economy: Some current research issues in the archaeology of Old-World complex societies (With bibliography). Journal of Archaeological Research 6(1):1-44.

Stein, M. 2001 The sedimentary and geochemical record of Neogene-Quaternary water bodies in the Dead Sea Basin - inferences for the regional paleoclimatic history. Journal of Paleolimnology 26(3):271-282.

267

Stein, M., A. Torfstein, I. Gavrieli and Y. Yechieli 2010 Abrupt aridities and salt deposition in the post-glacial Dead Sea and their North Atlantic connection. Quaternary Science Reviews 29:567-575.

Stekelis, M. 1935 Les monuments Mégalithiques de Palestine. Archives de l'Institut de Paléontologie Humaine, Mémoire 15. Masson et Cie, Éditeurs, Paris.

Stuiver, M., P.J. Reimer, E. Bard, J.W. Beck, G.S. Burr, K.A. Hughen, B. Kromer, G. McCormac, J. Van der Plicht and M. Spurk 1998 INTCAL98 radiocarbon age calibration, 24,000-0 cal BP. Radiocarbon 40(3):1041-1083.

Thomas, E. R., E. W. Wolff, R. Mulvaney, J. P. Steffensen, S. J. Johnsen, C. Arrowsmith, J. W. C. White, B. Vaughn and T. Popp 2007 The 8.2 ka event from Greenland ice cores. Quaternary Science Reviews 26(1- 2):70-81.

Thomas, M., W. Rijm and A. F. B. van der Poel 2001 Functionality of raw materials and feed composition. In Feed manufacturing in the Mediterranean region. Improving safety: From feed to food, edited by J. Brufau, pp. 87-102. Cahiers Options Méditerranéennes. vol. 54. Zaragoza : CIHEAM.

Thomsen, C.J. 1848 Guide to Northern Archaeology English ed. Translated by E. o. Ellesmere, London.

Torfstein, A, S.L Goldstein, M. Stein, Y. Enzel and R. K. Sharaiha 2013 Impacts of abrupt climate changes in the Levant from Last Glacial Dead Sea levels. Quaternary Science Reviews 69:1-7.

Twiss, K. C. 2008 Transformations in an early agricultural society: Feasting in the southern Levantine Pre-Pottery Neolithic. Journal of Anthropological Archaeology 27(4):418-442.

Twiss, K.C., A. Bogaard, M. Charles, J. Henecke, N. Russell, L. Martin and G. Jones 2009 Plants and Animals Together. Current Anthropology 50(6):885-895.

Tzori, N. 1958 Neolithic and Chalcolithic Sites in the Valley of Beth-Shan. Palestine Exploration Quarterly 90:44-51.

Ussishkin, D. 1971 The" Ghassulian" temple in Ein Gedi and the origin of the hoard from Nahal Mishmar. The Biblical Archaeologist:23-39.

1980 The Ghassulian Shrine at En-Gedi. Tel Aviv 7(1-2):1-44.

268

Ussishkin, David 1980 The Ghassulian shrine at En-Gedi. Tel Aviv 7:1-44. van Beek, G.W. 1972 Notes and News: Tel Gamma. Israel Exploration Journal 22.

Van den Brink, E.C.M. 1998 An index to Chalcolithic mortuary caves in Israel. Israel Exploration Journal:165-173.

Varien, M.D. and R.H. Wilshusen (editors) 2002 Seeking the Center Place: Archaeology and Ancient Communities in the Mesa Verde Region. University of Utah Press, Salt Lake City.

Verhoeven, M. 1999 An Archaeological Ethnography of a Neolithic Community: Space, Place and Social Relations in the Burnt Village at Tell Sabi Abyah, Syria. Nederlands Historisch- Archaeologisch Institut, Istanbul.

Verhoeven, M. and P. Kranendonk 1996 The excavations: stratigraphy and architecture. In Tell Sabi Abyad. The late neolithic settlement. Report on the excavations of the university of Amsterdam (1988) and the national museum of antiquities Leiden (1991-1993) in Syria, Istanbul, edited by P. M. M. G. Akkermans, pp. 25-118.

Walker, R. 2011 SI metric: density of bulk materials (webpage). accessed 15/01/2013, http://www.simetric.co.uk/si_materials.htm.

Watson, P.J. 1979 Archaeological Ethnography in Western Iran. Viking Fund Pulbications in Anthropology 57. University of Arizona Press, Tuscon.

Weaver, A.J., O.A. Saenko, P.U. Clark, and J.X. Mitrovica 2003 Meltwater Pulse 1A from Antarctica as a Trigger of the Bolling-Allerod Warm Interval. Science 299:1709-1713.

Weippert, H. 1998 Sanctuary as meeting place-An example of a chalcolithic sacred shrine at Gilat. Zeitschrift des Deutschen Palastina-Vereins 114(2):106-136.

Wernke, S.A. 2007 Negotiating Community and Landscape in the Peruvian Andes: A Transconquest View. American Anthropologist:130-152.

269

Whitcher, S.E., C. Grigson and T.E. Levy 1998 Recent faunal analysis at Shiqmim, Israel. In Archaeozoology of the Near East III, Proceedings of the third international symposium on the archaeozoology of Southwestern Asia and adjacent areas. , edited by H. Buitenhuis, L. Bartosiewicz and A. Choyke, pp. 103-116. The Institute of Archaeology, Gronigen.

Wilkinson, T.J. 2000 Regional approaches to Mesopotamian archaeology: The contribution of archaeological surveys. Journal of Archaeological Research 8(3):219-267.

Willcox, G. and S. Fornite 1999 Impressions of wild cereal chaff in pisé from 10th millennium uncal BP at Jerf el Ahmar and Mureybit, northern Syria. Vegetation History and Archaeobotany 8:21-24.

Willcox, G. and D. Stordeur 2012 Large-scale cereal processing before domestication during the tenth millenium cal BC in northern Syria. Antiquity 86:99-114.

Wills, W.H. and R.D. Leonard (editors) 1994 The Ancient Southwestern Community: Models and Methods for the Study of Prehistoric Social Organization. University of New Mexico Press, Albuquerque.

Winter-Livneh, R., T. Svoray and I. Gilead 2012 Secondary burial cemeteries, visibility and land tenure: A view from the southern Levant Chalcolithic period. Journal of Anthropological Archaeology 31(4):423-438.

Worsaae, J.J.A. 1849 The primeval antiquities of Denmark. Translated by W.Thoms, London.

Wreschner, E.E. 1977 New Yam--a Submerged Late-Neolithic Settlement Near Mount Carmel. Eretz- Israel 13:260-271.

Wright, H 1984 Prestate political formations. In On the Evolution of Complex Societies, edited by T. Earle, pp. 47-77. Undena, Malibu.

Yasuda, Y., H. Kitagawa and T. Nakagawa 2000 The earliest record of major anthropogenic deforestation in the Ghab Valley, northwest Syria: a palynological study. Quaternary International 73:127-136.

Yoffee, N. 1993 Too many chiefs? (or , Safe texts for the '90s). In Archaeological Theory: Who Sets the Agenda?, edited by N. Yoffee and A. Sherratt, pp. 60-78. Cambridge University Press, Cambridge.

270

Zeder, M.A. 1996 The role of pigs in Near Eastern subsistence: A view from the Southern Levant. Retrieving the Past: Essays on Archaeological Research and Methodology in Honor of Gus W. Van Beek, Eisenbrauns, Winona Lake, IN:297-312.

1998 Pigs and emergent complexity in the ancient Near East. MASCA research papers in science and archaeology 15:109-122.

2003 Food provisioning in urban societies: A view from northern Mesopotamia. In The Social Construction of Ancient Cities, edited by M. L. Smith, pp. 156-183. Smithsonian Institution Press, Washington, DC.

Zeder, M.A., E. Emshwiller, B.D. Smith and D.G. Bradley 2006 Documenting domestication: the intersection of genetics and archaeology. TRENDS in Genetics 22(3):139-155.

Zohary, M. 1962 Plant Life of Palestine, Israel and Jordan. Chronica Botanica, New Series of Plant Science Books 33. Ronald Press, New York.

271

Appendices

Appendix I: Micromorphology Sample Analysis Forms

272

273

TS06-1 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Sand Moderate – good Mod Sandy clay 90 Calcite Unsorted 50+

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

10 enaulic Unsorted sharp

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

<5% Chambers

Orientation/Distribution Patterns:

Unoriented Random distribution of inclusions/features

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

Charred wood/seed

c. 7 pieces of bone frags

Mircroartefacts: Rock and Mineral Inclusions: Others: 28+ flint sherds (1500–2000 µm) Pottery frags – small, limited Basalt/gabbro < 2% of slide Burnt mudbrick -55% of slide Comments and Summary: Taken from Room 70, baulk between 2005 and 2006 excavation levels. Visible floor leves in eastern section but sampled from western section due to density of visible artefacts in east. Associated sampes: TS06-2, TS06-3, TS06-4, TS06-5 Upper most sample.

No floor level visible, post-occupational fill (sand + brick/mud render frags) Flint – micro flakes, not in situ, likely part of fill rather than knapping area.

274

TS06-02 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Poor - Moderate Poor Sandy clay 20 Sand Calcite poor

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

80 enaulic Poor sharp

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

Cracks 20-30% Plant voids Plant voids in mudbrick fragments Channels

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

1 piece of charcoal Plant voids 3 bone fragments

Mircroartefacts: Rock and Mineral Inclusions: Others: Building frags: Basalt/gabbro <5% >10 flint fragments Upper = mud render frag? Anhydrite/gypsum in voids/channels Lower = mudbrick frag? Comments and Summary: Taken from Room 70, baulk between 2005 and 2006 excavation levels. Visible floor leves in eastern section but sampled from western section due to density of visible artefacts in east. Associated sampes: TS06-1, TS06-3, TS06-4, TS06-5 From fill above first visible floor level, right and slightly below TS06-1

No floor level visible Flint present in very low density Post-depositional fill – predominantly building fragments of various sizes, no contiguous floor/brick

275

TS03-3 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting: 1 – 1 – Moderate 1- mod 1 - Sandy clay 1 – 40% 1 – unsorted loamy calcite 2 – Very Good 2- well 2 - clay 2 – 5% 2 – very well sand Fine Fraction – microscopic Texture: C/F related Distribution Sorting: Boundary: Notes: 1- 1 – 60 1 – enaulic 1 = fill level Unsorted Sharp 2 - 95 2 - porphyric 2 = mud-render 2- v.well Void Pattern % Total Voids: Dominant Type: Other Voids: Comments: 2 – 1 – 1% Horizontally oriented plans through planes/cracks, 2 – 25% mud-render surface plant voids

Orientation/Distribution Patterns:

1 – unoriented, random 2 – laminate, well oriented, random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

1- charcoal bits , <50 µm Charcoal 2 – plant voids 20%, 60– 1500 µm

Mircroartefacts: Rock and Mineral Inclusions: Others: 1 – 1 large inclusion = clay oven 1 – 1 flint flake frag? +mudbrick frags 2 – grog frags – temper?

Comments and Summary: Taken from Room 70, baulk between 2005 and 2006 excavation levels. Visible floor leves in eastern section but sampled from western section due to density of visible artefacts in east. Associated sampes: TS06-1 TS06-2, TS06-4, TS06-5 First floor level, mid-section of baulk, to the right of TS06-2

Fill: very similar to TS06-1 and TS06-2 Floor: clay with plant and pottery temper, evidence of 1-2 re-surfacing events including area of undulation that is likely a woven reed mat impression.

276

TS06-4 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting: Sandy Poor Sandy clay, Calcite Poor 30% silt poor silty clay loam Fine Fraction – microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Gradual, Clay largely present Silt, clay enaulic Moderate some sharp in floor/brick frags

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

Voids, 5% Verticle cracks chambers

Orientation/Distribution Patterns:

Horizontal floor level in upper portion Fill material: random orientation, random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

charcoal

Mircroartefacts: Rock and Mineral Inclusions: Others:

Pottery fragments Disturbed mud-rendered surface Flint frags

Comments and Summary: Taken from Room 70, baulk between 2005 and 2006 excavation levels. Visible floor leves in eastern section but sampled from western section due to density of visible artefacts in east. Associated sampes: TS06-1, TS06-2, TS06-3, TS06-5 Upper baulk – to the right of TS06-2, possible floor levels visible

Lower portion dominated by large mudbrick fragments with plant voids visible. Some areas of horizontally oriented fill 4 sections from top: fill, floor, fill (loose packed charred material as well), mudbrick or potential disturbed floor.

277

TS06-5 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Poor Sandy Poor Sandy clay 70 Calcite Mod clay

Fine Fraction – microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Clay and silt Enaulic Mod Difussed

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

< 5% chambers

Orientation/Distribution Patterns:

Mostly unoriented, random distribution Some areas of potential plant voids horizontally oriented, no phytoliths

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

Plant voids, horizontal Few bone frags

Mircroartefacts: Rock and Mineral Inclusions: Others: Basalt/gabbro Compact mud-rendered floor Flint frags – high density in centre Anhydrite/gypsum (post- frag. depositional) Building material frags Comments and Summary:

Taken from Room 70, baulk between 2005 and 2006 excavation levels. Visible floor leves in eastern section but sampled from western section due to density of visible artefacts in east. Associated sampes: TS07-1, TS06-2, TS06-3, TS06-4 From directly beneath TS06-1

Mud-rendered floor frag in loose packed matrix, not contiguous. Dense, angular fragments as well, likely building material frags. Fill material less dense towards lower portion of slide.

278

TS06-7 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Calcite, some Very Good Well Silty clay <5% Sand Well sorted limestone

Fine Fraction – microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Silty clay Porphyric Well Diffused

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

10% Chambers Channels

Orientation/Distribution Patterns:

Unoriented, random

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

Scattered charcoal

Mircroartefacts: Rock and Mineral Inclusions: Others:

Small pottery frag Basalt/gabbro Few flint frags

Comments and Summary:

From test pit to the west of main excavation area (UU16B). No visible floor levels, compact/dry mud layer atop loose, darker sediment.

General fill, anthropomorphic material mixed into natural environment.

279

TS06-8 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Moderate Sandy silt Sand/ Calcite Poor 70 Poor loam silt

Fine Fraction – microscopic Texture: C/F related Distribution Sorting: Boundary: Notes: clay Enaulic poor gradual, Some organic staining

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

Cracks, Channels, Long, horizontal channel at base of 10% vertical chambers slide likely from plant root

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions: Charred wood, charcoal from Very few spherulites in amorphous Phytoliths (horizontal reed stems) shrub organic matter shell Mircroartefacts: Rock and Mineral Inclusions: Others: Basalt/gabbro < 2% of slide Flint frags, generally horizontally Gypsum 25% Baked clay frags oriented Microfossils Comments and Summary:

From southern wall of Byzantine burial pit in eastern BC I courtyard. Taken from lower level of stepped excavation area – southern wall. Visible ash lines, possible floor level?? Associated samples: TS06-9, TS07-6

Multiple levels of use: two rough laid floor layers, 2 loose charcoal rake-out, some fill material Evidence of post-depositional plant/root disturbance.

280

TS06-9 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Sandy clay Poor Poor 20% sand Calcite Poor loam

Fine Fraction – microscopic Texture: C/F related Distribution Sorting: Boundary: Notes: Areas of dense Clay loam Enaulic Unsorted sharp material embedded in loose fill Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

Horizontal 10% chambers cracks

Orientation/Distribution Patterns:

Loose horizontal orientation Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions: Charred reed stems Melted corprolitic material in ashy Ash Dense horizontal phytoliths area bone

Mircroartefacts: Rock and Mineral Inclusions: Others:

Some flint Gypsum/andhydrite Building fragments

Comments and Summary:

From southern wall of Byzantine burial pit in eastern BC I courtyard. Lower step level – southern wall. Associated samples: TS06-8, TS07-6

Lower portion dominated by burnt material, similar to rake-out levels of TS06-8 Horizontal reed stem – charred and phytoliths Calcined bone fragment visible on the macro-scale present Dense clay building fragments embedded in fill

281

TS07-10 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Moderate Mod Sand 90 Sand Calcite Mod

Fine Fraction – microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Silt Monic Unsorted Sharp

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

15% Chambers Cracks

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

Plant voids in floor

Mircroartefacts: Rock and Mineral Inclusions: Others:

Gypsum/anhydrite, post-depositional

Comments and Summary: From testpit in presumed Byzantine level in UU16B – northeast corner of excavation area above Chalcolithic levels. From west corner of south wall, visible floor levels interspersed with reddish sediment. Associated sample: TS06-11.

Slide almost entirely anthropogenic material. Large and small building fragments with some sand grains as fill between features. Byzantine floor level, very dense fine calcareous material, few plant voids but many plane voids/cracks. Visibly different from Chalcolithic mud-render floors.

282

TS06-11 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Good Well Sandy clay 10% Sand Calcite Well

Fine Fraction – microscopic Texture: C/F related Distribution Sorting: Boundary: Notes: Anthropomorphic clay monic Unsorted Sharp clay with sharp boundary with sand Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

10$ Plane/cracks Chambers

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

Mircroartefacts: Rock and Mineral Inclusions: Others:

Few basalt/gabbro Few flint frags microfossils

Comments and Summary:

From testpit in presumed Byzantine level in UU16B – northeast corner of excavation area above Chalcolithic levels. From of western wall, visible floor levels interspersed with reddish sediment. Associated sample: TS06-10.

Possible floor fragment and mudbrick fragment within sandy matrix Byzantine – taken from test pit, visibly different from Chalcolithic building frags Fewer plant voids

283

TS06-13 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Moderate Mod Clay loam 5% Sand Calcite Mod

Fine Fraction – microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Clay/loam Prophyric Mod Diffused

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

<5% Voids

Orientation/Distribution Patterns:

Moderate horizontal orientation Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

Abundant Phytoliths – reed stem Spherulites – clustered and scattered Scattered charcoal frags fragments

Mircroartefacts: Rock and Mineral Inclusions: Others: some bone and flint, limited Building frags

Comments and Summary: From baulk left in Room 230. Visible stratigraphy through eastern section, western section bisects pit. From upper eastern portion, 3 visible ‘whitish’ levels, multiple visible reddish/white levels – very long and thin. Associated sample: TS06-14

Dense fine and organic material Some horizontal orientation, but generally unoriented material Spherulites indicated dung pellets and dispersed fecal material

284

TS06-14 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Poor mod Silty clay 10 Sand Calcite poor

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Irregular/ Multiple boundaries Silty clay Enaulic Poor sharp with varied materials

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

30% Cracks Channels Some trampling

Orientation/Distribution Patterns:

Strong bowlike orientation Parallel linear distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

Charred flecks Phytoliths – reed stems Spherulites – clustered and scattered Charred seeds

Mircroartefacts: Rock and Mineral Inclusions: Others:

Limited flint Basalt/gabbro – limited

Comments and Summary: From baulk left in Room 230. Visible stratigraphy through eastern section, western section bisects pit. From lower eastern portion, 3 visible ‘whitish’ levels, multiple visible reddish/white levels – very long and thin. Associated sample: TS06-13

Multiple elements including some building fragments, but no indication of a floor surface – mudbrick and ceramic dominate Dense clusters of spherulites, often in association with long reed stem phytoliths, indicate in situ dung deposits – key indicator of penning. Voids indicating trampling support this. Other indications of organic material are present in organic staining in some areas of slide Some diffused ash also present.

285

TS06-15 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Good Mod Sand 90 Sand Calcite Mod

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Loam Enaulic Unsorted Gradual

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

20% Chambers

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions: Charcoal Phytoliths - moderate Shell

bone Mircroartefacts: Rock and Mineral Inclusions: Others:

Flint frags Basalt/gabbro Building frags Pottery

Comments and Summary: From baulk left in Room 230. Visible stratigraphy through eastern section, western section bisects pit. From upper western portion, above pit. Larger visible layers of compressed mud and potential ashy areas. Associated samples: TS06-16, TS06-17

Mostly building frags in sandy fill Moderate phytolith density General fill with anthropogenic features

286

TS06-16 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Good Mod Loamy sand 15% Sand Calcite moderate

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Loam enaulic mod Sharp

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

5% Chambers

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions: Charcoal- woody and frags High density of Phytoliths, small fragments

Mircroartefacts: Rock and Mineral Inclusions: Others: Flint Basalt/gabbro Building frags

Comments and Summary:

From baulk left in Room 230. Visible stratigraphy through eastern section, western section bisects pit. From western portion, straddling eastern floor/below floor level of pit. Larger visible layers of compressed mud and potential ashy areas. Associated samples: TS06-15, TS06-17

Loosely packed fill with abundant phytoliths and charcoal - unoriented embedded building fragments General fill material - possibly from hearth/oven? No spherulites.

287

TS06-17 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Sandy silt- Moderate Mod 30% Sand Calcite Moderate loam

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes: Upper portion sandy, Silt loam enaulic porphyric well sharp lower exclusively silt/loam Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

10% Channels

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

Charcoal scatters Dense Phytoliths ash

Mircroartefacts: Rock and Mineral Inclusions: Others: Marl Pottery Basalt/gabbro Small frags of unworked clay Flint frags Building frags Comments and Summary:

Similar to TS06-16, dense fine material with charcoal and phytoliths – densities increase towards bottom of slide General fill including anthropogenic material (charcoal + phytolith) – hearth disposal?

288

TS07-1 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting: good Mod Silty clay 5% Sand Calcite Mod

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Silty clay Enaulic Well Diffused

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

Chambers Cracks Voids have general vertical orientation 25%

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

Some charcoal frags Same phytoliths – towards bottom A few near bottom shell

Mircroartefacts: Rock and Mineral Inclusions: Others:

Basalt/gabbro marls microfossils

Comments and Summary:

Extensive bioturbation as indicated by microaggregates. Well sorted, homogenous material Phytoliths and spherulites increase towards bottom of the slide Sand only present in the upper portion, fine material below Disturbed penning area? Or re-deposited penning material?

289

TS07-2 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Moderate Mod Silty clay 5% Sand Calcite Well

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Sharp boundary with Silty clay Enaulic Well Gradual mudbrick frag

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

5% Chambers

Orientation/Distribution Patterns: Unoriented Random distribution Moderately horizontally oriented, laminate organic material with linear distribution in middle-lower portion Inclusions Plant Evidence: Dung Evidence: Biological Inclusions: Charcoal/ ash Phytoliths – laminate and Spherulites – clustered and scattered bone scattered Amorphous organics Mircroartefacts: Rock and Mineral Inclusions: Others:

Basalt/gabbro Mudbrick frag microfossils

Comments and Summary:

Upper portion – re-deposited or disturbed dung, not vertically compressed Mid-Upper – melted dung (amorphous organics), ash deposits, higher biofringence Mid-Lower – laminated, in situ penning material (similar to TS06-14) Lower – disturbed material (see upper portion), some ash Bottom – foundation layer? Or highly disturbed in situ penning

290

TS07-6 Macroscopic Course Fraction – microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Sandy silt Good Well 20% Sand Calcite Poor loam

Fine Fraction – microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Clay Porphyric Mod Diffused marl

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

20% Chambers Vesicles

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

Charcoal – woody and scattered Bone

Mircroartefacts: Rock and Mineral Inclusions: Others: Basalt/gabbro - significant Flint frags Feltspar Clay chunks Ceramic/oven piece Obsidian? V. small frag Comments and Summary:

Taken from southwestern wall of byzantine burial pit (second one) that spans BC I and BC II (through wall). Sample from BCI courtyard. Associated samples: TS06-8, TS06-9

Very little anthropogenic material, homogenous material More sand in the upper portion of slide

291

TS07-8 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting: Calcite, Sandy clay Moderate Poor 60% Sand igneous, Mod loam limestone Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Large anthropogenic Clay-loam Enaulic Mod Diffused frags in loose matrix

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

50% Vughs Chambers Many, rounded vughs in loose matrix

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

Phytoliths in loose fill, grass Charcoal – woody, scattered Limited scattered spherulites forms Charred seeds

Mircroartefacts: Rock and Mineral Inclusions: Others: Basalt/gabbro frequent Mudbrick? Oven fragments? Few flint frags Marls Rough floor? microfossils Comments and Summary: Taken from the southern end of the pit found in Room 230. No architecture associated with these levels (circular walls of R230 do not continue below upper level of pit) Associated samples:TS07-9, TS07-10, TS07-12, TS07-13

No floor level visible, post-occupational fill? A number of large, consolidated sections with sharp boundaries – likely mudbrick fragments, oven fragments? Rough floor? Some plant voids visible – unoriented. Predominantly fine fraction. Areas of dense fine material including clays and marls.

292

TS07-9 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting: Well Calcite Moderate Mod Sandy clay 40% Sand poor Igneous Poor Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Diffused and Loose and dense Clay, marl enaulic, porphyric, Unsorted sharp sediment - defussed

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

Plans in dense material in lower 30% Vughs Chambers, planes portion

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions: Charred wood Phytoliths – few Small clusters of spherulites, rare Bone Shell Mircroartefacts: Rock and Mineral Inclusions: Others:

Basalt/gabbro Potential anthropogenic material Other igneous

Comments and Summary: Taken from the southern end of the pit found in Room 230. No architecture associated with these levels (circular walls of R230 do not continue below upper level of pit) Associated samples:TS07-8, TS07-10, TS07-12, TS07-13

Lower portion of slide has large fragments of dense material with large vughs between. The sediment has sharp boundaries with vughs and are clay lined. The densly pack material has long generally vertical planes, and differs in density from the upper portion which contains horizontally oriented chambers. Potentially anthropogenic, but diffused boundary with upper, looser sediment of same material makes it difficult to confirm. Clay lining water laid.

Unoriented, random distribution of calcitic sand with significant igneous inclusions suggests non-anthropomorphic deposition. Different composition to fill levels found in anthropomorphic contexts.

293

TS07-10 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting: Calcite Very good Poor Sandy clay 70 Sand Limestone unsorted Igneous Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Clay Porphyric Unsorted N/A

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

30% Vughs Chambers General verticle orientation

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

Shell Limited charcoal

Mircroartefacts: Rock and Mineral Inclusions: Others: Basalt/gabbro Microfossils flint Feldspar marl Comments and Summary: Taken from the southern end of the pit found in Room 230. No architecture associated with these levels (circular walls of R230 do not continue below upper level of pit) Associated samples:TS07-8, TS07-10, TS07-12, TS07-13

Well consolidated, homogenous sediment with a high percent of sand composed of both sedimentary and igneous rock. Similar to TS07-9 material. Not like anthropogenic fill levels found elsewhere. Natural, aeolean deposition?

294

TS07-12 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Sand Calcite Moderate Poor Sandy clay 30 unsorted Igneous

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes: Monic (loose material) Some sharp Sharp Clay Porphyric (dense Unsorted boundaries, some Diffused material) diffused Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

20% Vughs Chambers General verticle orientation

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

Bone Phytoliths – in loose material only Charred seeds

Mircroartefacts: Rock and Mineral Inclusions: Others: Basalt/gabbro Limestone Dense – mudbrick??? marls Other igneous Comments and Summary: Taken from the southern end of the pit found in Room 230. No architecture associated with these levels (circular walls of R230 do not continue below upper level of pit) Associated samples:TS07-8, TS07-9, TS07-10, TS07-13

Similar to TS07-10, dense material does not include plant voids – higher sand content than other mudbrick. Some sharp boundaries with loose fill/phytolith material, but others are more gradual. Not like anthropogenic fill from other contexts. Disturbed aeolean desposits???

295

TS07-13 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Calcite Very Good poor Sandy clay 40% Sand unsorted igneous

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Clay/marl porphyric Unsorted None

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

15% Vughs Large, unoriented

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

Shell Charcoal

Mircroartefacts: Rock and Mineral Inclusions: Others: Basalt/gabbro Large pottery fragment Marl

10mm X 8mm Limestone feldspar Comments and Summary:

Taken from the southern end of the pit found in Room 230. No architecture associated with these levels (circular walls of R230 do not continue below upper level of pit) Associated samples:TS07-8, TS07-9, TS07-10, TS07-12

Other than pottery fragment, not anthropogenic material. Very similar to TS07-10. Homogenous matrix, no boundaries. Eolean? Not like anthropogenic fill from other contexts.

296

TS07-18 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Sandy clay Moderate Poor 5% Sand Calicite poor loam

Fine Fraction – microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Clay Enaulic Unsorted Sharp marl porphyric

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

Cracks generally lower portion, 15% Planes/cracks Vughs horizontal

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

Limited charcoal

Charred seeds

Mircroartefacts: Rock and Mineral Inclusions: Others:

Basalt/gabbro Stuf Large limestone frag – 9mm X 4mm

Comments and Summary:

Taken from “wall” of Room 662, below mudbrick spill. Associated samples: TS07-19, TS07-20B

Homeogenous, varying density. Dense clumps with sharp boundaries in loose matrix. Material is generally the same in each. Potential rough mudbrick fragments? No plant voids.

Vertical planes may indicate trampling or impact compression from mudbrick spill? Definitely not a wall.

297

TS07-19 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Sandy clay Calcite Moderate poor 5% Sand Poor loam Igneous

Fine Fraction – microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Clay Porphyric Unsorted Gradual marl

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

General horizontal orientation of 5-10% Planes Vughs planes

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions: Charred wood/seed Bone

Shell Mircroartefacts: Rock and Mineral Inclusions: Others:

Basalt/gabbro

Comments and Summary: Taken from “wall” of Room 662, below mudbrick spill. Associated samples: TS07-18, TS07-20B

Similar composition to TS07-18, dense and consolidated sediment in loose matrix of similar composition. Some sharp and some defused boundaries.

Horizontal orientation of voids (esp planes) indicate trampling or compression (from mudbrick spill?)

298

TS07-20B Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Sandy clay Calcite Moderate Mod 5% Sand Poor loam Igneous

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes: Loose material at top, Clay Enaulic Unsorted Sharp consolidated in lower Marl Porphyric portion Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

5% Planes Vugh Horizontal orientation of planes

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

Bone

shell

Mircroartefacts: Rock and Mineral Inclusions: Others:

Basalt/gabbro Mudbrick? microfossils

Comments and Summary:

Taken from “wall” of Room 662, below mudbrick spill. Associated samples: TS07-18, TS07-19

Similar to TS07-18 and 19. Loose and consolidated material of similar make-up. Some sharp boundaries between the two. Horizonal cracks may indicate trampling or compression (mud brick spill?).

299

TS07-20 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Poor Poor Sandy clay Sand Calcite unsorted

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Enaulic Clay Unsorted Sharp Porphyric

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

10% Chambers

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

Plant voids in mudrender

Mircroartefacts: Rock and Mineral Inclusions: Others:

Mud brick frags Basalt/gabbro < 2% of slide Mud-rendered wall layer

Comments and Summary: Taken from south wall of Room 70 interior, visible plaster layer

Outer surface, unconsolidated enaulic sediment, some mudbrick fragments, loose fill similar to other R70 fill material

Mudrender layer, dense homogeneous clay/mud, some cracks, plant voids not as horizontally oriented as floor levels Beneath Mudrender, porphyric sediment of similar composition to outer surface material, densly packed

300

TS07-22 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Moderate Poor Silty clay 5% Sand Calcite Poor

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Enaulic Clay Unsorted Sharp Porphyric

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

10% Chambers Planes Vertical orientation of planes

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions: Charcoal Ash

Amorphous organics Mircroartefacts: Rock and Mineral Inclusions: Others:

Basalt/gabbro

microfossils

Comments and Summary:

Taken from baulk in interior ‘floor’ of Room 662. Associated sample TS07-23

Consolidated fragments with sharp boundaries within loose matrix of similar material.

Not a floor level – potential fill? Building frags in fill level?

301

TS-23 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Mod Mod Sandy clay 5% Sand Calcite poor

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Clay Porphyric Unsorted Gradual

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

15% Planes Vughs Strong horizontal plane orientation

Orientation/Distribution Patterns:

Locally oriented – horizontal Linear distribution of some areas

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

Charred wood Phytoliths ash

Mircroartefacts: Rock and Mineral Inclusions: Others: Basalt/gabbro Limestone

Gypsum (post-depositional) Comments and Summary: Taken from baulk in interior ‘floor’ of Room 662. Associated sample TS07-22.

Upper portion demonstrates similar features and orientation/distribution as courtyard floor surface in Room 70. Plant voids have post-depositional gypsum.

Lower portion dense, clay with basalt, ash and phytoliths.

302

TS07-24 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Poor Poor Clay loam 10% Silt Calcite Poor

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Clay Enaulic Unsorted Diffused

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

<5% Vesicles Chambers General verticle orientation of vesicles

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions: Charred seeds common Spherulites in amorphous organic Phytoliths Charcoal – woody matter plus scattered Calcined bone Mircroartefacts: Rock and Mineral Inclusions: Others:

Basalt/gabbro chert Unworked clay fragments microfossils

Comments and Summary: Taken from western side of wall 580 – southern courtyard area of BC IV, oven feature. Associated sample TS07-25

Sediment in upper portion less dense.

Evidence of melted/heated organic matter including dung, charred wood and seeds, calcined bone all consistent with cooking areas. Dominance of geological material (no dense ash/charcoal) may indicate rack out or upper fill?

303

TS07-25 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Very poor Poor Silty clay <5% Calcite Unsorted

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes: Some sharp Silty clay Enaulic Unsorted Graduated boundaries with anthro. frags Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

5% Chambers

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions: Charred wood Spherulites with phytoliths, Charcoal scattered Phytoliths abundant amorphous organic material Ash Bone Mircroartefacts: Rock and Mineral Inclusions: Others:

Basalt/gabbro some Possible oven frags?

Comments and Summary:

Varying degrees of density throughout, mostly calcitic clay with amorphous organic matter (phytoliths and spherulites associated). Ash most dense in the upper portion.

Dark burnt areas

All consistent with cooking features.

304

TS07-30 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Poor Good Silty clay <5% Sand Calcite Unsorted

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Clay Enaulic Unsorted Sharp Marl

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

15% Vughs Plans

Orientation/Distribution Patterns:

Local parallel orientation Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions: Charcoal – large woody Abundant scattered Phytoliths – abundant Some spherulites Amorphous organic ash Mircroartefacts: Rock and Mineral Inclusions: Others:

Building frags with clay Basalt/gabbro < 2% of slide choating

Comments and Summary: Taken from lower portion of BC I courtyard baulk located next to the northeast corner of Room 70. Associated sample TS07-31

Upper portion dominated by fragments of building material (mud brick?) of various sizes. Clay coated with thin layer.

Sections of dense amorphous organic material with charcoal, phytoliths, ash and marl – partially ashed dung? Potential courtyard floor surface at the very lower boundary of the slide.

305

TS07-31 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Calcite Good Good Silty clay <5% sand Good limestone

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Clay Waterlaid clay Enaulic Well Diffused Marl Banded sand

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

5% Vughs

Orientation/Distribution Patterns:

Strong parallel orientation Linear/banded distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions: Charred wood Phytoliths Spherulites - clustered Shell bone Mircroartefacts: Rock and Mineral Inclusions: Others: Basalt/gabbro Flint frags Quartz

Microfossils Comments and Summary: Taken from lowest extreme of BC I courtyard baulk located next to the northeast corner of Room 70. Associated sample TS07-30

Anthropogenic material within water laid, sorted sediment Some corprolitic material – isolated, consolidated

Likely deposited prior to Phase 4 occupation of Building Complex I Some organic staining – bacterial?

306

TS07-33 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Very good Poorly Silty clay <1% Calcitic Unsorted

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Clay porphyric Unsorted marl

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

10% Vughs rounded

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

Mircroartefacts: Rock and Mineral Inclusions: Others:

Comments and Summary:

Mortor section between mud brick rows on SE wall of Room 70, clearly textured in situ

Consolidated clay/marl, appears to be oxidated. Baked mudbrick?

307

TS07-34 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Clay/silty Moderate Poor 5% Sand Calcite unsorted clay

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes: Clay Silt porphyric Unsorted Diffused Oxidized clay marl Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

15% Planes/cracks Vughs Unoriented planes

Orientation/Distribution Patterns:

Moderate bowlike orientation Linear, laminate distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

Mircroartefacts: Rock and Mineral Inclusions: Others:

v. few Unworked clay frags

Comments and Summary:

Taken from exposed wall of chalcolithic burial in Silo 339. Related samples TS07-36, TS07-37

Fine clay/marl, some oxidized clay (iron rich – reddish/orange)

Potentially a very poorly constructed mortar or mudbrick - unworked

308

TS07-36 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Calcite Good Mod Silty clay 5% Sand Poor Igneous

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Clay Porphoric Unsorted marl

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

Both horizontal and vertical 35% Planes/cracks Chambers orientation

Orientation/Distribution Patterns:

Unoriented Random distribution of inclusions/features

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

bone

Mircroartefacts: Rock and Mineral Inclusions: Others:

Basalt/gabbro Pottery frags Unworked clay frags microfossils

Comments and Summary:

Taken from exposed wall of chalcolithic burial in Silo 339. Related samples TS07-34, TS07-37

Well consolidated clay/marls – no plant voids. Rough mudbrick? Clay packing/lining?

309

TS07-37 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Moderate Mod Clay 1%

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Clay Fine monic Unsorted sharp

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

Large vughs in lower portion between 45% Vughs Planes/cracks sharp boundary fragments

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

Shell Phytoliths spherulites Charred seeds

Mircroartefacts: Rock and Mineral Inclusions: Others:

Basalt/gabbro Oxidised clay microfossils

Comments and Summary:

Taken from exposed wall of chalcolithic burial in Silo 339. Related samples TS07-34, TS07-36

Sharp edged consolidated fragments in lower portion same composition as consolidated section in upper portion. Fragmented unworked/rough made mudbrick or clay lining??

310

TS07-38 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Sandy silt Poor Mod 25% Sand Calcite unsorted loam

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes: Silt Enaulic Loam Unsorted Diffused Porphoric Marl Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

10$ Vughs

Orientation/Distribution Patterns:

Moderate parallel oriention Linear and clustered distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions: Charred wood, one especially Phytoliths clustered (pellets) Spherulites – clustered in pellets large frag: scattered Scattered. Calcined bone, shell, ash Mircroartefacts: Rock and Mineral Inclusions: Others: 28+ flint sherds (1500–2000 µm) Basalt/gabbro Pottery frags – small, limited Gypsum/anhydrite Organic staining Burnt mudbrick -55% of slide Comments and Summary: Taken from Room 612, western wall of room near burn level. Associated samples TS07-39, TS07-40, TS07- 41, TS07-42 Lower portion shows parallel orientation or loose/unconsolidated sediment with reed stem phytoliths and some charred organics. Significant compressed phytolith and spherulite cluster with strong parallel orientation represents dung pellet Central slide has unoriented clustered u-shape distribution of dense charred material including significant charcoal. Likely sediment related to burning activity – not defined or restrained by floor/mudbrick or other features

311

TS07-39 Macroscopic Course Fraction – microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Silty clay Moderate Mod 25 Sand Calcite Mod loam

Fine Fraction – microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Clay Enaulic Mod Gradual Marl

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

15% Plains vughs Vertical orientation of planes

Orientation/Distribution Patterns:

Local horizontal parallel orientation Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions: Charcoal – scattered Phytoliths Spherulites – compressed coprolite Shell Plant voids Amorphous organics Mircroartefacts: Rock and Mineral Inclusions: Others:

Basalt/gabbro Courtyard floor fragments

Comments and Summary: Taken from Room 612, western wall of room near burn level. Associated samples TS07-38, TS07-40, TS07- 41, TS07-42

Compressed coprolitic material in upper right quadrant between floor fragment layers, and lower left amorphous organics.

Three horizontal areas containing non-contiguous rough-laid courtyard floor fragments, similar to BC I courtyard surfaces.

Fill between floor “levels” similar to BC I courtyard fill – high charcoal, phytolith and amorphous organic fill

312

TS07-40 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Sandy clay Poor Mod 45% Sand Calcite poor loam

Fine Fraction – microscopic Texture: C/F related Distribution Sorting: Boundary: Notes: Sharp boundary Clay Enaulic Unsorted Sharp between fill and marl mudbrick frags Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

15% Cracks Vughs

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

Phytoliths – lots Charcoal Plant voids shell

Mircroartefacts: Rock and Mineral Inclusions: Others:

Building frags - floors Pottery Basalt/gabbro Clay coating - limited

Comments and Summary: Taken from Room 612, western wall of room near burn level. Associated samples TS07-38, TS07- 39, TS07-41, TS07-42

Mudbrick frags + unoriented rough-laid courtyard floors similar to BC I courtyard and TS07-39.

Loose organic rich fill, lots of phytoliths.

Lower portion of slide contains more sharp-edged building fragments.

313

TS07-41 Macroscopic Course Fraction – microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Sandy clay Poor Poor 40 Sand Calcite Unsorted loam

Fine Fraction – microscopic Texture: C/F related Distribution Sorting: Boundary: Notes: Sharp boundary Clay Enaulic Unsorted Sharp between loose and loam mudbrick frags Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

10% Planes Vughs

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

Charcoal Phytoliths – lots Shell

Mircroartefacts: Rock and Mineral Inclusions: Others:

Pottery frags Basalt/gabbro Mudbrick frags

Comments and Summary:

Taken from Room 612, western wall of room near burn level. Associated samples TS07-38, TS07- 39, TS07-41, TS07-42

Loose, phytolith fill material. Embedded pottery and mudbrick frags.

314

TS07-42 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Sandy clay Poor Mod 60% Sand Calcite Unsorted loam

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes: Sharp boundary with Clay Enaulic Unsorted Diffused anthropogenic Loam Porphyric features Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

<5% Cracks

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions: Charcoal Ash

shell Mircroartefacts: Rock and Mineral Inclusions: Others:

Macro ceramic frags Basalt/gabbro Oven/mudbrick frags Flint frags Quartz

Comments and Summary: Taken from Room 612, western wall of room near burn level. Associated samples TS07-39, TS07- 40, TS07-41, TS07-42

Upper limit – dense charcoal with loose, fine fraction

Fragments of building material in increasingly dense fill material from top to bottom of slide

Bottom consolidated clay/marl fine fraction

315

TS07-43 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Calcite Poor Poor Silty clay 5% Sand Poor Limestone

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Clay Monic Unsorted Sharp Silt Porphyric

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

10% Channels Planes Subrounded

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

Amorphous organic shell

Mircroartefacts: Rock and Mineral Inclusions: Others:

Microfossils Ceramic frags Organic staining Gypsum/anhydrite

Comments and Summary:

Taken from silo 272 section – starting from the top downwards. Associated samples TS07-44, 45, 46, 47, 48, 49, 50

Dense consolidated building fragments in loose, monic matrix of similar material

316

TS07-44 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Calcite Moderate Mod Silty clay 5-10% Sand unsorted Limestone

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes: Limited monic fill Clay Porphyric Unsorted Sharp between dense silt monic sections Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

40% Channels Cracks Generally voids between fragments

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

Shell

Burnt bone

Mircroartefacts: Rock and Mineral Inclusions: Others:

Basalt/gabbro – abundant Flint frags – few Mudbrick frags Gypsum/anhydrite

Comments and Summary:

Taken from silo 272 section – starting from the top downwards. Associated samples TS07-43, 45, 46, 47, 48, 49, 50

Mudbrick fragments, loosely consolidated in monic fill. Channels and voids indicate disturbance, non-contiguous brick.

317

TS07-45 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Moderate Poor Silty clay 20% Sand Calcite Unsorted

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Porphyric Sharp Sharp between frags clay Unsorted Enaulic Blended and fill

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

5% Chambers Vughs

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

Charcoal – scattered, few shell

Mircroartefacts: Rock and Mineral Inclusions: Others: Basalt/gabbro Flint frags Gypsum/anhydrite mudbrick microfossils Comments and Summary:

Taken from silo 272 section – starting from the top downwards. Associated samples TS07-43, 44, 46, 47, 48, 49, 50

Mudbrick fragments in loose fill – 70% slide mudbrick frags.

318

TS07-46 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Very Good Unsort Silty clay 10% Sand Calcite unsorted

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Clay Porphyric Unsorted Sharp

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

25% Planes

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions: Charred wood shell

Mircroartefacts: Rock and Mineral Inclusions: Others:

Flint frags Microfossils mudbrick

Comments and Summary:

No floor level visible, post-occupational fill (sand + brick/mud render frags) Flint – micro flakes, not in situ, likely part of fill rather than knapping area.

Highly fragmented mudbrick

319

TS07-47 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Calcite Poor Poor Silty clay 10% Sand Unsorted igneous

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Porphyric Clay Unsorted Sharp Monic

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

45% Chambers Highly fragmented building material

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

Charred wood Charred buliform reed

Mircroartefacts: Rock and Mineral Inclusions: Others: Basalt/gabbro frequent Pottery – macro 15mm X 8mm Microfossils stuff marl Comments and Summary:

Taken from silo 272 section – starting from the top downwards. Associated samples TS07-43, 44, 45, 46, 48, 49, 50

Mudbrick frag in loose fill.

320

TS07-48 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Good un Silty clay 5% Sand Calcite unsorted

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Clay Porphyric Unsorted Sharp

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

15% Vughs

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions: Charred wood frags Burnt bone

Mircroartefacts: Rock and Mineral Inclusions: Others:

Marl Pottery frag – 3mm Mudbrick Microfossil

Comments and Summary:

Taken from silo 272 section – starting from the top downwards. Associated samples TS07-43, 44, 45, 46, 47, 49, 50

Fragmented mudbrick with some loose fill in upper corner

321

TS07-49 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Good un Silty clay 10 Sand Calcite unsorted

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Clay Porphyric Unsorted Sharp

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

Large vugh in centre, planes radiate 25% Planes Vugh from there

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions: Charred wood Bone

shell

Mircroartefacts: Rock and Mineral Inclusions: Others:

Marl Pottery frags mudbrick Microfossils

Comments and Summary:

Taken from silo 272 section – starting from the top downwards. Associated samples TS07- 43, 44, 45, 46, 47, 48, 50

Highly fragmented mudbrick

322

TS07-50 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Good Un Silty clay 10 Sand Calcite unsorted

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Clay Porphyric Unsorted Sharp

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

45% Vughs

Orientation/Distribution Patterns:

Unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions: Charred wood Burnt bone

shell Mircroartefacts: Rock and Mineral Inclusions: Others:

Marl Mudbrick Microfossils

Comments and Summary:

Taken from silo 272 section – starting from the top downwards. Associated samples TS07-43, 44, 45, 46, 47, 48, 49. To the right of other samples.

Irregular mudbrick fragments, some clay coating

323

GS07-1 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Very good Well Clay <1% Calcite unsorted

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Very Calcareous clay Porphyric Gradual well

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

15% Planes Vughs

Orientation/Distribution Patterns:

Strong bowlike orientation Parallel banded distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

Mircroartefacts: Rock and Mineral Inclusions: Others:

Biotite

Zircon

Comments and Summary:

Taken from 2006 test pit c. 500 m NW of excavation area C. in situ Lisan Formation deposits visible on E and W walls of pit. Samples GS07-1, GS07-2 were both taken from Western wall.

Upper 5th of slide is loose, consolidated clay – broken up Remainder of slide is strongly oriented, waterlaid Lisan clay desposits.

324

GS07-2 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Very good Well Clay <1% Calcite unsorted

Fine Fraction - microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Very Calcareous clay Porphyric Gradual well

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

Large vughs separate clay clusters 45% Vughs Planes with sharp boundaries

Orientation/Distribution Patterns: unoriented Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

Mircroartefacts: Rock and Mineral Inclusions: Others:

biotite

Comments and Summary:

Taken from 2006 test pit c. 500 m NW of excavation area C. in situ Lisan Formation deposits visible on E and W walls of pit. Samples GS07-1, GS07-2 were both taken from Western wall.

Fragments of consolidated calcareous clays separated by vughs – similar to upper portion of GS07-1 Highly disturbed/weathered Lisan clays – caused by drying/exposure between seasons?

325

GS07-8 Macroscopic Course Fraction - microscopic Homogeneity: Sorting: Texture: % Total: Size: Dominant type: Sorting:

Calcite Moderate Poor Silty clay 30% Sand Unsorted Igneous

Fine Fraction – microscopic Texture: C/F related Distribution Sorting: Boundary: Notes:

Calcareous clay Enaulic Poor Diffused

Void Pattern % Total Voids: Dominant Type: Other Voids: Comments:

5% Chambers Vughs

Orientation/Distribution Patterns:

Locally oriented – pockets of parallel deposits Random distribution

Inclusions Plant Evidence: Dung Evidence: Biological Inclusions:

charcoal

Mircroartefacts: Rock and Mineral Inclusions: Others: basalt Plagioclase bearing gabbro

Iron oxides Comments and Summary:

From southern Hill of Tel Tsaf settlement to the left of the excavated wall. No visible artifacts or anthropomorphic material visible in sample location.

Re-deposited sediment from the Jordan Valley catchment area, including anthropogenic input (charcoal).

326

Appendix II: Bulk Sample Grain Size Analysis Data

Mass at each screen (g) Sample <250µm >250µm >500µm >710µm >1mm >2mm Total 70-2 155.99 62.55 33.39 33.11 87.98 117.38 490.4 70-3 93.9 31.7 16.8 17.4 25.3 313.4 498.5 70-4 94.7 20.6 8.2 8.4 15.2 12.8 159.9 70-5 149.9 66.8 27.6 33 62.2 47.3 386.8 70-6 123.95 57.69 30.42 25.7 63.69 45.25 346.7 230-1 181.55 76.32 37.86 31.59 73.63 62.52 463.47 230-2 279 55.81 23.23 18.64 33.44 28.23 438.35 230-3 249.04 83.86 37.07 22.19 38.55 35.94 466.65 230-4 274.2 93.4 31.7 32 40.4 34.1 505.8 230-5 232.1 62.8 22.9 18.9 31.6 34.6 402.9 263-1 454 139.3 49.4 22.6 49.9 20.7 735.9 263-3 295.8 120.6 48 31.1 74.5 33.7 603.7 263-4 257.7 114.3 55.6 53.8 116.5 73.6 671.5 263-5 189.4 83.3 38.5 47.9 93.7 45.5 498.3 662-1 80 11.5 3.8 3.5 4 3.8 106.6 662-2 336.6 83.6 28.3 21.5 33.7 17.3 521 CY-1 223.34 82.83 35.09 26 49.77 88.71 505.74 CY-2 293.32 1010.07 32.18 21.55 35.86 54.64 1447.62 CY-3 279.57 99.59 48.23 37.96 89.34 87.27 641.96 CY-4 325.23 133.75 62.06 49.12 113.28 127.4 810.84 CY-5 319.51 130.32 61.48 50.36 106.84 93.53 762.04 CY-6 230.4 56 22.2 24 38.4 39.4 410.4 CY-7 245.36 92.19 40.06 24.05 42.43 47.93 492.02 CY-9 146.9 69.5 41.8 36.9 70.9 32.8 398.8 CY-10 257.2 110.6 49.9 33.3 60.2 37.2 548.4 CY-11 45.2 10.5 4.7 4.7 9.6 9.7 84.4 CY-12 276.04 98.56 46.44 30.16 70.91 53.46 575.57

327

Appendix II continued

Mass at each screen (g) Sample <250µm >250µm >500µm >710µm >1mm >2mm Total

CY-13 232.03 99.68 48.34 37.49 78.6 66.58 562.72 CY-14 182.5 71.1 31.3 37.2 73.4 56.2 451.7 CY-15 180.5 46.4 21.2 25.5 51 52.9 377.5 CY-16 86.8 17.8 7.2 6.7 12 11.5 142 CY-17 237.8 50.2 19 17.7 35.4 68.3 428.4 CY-18 87.21 22.11 11.79 9.33 17.33 18.76 166.53 RP-1 167.19 64.18 30.04 25.55 46.093 45.09 378.143 RP-2 198.9 92.6 39.5 43.1 53 43.5 470.6 RP-3 186.7 73.6 30.4 33.6 63.9 64.1 452.3 SILO-1 290.8 141.3 53.8 57.1 89.1 72 704.1 SILO-2 193.6 104.6 55.2 44.4 98 59.8 555.6 SILO-4 351.6 180.9 92.1 123 291.2 328.9 1367.7